Irwin M. Arias (Editor), Harvey J. Alter (Editor), James L. Boyer (Editor), David E. Cohen (Editor), David A. Shafritz (Editor), Snorri S. Thorgeirsson (Editor), Allan W. Wolkoff (Editor) - The Liver

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 1126

The Liver

Dedication

This book is dedicated to Win Arias, whose enthusiasm, insight, tools for building bridges between basic and clinical hepatolo-
and scientific rigor have served as an inspiration to several gen- gists as they elucidate together the mysteries of liver function in
erations of investigators, providing the essential foundation and health and disease.
The Liver
Biology and Pathobiology
Sixth Edition

Edited by
Irwin M. Arias MD
Emeritus Senior Scientist, National Institutes of Health, Bethesda, MD, USA
Emeritus Professor of Physiology, Tufts University School of Medicine, Boston, MA, USA
Emeritus Professor of Medicine, Albert Einstein College of Medicine, Bronx, NY, USA

Harvey J. Alter MD, MACP


Distinguished NIH Scientist Emeritus, Department of Transfusion Medicine,
National Institutes of Health, Bethesda, MD, USA

James L. Boyer MD
Ensign Professor of Medicine, Department of Internal Medicine and Liver Center, Yale University School of Medicine,
New Haven, CT, USA

David E. Cohen MD, PhD


Vincent Astor Distinguished Professor of Medicine
Chief, Division of Gastroenterology and Hepatology, Joan & Sanford I. Weill Department of Medicine,
Weill Cornell Medical College
Co‐Director, Center for Advanced Digestive Care,
New York‐Presbyterian Hospital and Weill Cornell Medical Center, New York, NY, USA

David A. Shafritz MD
Professor of Medicine, Cell Biology & Pathology, Albert Einstein College of Medicine
Associate Director, Marion Bessin Liver Research Center, Bronx, NY, USA

Snorri S. Thorgeirsson MD, PhD


Senior Scientist, Laboratory of Human Carcinogenesis, Center for Cancer Research, National Cancer Institute,
National Institutes of Health, Bethesda, MD, USA

Allan W. Wolkoff MD
The Herman Lopata Chair in Liver Disease Research
Professor of Medicine and Anatomy and Structural Biology
Associate Chair of Medicine for Research
Chief, Division of Hepatology
Director, Marion Bessin Liver Research Center
Albert Einstein College of Medicine and Montefiore Medical Center, Bronx, NY, USA
This edition first published 2020
© 2020 John Wiley & Sons Ltd

Edition History
John Wiley & Sons Ltd. (5e, 2009)

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic,
mechanical, photocopying, recording or otherwise, except as permitted by law. Advice on how to obtain permission to reuse material from this title is
available at http://www.wiley.com/go/permissions.

The right of Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz, Snorri S. Thorgeirsson, and Allan W. Wolkoff to be
identified as the authors of the editorial material in this work has been asserted in accordance with law.

Registered Offices
John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA
John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ, UK

Editorial Office
9600 Garsington Road, Oxford, OX4 2DQ, UK

For details of our global editorial offices, customer services, and more information about Wiley products visit us at www.wiley.com.

Wiley also publishes its books in a variety of electronic formats and by print‐on‐demand. Some content that appears in standard print versions of this book
may not be available in other formats.

Limit of Liability/Disclaimer of Warranty


The contents of this work are intended to further general scientific research, understanding, and discussion only and are not intended and should not be
relied upon as recommending or promoting scientific method, diagnosis, or treatment by physicians for any particular patient. In view of ongoing research,
equipment modifications, changes in governmental regulations, and the constant flow of information relating to the use of medicines, equipment, and
devices, the reader is urged to review and evaluate the information provided in the package insert or instructions for each medicine, equipment, or device
for, among other things, any changes in the instructions or indication of usage and for added warnings and precautions. While the publisher and authors
have used their best efforts in preparing this work, they make no representations or warranties with respect to the accuracy or completeness of the contents
of this work and specifically disclaim all warranties, including without limitation any implied warranties of merchantability or fitness for a particular
purpose. No warranty may be created or extended by sales representatives, written sales materials or promotional statements for this work. The fact that an
organization, website, or product is referred to in this work as a citation and/or potential source of further information does not mean that the publisher and
authors endorse the information or services the organization, website, or product may provide or recommendations it may make. This work is sold with the
understanding that the publisher is not engaged in rendering professional services. The advice and strategies contained herein may not be suitable for your
situation. You should consult with a specialist where appropriate. Further, readers should be aware that websites listed in this work may have changed or
disappeared between when this work was written and when it is read. Neither the publisher nor authors shall be liable for any loss of profit or any other
commercial damages, including but not limited to special, incidental, consequential, or other damages.

Library of Congress Cataloging‐in‐Publication Data


Names: Arias, Irwin M., editor.
Title: The liver : biology and pathobiology / edited by Irwin M. Arias [and six others].
Other titles: Liver (Arias)
Description: Sixth edition. | Hoboken, NJ : Wiley-Blackwell, 2020. | Includes bibliographical references and index.
Identifiers: LCCN 2019024961 (print) | LCCN 2019024962 (ebook) | ISBN 9781119436829 (hardback) |
  ISBN 9781119436836 (epub) | ISBN 9781119436843 (adobe pdf)
Subjects: MESH: Liver–physiopathology
Classification: LCC QP185 (print) | LCC QP185 (ebook) | NLM WI 702 | DDC 612.3/52–dc23
LC record available at https://lccn.loc.gov/2019024961
LC ebook record available at https://lccn.loc.gov/2019024962

Cover images: main image and background image courtesy of Sandor Paku; top left image courtesy of Scott Friedman;
top middle images courtesy of Sandor Paku; top right image courtesy of Peter Kim
Cover design by Wiley

Set in 9.5/11.5pt Times LT Std by SPi Global, Pondicherry, India

10 9 8 7 6 5 4 3 2 1

ffirs.indd 4 11/30/2019 9:33:25 AM


Contents

List of Contributors x

Preface xx

Acknowledgments xxi

PART ONE:  INTRODUCTION 1

  1 Organizational Principles of the Liver 3


Peter Nagy, Snorri S. Thorgeirsson, and Joe W. Grisham

  2 Embryonic Development of the Liver 14


Kenneth S. Zaret, Roque Bort, and Stephen A. Duncan

PART TWO:  THE CELLS 23

SECTION A:  CELL BIOLOGY OF THE LIVER 25

  3 Cytoskeletal Motors: Structure and Function in Hepatocytes 27


Mukesh Kumar, Arnab Gupta, and Roop Mallik

  4 Hepatocyte Surface Polarity 36


Anne Müsch and Irwin M. Arias

  5 Primary Cilia 50
Carolyn M. Ott

  6 Endocytosis in Liver Function and Pathology 62


Micah B. Schott, Barbara Schroeder, and Mark A. McNiven

  7 The Hepatocellular Secretory Pathway 75


Catherine L. Jackson and Mark A. McNiven

  8 Mitochondrial Function, Dynamics, and Quality Control 86


Marc Liesa, Ilan Benador, Nathanael Miller, and Orian S. Shirihai

  9 Nuclear Pore Complex 94


Michelle A. Veronin and Joseph S. Glavy

10 Protein Maturation and Processing at the Endoplasmic Reticulum 108


Ramanujan S. Hegde
vi Contents

11 Protein Degradation and the Lysosomal System 122


Susmita Kaushik and Ana Maria Cuervo

12 Peroxisome Assembly, Degradation, and Disease 137


Rong Hua and Peter K. Kim

13 Organelle–Organelle Contacts: Origins and Functions 151


Uri Manor

14 Gap and Tight Junctions in Liver: Structure, Function, and Pathology 160


John W. Murray and David C. Spray

15 Ribosome Biogenesis and its Role in Cell Growth and Proliferation in the Liver 174
Katherine I. Farley‐Barnes and Susan J. Baserga

16 miRNAs and Hepatocellular Carcinoma 183


Yusuke Yamamoto, Isaku Kohama, and Takahiro Ochiya

17 Hepatocyte Apoptosis: Mechanisms and Relevance in Liver Diseases 195


Harmeet Malhi and Gregory J. Gores

SECTION B:  THE HEPATOCYTE 207

18 Copper Metabolism and the Liver 209


Cynthia Abou Zeid, Ling Yi, and Stephen G. Kaler

19 The Central Role of the Liver in Iron Storage and Regulation of Systemic Iron Homeostasis 215
Tracey A. Rouault, Victor R. Gordeuk, and Gregory J. Anderson

20 Disorders of Bilirubin Metabolism 229


Namita Roy Chowdhury, Yanfeng Li, and Jayanta Roy Chowdhury

21 Hepatic Lipid Droplets in Liver Function and Disease 245


Douglas G. Mashek, Wenqi Cui, Linshan Shang, and Charles P. Najt

22 Lipoprotein Metabolism and Cholesterol Balance 255


Mariana Acuña‐Aravena and David E. Cohen

SECTION C:  TRANSPORTERS, BILE ACIDS, AND CHOLESTASIS 269

23 Bile Acid Metabolism in Health and Disease: An Update 271


Tiangang Li and John Y.L. Chiang

24 TGR5 (GPBAR1) in the Liver 286


Verena Keitel, Christoph G.W. Gertzen, Lina Spomer, Holger Gohlke, and Dieter Häussinger

25 Bile Acids as Signaling Molecules 299


Thierry Claudel and Michael Trauner

26 Hepatic Adenosine Triphosphate‐Binding Cassette Transport Proteins and Their Role in Physiology 313
Peter L.M. Jansen

27 Basolateral Plasma Membrane Organic Anion Transporters 327


M. Sawkat Anwer and Allan W. Wolkoff

28 Hepatic Nuclear Receptors 337


Raymond E. Soccio

29 Molecular Cholestasis 351


Paul Gissen and Richard J. Thompson

30 Pathophysiologic Basis for Alternative Therapies for Cholestasis 364


Claudia D. Fuchs, Emina Halilbasic, and Michael Trauner
Contents vii

31 Adaptive Regulation of Hepatocyte Transporters in Cholestasis 378


James L. Boyer

SECTION D:  NON‐HEPATOCYTE CELLS 391

32 Cholangiocyte Biology and Pathobiology 393


Massimiliano Cadamuro, Romina Fiorotto, and Mario Strazzabosco

33 Polycystic Liver Diseases: Genetics, Mechanisms, and Therapies 408


Tatyana Masyuk, Anatoliy Masyuk, and Nicholas LaRusso

34 The Liver Sinusoidal Endothelial Cell: Basic Biology and Pathobiology 422


Karen K. Sørensen and Bård Smedsrød

35 Fenestrations in the Liver Sinusoidal Endothelial Cell 435


Victoria C. Cogger, Nicholas J. Hunt, and David G. Le Couteur

36 Stellate Cells and Fibrosis 444


Youngmin A. Lee and Scott L. Friedman

PART THREE:  FUNCTIONS OF THE LIVER 455

SECTION A:  METABOLIC FUNCTIONS 457

37 Non‐alcoholic Fatty Liver Disease and Insulin Resistance 459


Max C. Petersen, Varman T. Samuel, Kitt Falk Petersen, and Gerald I. Shulman

38 AMPK: Central Regulator of Glucose and Lipid Metabolism and Target of Type 2 Diabetes Therapeutics 472
Daniel Garcia, Maria M. Mihaylova, and Reuben J. Shaw

39 Insulin‐Mediated PI3K and AKT Signaling 485


Hyokjoon Kwon and Jeffrey E. Pessin

40 Ca2+ Signaling in the Liver 496


Mateus T. Guerra, M. Fatima Leite, and Michael H. Nathanson

41 Clinical Genomics of NAFLD 509


Frank Lammert

SECTION B:  LIVER GROWTH AND REGENERATION 521

42 Stem Cell‐Fueled Maturational Lineages in Hepatic and Pancreatic Organogenesis 523


Wencheng Zhang, Amanda Allen, Eliane Wauthier, Xianwen Yi, Homayoun Hani, Praveen Sethupathy,
David Gerber, Vincenzo Cardinale, Guido Carpino, Juan Dominguez‐Bendala, Giacomo Lanzoni,
Domenico Alvaro, Eugenio Gaudio, and Lola Reid

43 Developmental Morphogens and Adult Liver Repair 539


Mariana Verdelho Machado and Anna Mae Diehl

44 Liver Repopulation by Cell Transplantation and the Role of Stem Cells in Liver Biology 550
David A. Shafritz and Markus Grompe

45 Liver Regeneration 566


George K. Michalopoulos

46 β‐Catenin Signaling 585


Satdarshan P.S. Monga

47 Polyploidy in Liver Function, Mitochondrial Metabolism, and Cancer 603


Evan R. Delgado, Elizabeth C. Stahl, Nairita Roy, Patrick D. Wilkinson, and Andrew W. Duncan
viii Contents

PART FOUR:  PATHOBIOLOGY OF LIVER DISEASE 615

48 Hepatic Encephalopathy 617


Roger F. Butterworth

49 The Kidney in Liver Disease 630


Moshe Levi, Shogo Takahashi, Xiaoxin X. Wang, and Marilyn E. Levi

50 α1‐Antitrypsin Deficiency 645


David A. Rudnick and David H. Perlmutter

51 Pathophysiology of Portal Hypertension 659


Yasuko Iwakiri and Roberto J. Groszmann

52 Non‐alcoholic Fatty Liver Disease: Mechanisms and Treatment 670


Yaron Rotman and Devika Kapuria

53 Alcoholic Liver Disease 682


Bin Gao, Xiaogang Xiang, Lorenzo Leggio, and George F. Koob

54 Drug‐Induced Liver Injury 701


Lily Dara and Neil Kaplowitz

55 Oxidative Stress and Inflammation in the Liver 714


John J. Lemasters and Hartmut Jaeschke

56 The Role of Bile Acid‐Mediated Inflammation in Cholestatic Liver Injury 728


Shi‐Ying Cai, Man Li, and James L. Boyer

57 Toll‐like Receptors in Liver Disease 737


So Yeon Kim and Ekihiro Seki

PART FIVE:  LIVER CANCER 747

58 Experimental Models of Liver Cancer: Genomic Assessment of Experimental Models 749


Sun Young Yim, Jae‐Jun Shim, Bo Hwa Sohn, and Ju‐Seog Lee

59 Epidemiology of Hepatocellular Carcinoma 758


Hashem B. El‐Serag

60 Mutations and Genomic Alterations in Liver Cancer 773


Jessica Zucman‐Rossi and Jean‐Charles Nault

61 Treatment of Liver Cancer 782


Tim F. Greten

PART SIX:  HEPATITIS 793

62 Molecular Biology of Hepatitis Viruses 795


Christoph Seeger, William S. Mason, and Michael M.C. Lai

63 Immune Mechanisms of Viral Clearance and Disease Pathogenesis During Viral Hepatitis 821


Carlo Ferrari, Valeria Barili, Stefania Varchetta, and Mario U. Mondelli

64 Clinical Implications of the Molecular Biology of Hepatitis B Virus 851


Timothy M. Block, Ju‐Tao Guo, and W. Thomas London

65 Viral Escape Mechanisms in Hepatitis C and the Clinical Consequences of Persistent Infection 868
Marc G. Ghany, Christopher M. Walker, and Patrizia Farci

66 Tracking Hepatitis C Virus Interactions with the Hepatic Lipid Metabolism: A Hitchhiker’s


Guide to Solve Remaining Translational Research Challenges in Hepatitis C 889
Gabrielle Vieyres and Thomas Pietschmann
Contents ix

67 Nucleoside Antiviral Agents for HCV: What’s Left to Do? 906


Franck Amblard, Seema Mengshetti, Junxing Shi, Sijia Tao, Leda Bassit, and Raymond F. Schinazi

68 Hepatitis E Virus: An Emerging Zoonotic Virus Causing Acute and Chronic Liver Disease 915
Xiang‐Jin Meng

69 Biological Principles and Clinical Issues Underlying Liver Transplantation for Viral‐Induced
End‐Stage Liver Disease in the Era of Highly Effective Direct‐Acting Antiviral Agents 926
Michael S. Kriss, James R. Burton, Jr., and Hugo R. Rosen

70 Time for the Elimination of Hepatitis C Virus as a Global Health Threat 935


John W. Ward, Alan R. Hinman, and Harvey J. Alter

PART SEVEN:  HORIZONS 953

71 Genome Editing by Targeted Nucleases and the CRISPR/Cas Revolution 955


Shawn M. Burgess

72 Imaging Cellular Proteins and Structures: Smaller, Brighter, and Faster 965


Aubrey V. Weigel and Erik Lee Snapp

73 Liver‐Directed Gene Therapy 979


Patrik Asp, Chandan Guha, Namita Roy Chowdhury, and Jayanta Roy Chowdhury

74 Telomeres and Telomerase in Liver Generation and Cirrhosis 992


Sonja C. Schätzlein and K. Lenhard Rudolph

75 Toxins and Biliary Atresia 1000


Michael Pack and Rebecca G. Wells

76 The Dual Role of ABC Transporters in Drug Metabolism and Resistance to Chemotherapy 1007
Jean‐Pierre Gillet, Marielle Boonen, Michel Jadot, and Michael M. Gottesman

77 Stem Cell‐Derived Liver Cells: From Model System to Therapy 1015


Helmuth Gehart and Hans Clevers

78 Extracellular Vesicles and Exosomes: Biology and Pathobiology 1022


Gyongyi Szabo and Fatemeh Momen‐Heravi

79 Integrated Technologies for Liver Tissue Engineering 1028


Tiffany N. Vo, Amanda X. Chen, Quinton B. Smith, Arnav Chhabra and Sangeeta N. Bhatia

80 Pluripotent Stem Cells and Reprogramming: Promise for Therapy 1036


James A. Heslop and Stephen A. Duncan

81 Chromatin Regulation and Transcription Factor Cooperation in Liver Cells 1043


Ido Goldstein

82 Drug Interactions in the Liver 1050


Guruprasad P. Aithal and Gerd A. Kullak‐Ublick

83 Metabolic Regulation of Hepatic Growth 1058


Wolfram Goessling

84 The Gut Microbiome and Liver Disease 1062


Lexing Yu, Jasmohan S. Bajaj, and Robert F. Schwabe

85 Lineage Tracing: Efficient Tools to Determine the Fate of Hepatic Cells in Health and Disease 1069
Frédéric Lemaigre

86 The Hepatocyte as a Household for Plasmodium Parasites 1075


Vanessa Zuzarte‐Luis and Maria M. Mota

Index 1081
List of Contributors

Cynthia Abou Zeid M. Sawkat Anwer


Section on Translational Neuroscience, Molecular Medicine Tufts Clinical and Translational Science Institute
Branch Tufts University Cummings School of Veterinary Medicine
Intramural Research Program, National Institutes of Health, Department of Biomedical Sciences
Bethesda, MD, USA North Grafton, MA, USA
Mariana Acuña‐Aravena Irwin M. Arias
Division of Gastroenterology and Hepatology, Joan & National Institutes of Health
Sanford I. Weill Department of Medicine Bethesda, MD, USA
Weill Cornell Medical College
New York, NY, USA Patrik Asp
Marion Bessin Liver Research Center
Guruprasad P. Aithal Department of Surgery, Albert Einstein College of Medicine
Nottingham Digestive Diseases Centre, School of Medicine, Bronx, NY, USA
University of Nottingham, Nottingham, UK;
NIHR Nottingham Biomedical Research Centre, Nottingham Jasmohan S. Bajaj
University Hospitals NHS Trust and the University of Virginia Commonwealth University and McGuire
Nottingham, Nottingham, UK VA Medical Center
Richmond, VA, USA
Amanda Allen
Department of Cell Biology and Physiology Valeria Barili
University of North Carolina School of Medicine Unit of Infectious Diseases and Hepatology
Chapel Hill, NC, USA Department of Medicine and Surgery, University of Parma
Parma, Italy
Harvey J. Alter
Department of Transfusion Medicine, Clinical Center Susan J. Baserga
National Institutes of Health Department of Molecular Biophysics & Biochemistry
Bethesda, MD, USA Department of Genetics
Department of Therapeutic Radiology, Yale University School
Domenico Alvaro
of Medicine
Department of Medico‐Surgical Sciences and Biotechnologies,
New Haven, CT, USA
Sapienza University of Rome, Latina, Italy
Department of Medicine and Medical Specialties Leda Bassit
Sapienza University of Rome Laboratory of Biochemical Pharmacology, Emory University
Rome, Italy School of Medicine
Atlanta, GA, USA
Franck Amblard
Laboratory of Biochemical Pharmacology, Emory University Ilan Benador
School of Medicine Department of Medicine, Division of Endocrinology and
Atlanta, GA, USA Department of Molecular and Medical Pharmacology
David Geffen School of Medicine at UCLA
Gregory J. Anderson
Los Angeles, CA, USA
QIMR Berghofer Medical Research Institute
Brisbane, Queensland, Australia
List of Contributors xi

Sangeeta N. Bhatia Guido Carpino


Institute for Medical Engineering and Science, Massachusetts Department of Movement, Human and Health Sciences,
Institute of Technology, Cambridge, MA; Division of Health Sciences
Harvard‐MIT Department of Health Sciences and Technology, University of Rome “Foro Italico”;
Institute for Medical Engineering and Science, Massachusetts Department of Anatomical, Histological, Forensic Medicine
Institute of Technology, Boston, MA; and Orthopedics Sciences
Department of Electrical Engineering and Computer Science, Sapienza University of Rome
Massachusetts Institute of Technology, Cambridge, MA; Rome, Italy
David H. Koch Institute for Integrative Cancer Research,
Amanda X. Chen
Massachusetts Institute of Technology, Cambridge, MA;
Department of Biological Engineering, Massachusetts
Howard Hughes Medical Institute, Chevy Chase, MD, USA
Institute of Technology
Timothy M. Block Cambridge, MA, USA
Baruch S. Blumberg Institute
Arnav Chhabra
Doylestown, PA, USA
Harvard‐MIT Department of Health Sciences and Technology,
Marielle Boonen Institute for Medical Engineering and Science, Massachusetts
Laboratory of Intracellular Trafficking Biology, URPhyM, Institute of Technology
NARILIS Boston, MA, USA
Faculty of Medicine, University of Namur
John Y.L. Chiang
Namur, Belgium
Department of Integrative Medical Sciences
Roque Bort Northeast Ohio Medical University
Instituto de Investigación Sanitaria La Fe (IIS La Fe) Rootstown, OH, USA
Unidad de Hepatología experimental
València, Spain Thierry Claudel
Hans Popper Laboratory of Molecular Hepatology, Division
James L. Boyer of Gastroenterology and Hepatology Department of Internal
Department of Internal Medicine and Liver Center, Medicine III
Yale University School of Medicine Medical University of Vienna
New Haven, CT, USA Vienna, Austria
Shawn M. Burgess Hans Clevers
National Human Genome Research Institute Oncode Institute, Hubrecht Institute, Royal Netherlands
National Institutes of Health Academy of Arts and Sciences (KNAW) and University
Bethesda, MD, USA Medical Centre (UMC) Utrecht;
Princess Máxima Centre for Paediatric Oncology
James R. Burton, Jr.
Utrecht, The Netherlands
University of Colorado School of Medicine
Aurora, CO, USA Victoria C. Cogger
Roger F. Butterworth Centre for Education and Research on Ageing
Department of Medicine University of Sydney and Concord RG Hospital
University of Montreal Sydney, NSW, Australia
Montreal, QC, Canada David E. Cohen
Massimiliano Cadamuro Division of Gastroenterology and Hepatology,
Department of Molecular Medicine Joan & Sanford I. Weill Department of Medicine
University of Padua, Padova, Italy; Weill Cornell Medical College
International Center for Digestive Health (ICDH) New York, NY, USA
University of Milan‐Bicocca, Monza, Italy
Ana Maria Cuervo
Shi‐Ying Cai Department of Developmental and Molecular Biology
The Liver Center, Yale University School of Medicine Institute for Aging Research, Albert Einstein College of
New Haven, CT, USA Medicine
Bronx, NY, USA
Vincenzo Cardinale
Department of Medico‐Surgical Sciences and Wenqi Cui
Biotechnologies Department of Biochemistry, Molecular Biology and Biophysics
Sapienza University of Rome University of Minnesota
Latina, Italy Minneapolis, MN, USA
xii List of Contributors

Lily Dara Liver Center and Section of Digestive Diseases, Department


Research Center for Liver Disease, Department of Medicine, of Internal Medicine, Section of Digestive Diseases, Yale
Division of Gastrointestinal and Liver Diseases University School of Medicine
Keck School of Medicine, University of Southern California New Haven, CT, USA
Los Angeles, CA, USA
Scott L. Friedman
Evan R. Delgado Division of Liver Diseases
Department of Pathology Icahn School of Medicine at Mount Sinai
McGowan Institute for Regenerative Medicine, Pittsburgh New York, NY, USA
Liver Research Center
Claudia D. Fuchs
University of Pittsburgh
Hans Popper Laboratory of Molecular Hepatology, Division
Pittsburgh, PA, USA
of Gastroenterology and Hepatology Department of Internal
Anna Mae Diehl Medicine III
School of Medicine, Duke University Medical University of Vienna
Durham, NC, USA Vienna, Austria
Juan Dominguez‐Bendala Bin Gao
Department of Medicine and Medical Specialties Laboratory of Liver Diseases, National Institute on Alcohol
Sapienza University of Rome Abuse and Alcoholism
Rome, Italy; National Institutes of Health
Diabetes Research Institute, Miller School of Medicine Bethesda, MD, USA
University of Miami
Daniel Garcia
Miami, FL, USA
Molecular and Cell Biology Laboratory
Andrew W. Duncan The Salk Institute for Biological Studies
Department of Pathology, McGowan Institute for Regenerative La Jolla, CA, USA
Medicine, Pittsburgh Liver Research Center, University of
Eugenio Gaudio
Pittsburgh
Department of Anatomical, Histological, Forensic Medicine
Pittsburgh, PA, USA
and Orthopedics Sciences
Stephen A. Duncan Sapienza University of Rome
Department of Regenerative Medicine and Cell Biology Rome, Italy
Medical University of South Carolina
Helmuth Gehart
Charleston, SC, USA
Oncode Institute, Hubrecht Institute, Royal Netherlands
Hashem B. El‐Serag Academy of Arts and Sciences (KNAW) and University
Department of Medicine Medical Centre (UMC) Utrecht
Baylor College of Medicine and Michael E. DeBakey Veterans Utrecht, The Netherlands
Affairs Medical Center
David Gerber
Houston, TX, USA
Department of Surgery
Patrizia Farci University of North Carolina School of Medicine
Hepatic Pathogenesis Section, Laboratory of Infectious Chapel Hill, NC, USA
Diseases, National Institute of Allergy and Infectious Diseases,
Christoph G.W. Gertzen
National Institutes of Health
Institute of Pharmaceutical and Medicinal Chemistry,
Bethesda, MD, USA
Heinrich‐Heine University Düsseldorf, Düsseldorf, Germany
Katherine I. Farley‐Barnes
Marc G. Ghany
Department of Molecular Biophysics & Biochemistry
Liver Diseases Branch, National Institute of Diabetes and
Yale University School of Medicine
Digestive and Kidney Diseases
New Haven, CT, USA
National Institutes of Health
Carlo Ferrari Bethesda, MD, USA
Unit of Infectious Diseases and Hepatology
Jean‐Pierre Gillet
Department of Medicine and Surgery, University of Parma
Laboratory of Molecular Cancer Biology, URPhyM, NARILIS
Parma, Italy
Faculty of Medicine, University of Namur
Romina Fiorotto Namur, Belgium
International Center for Digestive Health (ICDH), University
Paul Gissen
of Milan‐Bicocca
UCL Great Ormond Street Institute of Child Health and Great
Monza, Italy;
Ormond Street Hospital for Children, London, UK
List of Contributors xiii

Joseph S. Glavy Chandan Guha


University of Texas at Tyler Marion Bessin Liver Research Center, Albert Einstein College
Fisch College of Pharmacy of Medicine, Bronx, NY;
Tyler, TX, USA Departments of Radiation Oncology and Pathology, Albert
Einstein College of Medicine
Wolfram Goessling
Bronx, NY, USA
Division of Gastroenterology
Massachusetts General Hospital Ju‐Tao Guo
Harvard‐MIT Division of Health Sciences and Technology Baruch S. Blumberg Institute
Harvard Medical School Doylestown, PA, USA
Boston, MA, USA
Arnab Gupta
Holger Gohlke Department of Biological Sciences, Indian Institute of Science
Institute of Pharmaceutical and Medicinal Chemistry, Education and Research Kolkata, Mohanpur, India
Heinrich‐Heine University Düsseldorf, Düsseldorf, Germany
Emina Halilbasic
Ido Goldstein Hans Popper Laboratory of Molecular Hepatology, Division
Institute of Biochemistry, Food Science and Nutrition of Gastroenterology and Hepatology Department of Internal
The Robert H. Smith Faculty of Agriculture, Food and Medicine III
Environment Medical University of Vienna
The Hebrew University of Jerusalem Vienna, Austria
Rehovot, Israel
Homayoun Hani
Victor R. Gordeuk Department of Cell Biology and Physiology
University of Illinois at Chicago University of North Carolina School of Medicine
Chicago, IL, USA Chapel Hill, NC, USA
Gregory J. Gores
Dieter Häussinger
College of Medicine, Division of Gastroenterology
Clinic for Gastroenterology, Hepatology and Infectious
and Hepatology
Diseases, University Hospital Düsseldorf Medical Faculty at
Mayo Clinic
Heinrich‐Heine‐University
Rochester, MN, USA
Düsseldorf, Germany
Michael M. Gottesman
Ramanujan S. Hegde
Laboratory of Cell Biology, Center for Cancer Research
MRC Laboratory of Molecular Biology
National Cancer Institute, National Institutes of Health
Cambridge, UK
Bethesda, MD, USA
Tim F. Greten James A. Heslop
Thoracic and Gastrointestinal Malignancies Branch Department of Regenerative Medicine and Cell Biology
Center for Cancer Research, National Cancer Institute Medical University of South Carolina
Bethesda, MD, USA Charleston, SC, USA

Joe W. Grisham Alan R. Hinman


Department of Pathology and Laboratory Medicine Task Force for Global Health
University of North Carolina Decatur, GA, USA
Chapel Hill, NC, USA Rong Hua
Markus Grompe Cell Biology Program, Hospital for Sick Children
Papé Pediatric Research Institute Toronto, ON, Canada
Oregon Health Sciences University
Nicholas J. Hunt
Portland, OR, USA
Centre for Education and Research on Ageing
Roberto J. Groszman University of Sydney and Concord RG Hospital
Section of Digestive Diseases, Yale School of Medicine Sydney, NSW, Australia
New Haven, CT, USA
Yasuko Iwakiri
Mateus T. Guerra Section of Digestive Diseases
Departments of Medicine and Cell Biology Yale University School of Medicine
Yale University School of Medicine New Haven, CT, USA
New Haven, CT, USA
xiv List of Contributors

Catherine L. Jackson Isaku Kohama


Institut Jacques Monod, UMR7592 CNRS Université Paris‐ Division of Molecular and Cellular Medicine, National Cancer
Diderot, Sorbonne Paris Cité Center Research Institute
Paris, France Tsukiji, Tokyo, Japan
Michel Jadot George F. Koob
Laboratory of Physiological Chemistry, URPhyM, NARILIS, National Institute on Alcohol Abuse and Alcoholism and
Faculty of Medicine National Institute on Drug Abuse
University of Namur, Belgium National Institutes of Health
Bethesda, MD, USA
Hartmut Jaeschke
Department of Pharmacology, Toxicology and Therapeutics Michael S. Kriss
University of Kansas Medical Center Division of Gastroenterology & Hepatology
Kansas City, KS, USA University of Colorado School of Medicine
Peter L.M. Jansen Aurora, CO, USA
Department of Hepatology and Gastroenterology Gerd A. Kullak‐Ublick
Academic Medical Center University Hospital Zurich and University of Zurich
Amsterdam, The Netherlands; Zurich, Zurich, Switzerland;
LiSyM research network Mechanistic Safety, Chief Medical Office and Patient Safety,
University of Freiburg Novartis Global Drug Development
Freiburg, Germany Basel, Switzerland
Stephen G. Kaler
Mukesh Kumar
Section on Translational Neuroscience, Molecular
Department of Biological Sciences, Tata Institute of
Medicine Branch
Fundamental Research, Navy Nagar, Colaba, Mumbai, India
Intramural Research Program, National Institutes of Health
Bethesda, MD, USA Hyokjoon Kwon
Neil Kaplowitz Department of Medicine, Division of Endocrinology,
Research Center for Liver Disease, Department of Medicine, Metabolism and Nutrition
Division of Gastrointestinal and Liver Diseases Rutgers‐Robert Wood Johnson Medical School
Keck School of Medicine, University of Southern California New Brunswick, NJ, USA
Los Angeles, CA, USA Michael M.C. Lai
Devika Kapuria China Medical University and China Medical
Liver Energy and Metabolism Section, Liver Diseases Branch, University Hospital
National Institute of Diabetes and Digestive and Kidney Taichung, Taiwan
Diseases
Frank Lammert
National Institutes of Health
Department of Medicine II
Bethesda, MD, USA
Saarland University Medical Center
Susmita Kaushik Homburg, Germany
Department of Developmental and Molecular Biology
Institute for Aging Research, Albert Einstein College of Giacomo Lanzoni
Medicine Diabetes Research Institute, Miller School of Medicine
Bronx, NY, USA University of Miami
Miami, FL, USA
Verena Keitel
Clinic for Gastroenterology, Hepatology and Infectious Nicholas LaRusso
Diseases, University Hospital Düsseldorf Medical Faculty at Division of Gastroenterology and Hepatology
Heinrich‐Heine‐University Mayo Clinic College of Medicine
Düsseldorf, Germany Rochester, MN, USA

Peter K. Kim David G. Le Couteur


Cell Biology Program, The Hospital for Sick Children; Centre for Education and Research on Ageing
Department of Biochemistry, University of Toronto University of Sydney and Concord RG Hospital
Toronto, ON, Canada Sydney, NSW, Australia

So Yeon Kim Ju‐Seog Lee


Division of Digestive and Liver Diseases, Department of Department of Systems Biology
Medicine The University of Texas M.D. Anderson Cancer Center
Cedars‐Sinai Medical Center Houston, TX, USA
Los Angeles, CA, USA
List of Contributors xv

Youngmin A. Lee Mariana Verdelho Machado


Department of Surgery, Vanderbilt University Medical Center Gastroenterology and Hepatology Department
Nashville, TN, USA Hospital de Santa Maria
CHLN, Lisbon;
Lorenzo Leggio Faculty of Medicine
Section on Clinical Psychoneuroendocrinology and Lisbon University
Neuropsychopharmacology, National Institute on Alcohol Lisbon, Portugal
Abuse and Alcoholism and National Institute on Drug Abuse
National Institutes of Health Harmeet Malhi
Bethesda, MD, USA College of Medicine, Division of Gastroenterology and Hepatology
Mayo Clinic
M. Fatima Leite Rochester, MN, USA
Department of Physiology and Biophysics
UFMG Roop Mallik
Belo Horizonte, Brazil Department of Biological Sciences, Tata Institute of
Fundamental Research, Navy Nagar, Colaba, Mumbai, India
Frédéric Lemaigre
Université catholique de Louvain, de Duve Institute Uri Manor
Brussels, Belgium Waitt Advanced Biophotonics Center, Salk Institute for
Biological Studies
John J. Lemasters La Jolla, CA, USA
Departments of Drug Discovery and Biomedical Sciences and
Biochemistry and Molecular Biology Douglas G. Mashek
Medical University of South Carolina Department of Biochemistry, Molecular Biology and
Charleston, SC, USA Biophysics and Department of Medicine, Division of Diabetes,
Endocrinology and Metabolism
Marilyn E. Levi University of Minnesota
Department of Medicine, Division of Infectious Diseases, Minneapolis, MN, USA
University of Colorado
Aurora, CO, USA William S. Mason
Fox Chase Cancer Center
Moshe Levi Philadelphia, PA, USA
Department of Biochemistry and Molecular and Cellular
Biology Anatoliy Masyuk
Georgetown University Division of Gastroenterology and Hepatology
Washington, DC, USA Mayo Clinic College of Medicine
Rochester, MN, USA
W. Thomas London (deceased)
Formerly, Fox Chase Cancer Center Tatyana Masyuk
Philadelphia, PA, USA Division of Gastroenterology and Hepatology
Mayo Clinic College of Medicine
Man Li Rochester, MN, USA
The Liver Center, Yale University School of Medicine
New Haven, CT, USA Mark A. McNiven
Department of Biochemistry and Molecular Biology, Division
Tiangang Li of Gastroenterology and Hepatology, Mayo Clinic
Department of Pharmacology, Toxicology and Therapeutics Rochester, MN, USA
University of Kansas Medical Center
Kansas City, KS, USA Xiang‐Jin Meng
Department of Biomedical Sciences and Pathobiology
Yanfeng Li Virginia Polytechnic Institute and State University
Departments of Medicine and Genetics and Marion Bessin Blacksburg, VA, USA
Liver Research Center
Albert Einstein College of Medicine, Bronx, NY, USA Seema Mengshetti
Laboratory of Biochemical Pharmacology, Emory University
Marc Liesa School of Medicine
Department of Medicine, Division of Endocrinology and Atlanta, GA, USA
Department of Molecular and Medical Pharmacology
David Geffen School of Medicine at UCLA George K. Michalopoulos
Los Angeles, CA, USA Department of Pathology
University of Pittsburgh School of Medicine
Pittsburgh, PA, USA
xvi List of Contributors

Maria M. Mihaylova Liver Unit, Hôpital Jean Verdier, Hôpitaux Universitaires


Molecular and Cell Biology Laboratory Paris‐Seine‐Saint‐Denis, Assistance‐Publique Hôpitaux de
The Salk Institute for Biological Studies Paris, Bondy, France
La Jolla, CA, USA Takahiro Ochiya
Nathanael Miller Division of Molecular and Cellular Medicine, National Cancer
Department of Medicine, Division of Endocrinology and Center Research Institute
Department of Molecular and Medical Pharmacology Tsukiji, Tokyo
David Geffen School of Medicine at UCLA Institute of Medical Science, Tokyo Medical University
Los Angeles, CA, USA Shinjuku, Tokyo, Japan
Janelia Research Campus, Ashburn, VA, USA
Fatemeh Momen‐Heravi
Department of Medicine Carolyn M. Ott
University of Massachusetts Medical School Janelia Research Campus
Worcester, MA, USA Ashburn, VA, USA

Mario U. Mondelli Michael Pack


Division of Infectious Diseases and Immunology, Fondazione Departments of Medicine (GI Division) and Cell and
IRCCS Policlinico S.Matteo, Pavia; Developmental Biology
Department of Internal Medicine and Therapeutics Perelman School of Medicine, University of Pennsylvania
University of Pavia, Pavia, Italy Philadelphia, PA, USA

Satdarshan P.S. Monga David H. Perlmutter


Pittsburgh Liver Research Center Department of Pediatrics, Division of Gastroenterology,
University of Pittsburgh, School of Medicine and UPMC Hepatology, and Nutrition and Department of
Pittsburgh, PA, USA Developmental Biology, Washington University School
of Medicine in St. Louis, St. Louis Children’s Hospital,
Maria M. Mota St. Louis, MO, USA
Instituto de Medicina Molecular João Lobo Antunes,
Jeffrey E. Pessin
Faculty of Medicine, University of Lisbon, Lisbon,
Albert Einstein‐Mount Sinai Diabetes Research Center and
Portugal
the Fleischer Institute for Diabetes and Metabolism, Albert
Anne Müsch Einstein College of Medicine, Bronx, NY;
Department of Developmental & Molecular Biology Department of Medicine and Department of Molecular
Albert Einstein College of Medicine Pharmacology, Albert Einstein College of Medicine,
Bronx, NY, USA Bronx, NY, USA
John W. Murray Kitt Falk Petersen
Marion Bessin Liver Research Center Department of Internal Medicine, Section of Endocrinology
Albert Einstein College of Medicine; Yale University School of Medicine
Department of Anatomy and Structural Biology, Albert New Haven, CT, USA
Einstein College of Medicine Max C. Petersen
Bronx, NY, USA Department of Internal Medicine, Section of Endocrinology,
Peter Nagy and Department of Molecular and Cellular Physiology
First Department of Pathology and Experimental Cancer Yale University School of Medicine
Research, Semmelweis University, Budapest, Hungary New Haven, CT, USA
Thomas Pietschmann
Charles P. Najt Institute of Experimental Virology, TWINCORE,
Department of Biochemistry, Molecular Biology and Centre for Experimental and Clinical Infection Research,
Biophysics Hannover;
University of Minnesota German Centre for Infection Research (DZIF)
Minneapolis, MN, USA Braunschweig, Germany
Michael H. Nathanson Lola Reid
Departments of Medicine and Cell Biology Department of Cell Biology and Physiology
Yale University School of Medicine Program in Molecular Biology and Biotechnology
New Haven, CT, USA University of North Carolina School of Medicine
Chapel Hill, NC, USA
Jean‐Charles Nault
Centre de Recherche des Cordeliers, Sorbonne Université, Hugo R. Rosen
Inserm, Université de Paris, Functional Genomics of Solid University of Southern California Keck School of Medicine
Tumors Laboratory, Paris, France Los Angeles, CA, USA
List of Contributors xvii

Yaron Rotman Barbara Schroeder


Liver Energy and Metabolism Section, Liver Diseases Branch, Department of Biochemistry and Molecular Biology, Division
National Institute of Diabetes and Digestive and Kidney of Gastroenterology and Hepatology, Mayo Clinic
Diseases Rochester, MN, USA
National Institutes of Health
Bethesda, MA, USA Robert F. Schwabe
Columbia University
Tracey A. Rouault New York, NY, USA
Eunice Kennedy Shriver National Institute of Child Health and
Human Development Christoph Seeger
National Institutes of Health Fox Chase Cancer Center
Bethesda, MD, USA Philadelphia, PA, USA

Nairita Roy Ekihiro Seki


Department of Pathology Division of Digestive and Liver Diseases, Department of
McGowan Institute for Regenerative Medicine, Pittsburgh Medicine, Cedars‐Sinai Medical Center, Los Angeles, CA;
Liver Research Center Department of Biomedical Sciences, Cedars‐Sinai
University of Pittsburgh Medical Center,
Pittsburgh, PA, USA Los Angeles, CA, USA
Praveen Sethupathy
Jayanta Roy Chowdhury
Department of Genetics
Departments of Medicine and Genetics and Marion Bessin
University of North Carolina School of Medicine
Liver Research Center
Chapel Hill, NC;
Albert Einstein College of Medicine
Department of Biomedical Sciences
Bronx, NY, USA
Cornell University
Namita Roy Chowdhury Ithaca, NY, USA
Departments of Medicine and Genetics and Marion Bessin
David A. Shafritz
Liver Research Center
Marion Bessin Liver Research Center
Albert Einstein College of Medicine
Albert Einstein College of Medicine
Bronx, NY, USA
Bronx, NY, USA
David A. Rudnick
Linshan Shang
Department of Pediatrics, Division of Gastroenterology,
Department of Biochemistry, Molecular Biology and Biophysics
Hepatology, and Nutrition and Department of Developmental
University of Minnesota
Biology, Washington University School of Medicine in St.
Minneapolis, MN, USA
Louis, St. Louis Children’s Hospital St. Louis, MO, USA
Reuben J. Shaw
K. Lenhard Rudolph
Molecular and Cell Biology Laboratory
Research Group on Stem Cell Aging, Leibniz Institute on
The Salk Institute for Biological Studies
Aging – Fritz Lipmann Institute (FLI)
La Jolla, CA, USA
Jena, Germany
Junxing Shi
Varman T. Samuel Laboratory of Biochemical Pharmacology, Emory University
Department of Internal Medicine, Section of Endocrinology School of Medicine
Yale University School of Medicine Atlanta, GA, USA
New Haven;
Veterans Affairs Medical Center Jae‐Jun Shim
West Haven, CT, USA Department of Internal Medicine, Kyung Hee University
College of Medicine
Sonja C. Schätzlein Seoul, Korea
Spark@FLI, Leibniz Institute on Aging – Fritz Lipmann
Institute (FLI) Orian S. Shirihai
Jena, Germany Department of Medicine, Division of Endocrinology and
Department of Molecular and Medical Pharmacology
Raymond F. Schinazi David Geffen School of Medicine at UCLA
Laboratory of Biochemical Pharmacology, Emory University Los Angeles, CA, USA
School of Medicine
Atlanta, GA, USA Gerald I. Shulman
Department of Internal Medicine, Section of Endocrinology
Micah B. Schott Department of Molecular and Cellular Physiology and Howard
Department of Biochemistry and Molecular Biology, Division Hughes Medical Institute
of Gastroenterology and Hepatology, Mayo Clinic Yale University School of Medicine
Rochester, MN, USA New Haven, CT, USA
xviii List of Contributors

Bård Smedsrød Gyongyi Szabo


Vascular Biology Research Group, Department of Beth Israel Deaconess Medical Center
Medical Biology Harvard Medical School
UiT The Arctic University of Norway Boston, MA, USA
Tromsø, Norway
Shogo Takahashi
Quinton B. Smith Department of Biochemistry and Molecular and Cellular Biology
Institute for Medical Engineering and Science, Massachusetts Georgetown University
Institute of Technology, Cambridge, MA, USA Washington, DC, USA
Erik Lee Snapp Sijia Tao
Janelia Research Campus of the Howard Hughes Medical Laboratory of Biochemical Pharmacology, Emory University
Institute School of Medicine
Ashburn, VA, USA Atlanta, GA, USA
Raymond E. Soccio Richard J. Thompson
University of Pennsylvania Institute of Liver Studies, King’s College London
Perelman School of Medicine, Department of Medicine London, UK
Division of Endocrinology, Diabetes, and Metabolism
Institute for Diabetes, Obesity, and Metabolism Snorri S. Thorgeirsson
Philadelphia Crescenz VA Medical Center Laboratory of Human Carcinogenesis, Center for Cancer
Philadelphia, PA, USA Research, National Cancer Institute, NIH, Bethesda, MD, USA
Bo Hwa Sohn Michael Trauner
Department of Systems Biology Hans Popper Laboratory of Molecular Hepatology, Division
The University of Texas M.D. Anderson Cancer Center of Gastroenterology and Hepatology Department of Internal
Houston, TX, USA Medicine III
Medical University of Vienna
Karen K. Sørensen
Vienna, Austria
Vascular Biology Research Group, Department of Medical
Biology Stefania Varchetta
UiT The Arctic University of Norway Division of Infectious Diseases and Immunology, Fondazione
Tromsø, Norway IRCCS Policlinico S.Matteo, Pavia; Department of Internal
Medicine and Therapeutics
Lina Spomer
University of Pavia, Italy
Clinic for Gastroenterology, Hepatology and Infectious
Diseases, University Hospital Düsseldorf Medical Faculty at Michelle A. Veronin
Heinrich‐Heine‐University University of Texas at Tyler
Düsseldorf, Germany Fisch College of Pharmacy
David C. Spray Tyler, TX, USA
Marion Bessin Liver Research Center, Albert Einstein Gabrielle Vieyres
College of Medicine Institute of Experimental Virology, TWINCORE, Centre for
Department of Neuroscience, Albert Einstein College of Experimental and Clinical Infection Research
Medicine Hannover, Germany
Department of Medicine, Albert Einstein College of Medicine
Bronx, NY, USA Tiffany N. Vo
Institute for Medical Engineering and Science, Massachusetts
Elizabeth C. Stahl Institute of Technology
Department of Pathology Cambridge, MA, USA
McGowan Institute for Regenerative Medicine, Pittsburgh
Liver Research Center Christopher M. Walker
University of Pittsburgh Center for Vaccines and Immunity, The Research Institute at
Pittsburgh, PA, USA Nationwide Children’s Hospital, and College of Medicine, The
Ohio State University
Mario Strazzabosco
Columbus, OH, USA
International Center for Digestive Health (ICDH), University
of Milan‐Bicocca, Monza, Italy; Xiaoxin X. Wang
Liver Center and Section of Digestive Diseases, Department Department of Biochemistry and Molecular and Cellular Biology
of Internal Medicine, Section of Digestive Diseases, Yale Georgetown University
University School of Medicine, New Haven, CT, USA Washington, DC, USA
List of Contributors xix

John W. Ward Ling Yi


Task Force for Global Health, Decatur, GA; Section on Translational Neuroscience, Molecular
Centers for Disease Control and Prevention Medicine Branch
Atlanta, GA, USA Intramural Research Program, National Institutes of Health,
Bethesda, MD, USA
Eliane Wauthier
Department of Cell Biology and Physiology Xianwen Yi
University of North Carolina School of Medicine Department of Surgery
Chapel Hill, NC, USA University of North Carolina School of Medicine
Chapel Hill, NC, USA
Aubrey V. Weigel
Janelia Research Campus of the Howard Hughes Medical Institute Sun Young Yim
Ashburn, VA, USA Department of Internal Medicine, Korea University
College of Medicine
Rebecca G. Wells
Seoul, Korea
Departments of Medicine (GI Division) and Pathology and
Laboratory Medicine, Perelman School of Medicine, and Lexing Yu
Bioengineering, School of Engineering and Applied Sciences Second Military Medical University
University of Pennsylvania Shanghai, China
Philadelphia, PA, USA
Kenneth S. Zaret
Patrick D. Wilkinson Institute for Regenerative Medicine, Department of Cell and
Department of Pathology Developmental Biology
McGowan Institute for Regenerative Medicine, Pittsburgh Perelman School of Medicine, University of Pennsylvania
Liver Research Center Philadelphia, PA, USA
University of Pittsburgh
Wencheng Zhang
Pittsburgh, PA, USA
Department of Cell Biology and Physiology
Allan W. Wolkoff University of North Carolina School of Medicine
Division of Gastroenterology and Liver Diseases Chapel Hill, NC, USA
Marion Bessin Liver Research Center
Jessica Zucman‐Rossi
Albert Einstein College of Medicine and Montefiore
Centre de Recherche des Cordeliers, Sorbonne Université,
Medical Center
Inserm, Université de Paris, Functional Genomics of Solid
Bronx, NY, USA
Tumors Laboratory, Paris, France;
Xiaogang Xiang Hôpital Européen Georges Pompidou, Assistance Publique‐
Laboratory of Liver Diseases, National Institute on Alcohol Hôpitaux de Paris
Abuse and Alcoholism Paris, France
National Institutes of Health
Vanessa Zuzarte‐Luis
Bethesda, MD, USA;
Instituto de Medicina Molecular João Lobo Antunes
Department of Infectious Diseases, Ruijin Hospital,
Faculty of Medicine, University of Lisbon
School of Medicine
Lisbon, Portugal
Shanghai Jiao Tong University
Shanghai, China
Yusuke Yamamoto
Division of Molecular and Cellular Medicine, National Cancer
Center Research Institute
Tsukiji, Tokyo, Japan
Preface

The pace of discoveries in basic biomedical sciences and engi- analysis, engineered drug design, signaling networks, immune
neering and their application to diagnosis and treatment of liver mechanisms and tolerance, the brain, and metabolic/digestive
disease continues to exceed greatly the expectations expressed functions. Discoveries in these disciplines have already facili-
in the Preface to the previous editions published in 1982, 1988, tated diagnosis, treatment, and improved clinical outcome of
1994, 1999, and 2009. Concomitantly, the challenge addressed many liver diseases. Much more is undoubtedly yet to come.
by this book has not changed since first appearing in the Preface This sixth edition contains new chapters that present major
to the first edition over 30 years ago: progress that has been achieved in research laboratories and
clinics around the world. All other chapters have been com-
The amazing advances in fundamental biology that have pletely revised and updated. Following the death of our col-
occurred within the past two decades have brought hepatol- league Nelson Fausto, Snorri Thorgeirsson became an Associate
ogy and other disciplines into new, uncharted and exciting Editor. Previous editions included a section called “Horizons,”
waters. The dynamic changes in biology will profoundly influ- devoted to extraordinary advances in areas of potentially major
ence our ability to diagnose, treat and prevent liver disease. importance to the liver. Virtually all of these fields have rapidly
How can a student of the liver and its diseases maintain a link expanded and become topics for later chapters. Sixteen new
to these exciting advances? Most physicians lack the time to “Horizons” chapters are presented in this edition. One may
take post‐graduate courses in basic biology; most basic safely predict that their impact on the field of hepatology will be
researchers lack an understanding of liver physiology and considerable. As stated in the Preface to previous editions:
disease. This book strives to bridge the ever‐increasing gap
between the advances in basic biology and their application
The amazing advance in science proceeds at an ever‐increasing
to liver structure, function and disease.
pace. The implications for students of liver disease are
­considerable. The authors and editors will have achieved our
Molecular biology was not the only great wave in contempo- goals if the reader finds within this volume glimpses into the cur-
rary science, nor is it surely the last. Remarkable advances in rent state and future direction of our discipline and p­ erspectives
genetics and various omics are increasingly linked with dynamic that lead to better understanding of liver ­function and disease.
super‐resolution light microscopy, which permits the study of
cellular, molecular, and organ‐based physiology at nano‐levels.
The expanding worlds of RNA structure and function, CRISPR‐ Irwin M. Arias
type gene editing, and chromatin biology coupled with single‐ Harvey J. Alter
cell and single‐molecule genomic analyses are facilitating James L. Boyer
discoveries with great importance in organ physiology and David E. Cohen
medicine, including personalized diagnosis and treatment.
­ David A. Shafritz
Unexpected discoveries are certain to emerge from the ongoing Snorri S. Thorgeirsson
bridge‐building between chemical and physical structural Allan W. Wolkoff
Acknowledgments

We thank the distinguished authors for their expertise, of Wiley Blackwell and also to the freelance project manager
­enthusiastic participation, and patience in responding to edi- Gillian Whitley.
torial suggestions. Appreciation is also extended to the staff
PART ONE:
INTRODUCTION
Organizational Principles
1 of the Liver
Peter Nagy1, Snorri S. Thorgeirsson2, and Joe W. Grisham3
1
First Department of Pathology and Experimental Cancer Research, Semmelweis University, Budapest, Hungary
2
Laboratory of Human Carcinogenesis, Center for Cancer Research, National Cancer Institute, NIH, Bethesda,
MD, USA
3
Department of Pathology and Laboratory Medicine, University of North Carolina, Chapel Hill, NC, USA

PRINCIPLES OF LIVER STRUCTURE liver. In addition, they perform the most complex metabolic
AND FUNCTION tasks of the mammalian organism.
The cholangiocytes form the channels that constitute the
The liver is the largest organ of the mammalian body and has a ­biliary system, which drains the parenchyma and guarantees the
highly versatile and complex function. Its specialized role is permanent flow of the bile, a highly toxic solution. Cholangiocytes
shown by the fact that, despite intense efforts, the activity of the also modify the composition of the bile and, in case of adverse
liver cannot be replaced by artificial equipment. The liver par­ conditions, can participate in repair mechanisms. These liver
ticipates in the maintenance of the organism’s homeostasis as cells could not carry out their specific functions, of course,
an  active, bidirectional biofilter. It is classed as bidirectional ­without the support of several “communal” cell types, which
because it filters the portal blood that transports nutritional and are  highly adapted to the special function and architecture of
toxic compounds from the environment through the gastrointes­ the liver. The endothelial cells of the parenchyma have a unique
tinal tract and also filters the systemic blood (the body’s own fenestrated structure and various different ­subpopulations can
products, e.g. bilirubin), providing the only channel of the body, be  distinguished. There are several subpopulations of hepatic
the biliary system, through which non‐water‐soluble substances myofibroblasts as well. In addition to their mechanical functions
can be removed. It is classed as an active filter because it rapidly the myofibroblasts can store special substances (e.g. vitamin
metabolizes most nutritional compounds and neutralizes and A in stellate cells) and are a major source of growth factors and
prepares for removal toxic exogenous (xenobiotics) and endog­ cytokines. The Kupffer cells are the resident macrophages in the
enous (worn out) materials. Because of these major functions liver. In addition to filtering the blood, they perform their tradi­
the liver is constantly exposed to intense microbiological and tional immunregulatory function. The presence of almost all
antigenic stimuli which require function of the innate and adap­ subtypes of lymphocytes and dendritic cells makes the liver the
tive immune systems. These diversified functions are executed largest organ of the immune system. The mesothelial cells of the
by a structurally complex, multicellular tissue with a unique Glisson capsule are, beside their mechanical function, an impor­
angioarchitecture, and by the combined and integrated activities tant source of lymph production and can contribute to the gen­
of the participants. eration of other hepatic cell types. The features of the hepatic
There are only two unique cell types in the liver – hepatocytes extracellular matrix are unique. The components of the basement
and biliary cells (or cholangiocytes). The hepatocytes are “the membrane are present around the sinusoids in an “unstructured”
most valuable” parenchymal cells of the hepatic tissue. They do fashion, and cannot be detected by electron microscope, yet they
not constitute a homogeneous cell population. They are highly can perform certain functions.
polarized cells (i.e. molecular specializations of the various Another fundamental feature of liver organization is its
­surface membranes, including receptors, pumps, transport chan­ unique vascular pattern. Two afferent vessels supply blood to
nels and carrier proteins) and their functions and to a certain the liver: the portal vein and the hepatic artery. The blood of
extent morphology depend on their location in the parenchyma. the  portal vein, having already “drained” the stomach, gut,
This polarization makes the hepatocytes the logical center of the ­pancreas, and spleen, is reduced in oxygen and pressure, and is

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
4 THE LIVER:  FUNCTIONAL ANATOMY OF THE LIVER

enriched in nutrients and toxic materials absorbed from the hepatocytes runs in a centrifugal direction in the bile canalicules
alimentary tract and in viscerally generated hormones and
­ formed by neighboring hepatocytes and is collected by the inter­
growth factors. The arterial blood of the hepatic artery has sys­ lobular bile ducts of the portal triads. There is thus a countercur­
temic levels of oxygen, pressure, and composition. The major rent between the flow of the blood and bile at lobular level.
function of the hepatic artery is to supply the peribiliary vascu­
lar plexus, the portal tract interstitium, the hepatic capsule, and
the vasa vasorum of major vessels. In some species, the hepatic
artery forms anastomosis with the branches of the portal vein, FUNCTIONAL ANATOMY OF THE LIVER
but even then this blood also ends up in the sinusoids. The blood
of the liver is collected by one efferent draining system, the Macroanatomy
hepatic or “central” veins, which reach the systemic circulation
The liver is a continuous sponge‐like parenchymal mass
via the inferior vena cava. The sinusoids form a very special
­penetrated by tunnels (lacunae) that contain the interdigitating
vascular system, which is interposed between the afferent and
networks of afferent and efferent vessels [1]. The adult human
efferent vessels. The large number and capacity of the sinusoids
liver weighs from 1300 to 1700 g, depending on sex and body
and the special arrangement of the supplying vessels provide a
size. It is relatively small compared to other species (2% of the
large volume of blood at a high flow rate via the large vessels
body weight) – in rat and mouse the liver is 4–5% of the body.
with high compliance and capacity. At the same time the
In most mammalian species the liver is multilobed, the indi­
­sinusoids are perfused with blood at low pressure and flow rate.
vidual lobes reflecting the distribution of the major branches of
These arrangements (i.e. low flow, specifically fenestrated
afferent and efferent blood vessels. In contrast, the human liver
(­perforated) endothelial cells, and the lack of the structured
parenchyma is fused into a continuous parenchymal mass with
basement membrane) provide an especially efficient communi­
two major lobes, right and left, delineated only by being supplied
cation between the blood and hepatocytes. This is well illus­
and drained by separate first‐ and second‐order branches of the
trated by the pathological condition of liver cirrhosis, when the
portal and hepatic veins. Right and left lobes are topographically
changes in hemodynamic condition (i.e. the “capillarization” of
separated by the remnants of the embryonic umbilical vein (the
the sinusoids) disrupts this communication, resulting in severe
falciform ligament), but this landmark does not locate the true
dysfunction of the liver.
anatomic division. Anatomically, the medial segment of the left
Bile acids and their enterohepatic circulation are another good
lobe is located to the right side of the falciform ligament, ­centered
example of the cumulating functions. The bile acids are synthe­
on the anterior branches of the left portal vein. Interdigitation of
tized in the hepatocytes by a complex biochemical process that
first‐ and second‐order branches of the portal and hepatic veins
requires 16 different enzymes, which are further modified by the
produces eight macrovascular parenchymal segments centered
gut microbiota. The primary physiological function of the bile
on large portal veins and separated by large hepatic veins [2].
acids is to convert lipid bilayers into micelles. This makes pos­
Hemodynamic watersheds or fissures separating afferent and
sible the excretion of important waste products from the blood.
efferent macrovascular segments permit the surgical resection of
The bile acids also emulsify elements of the food in the gut and
individual or adjacent segments.
aid their ­absorption. In addition, bile acids act as signaling mol­
Liver transplantation and surgery has reached such a com­
ecules, synchronizing the cooperation of the liver and gut.
plexity, however, that the traditional eight‐segment scheme is no
The different types of cells and vessels mentioned above can
longer sufficient. Detailed histological and imaging investiga­
operate only if they are organized in a well “designed” structure.
tions have revealed that the number of second‐order branches
The most widely studied and analyzed morphological and func­
given off by the left and right portal veins is much higher, and
tional unit or module of the liver is the hepatic lobule. The popu­
the mean of their number is 20, leading to the “1–2–20” concept
larity of this structure for studies can be partly explained by the
of portal venous segmentation [3]. The recognition of the water­
fact that lobules are outlined nicely in some species (pig, camel,
shed septa between the variable actual segments is helped by
bear) by connective tissue septa, and can therefore be easily
intraoperative imaging techniques in real operative situations.
­recognized on the two‐dimensional histological sections com­
monly used in structural studies. The idealized lobule has a
polygonal (usually hexagonal) shape. The terminal branch of
the hepatic vein (central vein) is in the center of the lobule while
Microanatomy
the corners are occupied by the “portal triads.” The components Normal liver function requires the unique arrangement of basic
of the triad are the interlobular bile ducts and the terminal components of hepatic tissue: portal vein, hepatic artery, bile
branches of the portal vein and hepatic artery. The blood carried duct, hepatic vein, and hepatocytes. These form in two‐­
by these afferent vessels is distributed by the inlet venules and dimensional sheets the above‐mentioned hepatic (classical of
arteries along the virtual “vascular septa.” This vascular frame is Kiernan’s) lobules. Profiles of portal tracts and hepatic veins of
filled up columns (or sheets in three‐dimensional space) of the various sizes are a prominent feature of liver histology [4–6].
hepatocytes constructed as “plates” arranged in a radial fashion. Smaller branches of the afferent and efferent vessels (together
The hepatic plates are separated by the similarly distributed with their stromal components) predominate in tissue sections
sinusoids. The blood runs in a centripetal direction from the taken from peripheral, subcapsular locations, whereas tissue
­vascular septa to the central vein. The vascular septa secure sections taken from more proximal areas nearer to the hilum
the mixing of the portal venous and arterial blood and the more‐ contain larger vascular structures  [6]. These vascular/stromal
or‐less equal supply to the sinuses. The bile produced by the elements are contained in tunnels (lacunae) that penetrate the
1:  Organizational Principles of the Liver 5

contain preterminal hepatic venules. These vessels represent the


seventh‐ to tenth‐order branches from the hilar portal vein in
large mammals, such as humans. These small portal tracts and
hepatic (central) veins penetrate the parenchyma in nearly paral­
lel orientations about 0.5–1.0 mm apart. The portal inlet venules
are very short vessels with no smooth muscle in their walls. They
branch from preterminal and terminal venules at points on the
circumference of the lobules at about 120 radial degrees (­triradial
branching) and penetrate the parenchyma together with terminal
arteriolar branches approximately perpendicular to and midway
between two adjacent terminal hepatic venules [5, 6]. During
their course through the parenchyma portal inlet venules break
up completely into sinusoids, which are oriented more or less
perpendicularly to the veins. Because they are hardly larger than
sinusoids, the inlet venules are not conspicuous in humans and
Figure 1.1  Schematic drawing of the organization of blood vessels other mammals that lack a definite connective sheath around
(arteries, red; portal veins, purple; central veins, blue; bile ducts, green; them. However, in adult swine their course through the paren­
lumen of the biliary channels, including bile canaliculi, yellow) in two
chyma is clearly marked by connective tissue.
adjacent lobules of human liver. One sixth of a lobule is visible on the
right and one third of a lobule is visible on the left. A terminal portal Capillary‐size sinusoids occupy the smallest and most
venule, arteriole, and bile ductule (canal of Hering) are present in the ­numerous tunnels (lacunae) in the parenchymal mass [4]. Unlike
vascular septum between the lobules. The arteriole is connected directly capillaries elsewhere, liver sinusoids are composed of endothe­
to the sinusoids or enters the inlet venule. Bile is drained over the whole lial cells that are penetrated by holes (fenestrae) and lack a
surface of the lobule. Arterioles and bile ductules are not present in the structured basal membrane [12], features that allow free egress
vascular septum in rodents. The bile is drained through canals of Hering of the fluid components and solutes of the perfusing blood.
connected to hepatocytes of the limiting plate. The arterioles anasto­
mose with the portal system at higher level as well. Courtesy of Sandor
For example, tagged albumin has access to a space in the liver
Paku, Semmelweis University, Budapest, Hungary. that is about 48% larger than the sinusoidal volume, in contrast
to other tissues in which capillary space and albumin space are
nearly the same [13]. In favorably oriented histological sections,
parenchymal mass [4]. The hepatocytes arranged in plates fill in more or less parallel, longitudinal profiles of sinusoids alternate
the space between the portal tracts and hepatic veins (Figure 1.1). with hepatic plates  [14]. A narrow cleft, called the space of
The hepatic plates form brick‐like walls (muralia) of hepato­ Disse, separates sinusoids from hepatocytes located in adjacent
cytes one cell (one brick) thick. The first hepatocytes of the hepatic plates [12, 15]. At their proximal (portal venous) ends,
hepatic plates form a virtual barrier between the periportal con­ sinusoids are narrow and somewhat tortuous, whereas their mid­
nective tissue and the liver parenchyma called a limiting plate. dle and distal (hepatic venous) portions are larger and straighter
The blood vessels and their investments of connective tissue [9, 16, 17]. Sinusoids and hepatic plates are disposed radially
provide the soft, spongy liver with its major structural support, around the draining hepatic veins and extend directly to the
or “skeleton.” Larger afferent vessels, portal veins, and hepatic ­supplying inlet venules [17].
arteries are contained together with bile ducts in connective Three‐dimensional reconstruction of the interlobular zone
­tissue – the portal tracts – which are continuous with the mesen­ revealed the existence of a small vessel in this plane, the ­vascular
chymal components of the liver’s mesothelium‐covered surface septum, that serves as a starting pool for intralobular sinusoids.
capsule (Glisson’s capsule). Portal tracts also contain lymphatic This is a hemodynamic barrier, a “watershed” between the
vessels, nerves, and varying populations of other types of cells, two  neighboring lobules. This “interlobular” septum contains
such as macrophages, immunocytes, myofibroblasts, and pos­ ­connective tissue matrix in pig, camel, bear, etc. and outlines the
sibly hematopoietic stem cells (see [7] and references therein). lobules nicely; it also exists in a rudimentary form in human
The collagenous investment of the efferent vessels is less robust liver [18].
and lacks large numbers of adventitious cells. The bile  –  the excretion product of the hepatocytes  –  is
The hepatic artery is distributed to the tissues of portal tracts, ­collected and transported in bile canaliculi, which are formed by
the liver capsule, and the walls of large vessels [4–6]. In portal the apical sides of two adjacent hepatocytes in the hepatic plate.
tracts arterial branches form a capillary network (the peribiliary The network of canaliculi is drained into the interlobular bile
plexus) arborized around bile ducts [8, 9]. Efferent twigs from ducts through interface structures called canals of Hering. These
the peribiliary plexus empty into adjacent portal veins in rat and are intermediary structures constructed partly by hepatocytes or
mouse but not in human and hamster [10]. The portal vein sup­ cholangiocytes (Figure  1.2). Since these structures are the
plies blood to the parenchymal mass through the so‐called inlet ­primary candidates to harbor the hepatic stem cell compartment,
venules [9, 11]. they are the subject of intensive investigations [19]. The distribu­
In histological sections of mammalian liver, afferent and tion of canals of Hering shows variation among different spe­
efferent vessels interdigitate regularly in an approximate ratio of cies. They are characterized by a distinct (EMA−/CD56+/CD133+)
5–6 portal tracts for each profile of a hepatic vein, to form a pat­ immunophenotype in humans, leave the periportal space and
tern of cross‐sections of portal tracts and hepatic veins separated spread into the parenchyma along the rudimentary interlobular
by parenchyma [5, 6]. Most of the cross‐sectioned portal tracts septa, and thus do not enter into the hepatic lobule  [18].
6 THE LIVER:  FUNCTIONAL ANATOMY OF THE LIVER

Figure 1.2  Normal human liver stained for CD10. The hepatocytes
form trabeculae, 1–2 cells thick, radiating from a central vein. CD10 is
Figure 1.4  Normal human liver stained for pankeratin (green) and kera­
expressed on the canalicular domain, indicating the polarization of the
tin 7 (red). The dark lane in the center represents an interlobular vascular
cells. Courtesy of Sandor Paku, Semmelweis University, Budapest,
septum. The double positive (yellow) biliary ductules have several con­
Hungary.
nections with the limiting hepatocytes but do not enter into the lobules.
Courtesy of Sandor Paku, Semmelweis University, Budapest, Hungary.

modular arrangement can improve the interpretation of lesions,


especially in pathologically altered livers, but it is certainly not
easy to transform the two‐dimensional observations into three‐
dimensional space.

Functional unit of the liver


The concept of the primary functional unit of the liver has been
the subject of debate for more than 350 years since its descrip­
tion by Wepfler in 1664 [22]. The first and most widely accepted
traditional unit of the liver is “Kiernan’s lobule” [23], as
described earlier. This is the efferent microvascular segment,
being the smallest unit of parenchyma that is drained of blood by
a single efferent (terminal hepatic or central) vein. It is quite
Figure 1.3  Normal rat liver stained for pankeratin (green) and laminin easy to identify it, especially in species where they are outlined
(red), the nuclei are labeled by TOTO (blue). There is a cross‐section of
a canal of Hering beside a small portal vein. The cytoplasm of the chol­
by connective tissue. The major criticism of the concept is that
angiocytes is strongly positive for keratin and these cells are outlined by the terminal afferent vessels through the vascular septa contrib­
laminin (basement membrane) but basement membrane is absent at the ute to the blood supply of adjacent lobules, and therefore the
pole where the ductile is connected to an adjacent hepatocyte. Courtesy lobule cannot be a “basic functional unit.” Rappaport defined the
of Sandor Paku, Semmelweis University, Budapest, Hungary. basic unit as the compartment of the hepatic parenchyma sup­
plied with blood by a single terminal portal vein and called this
The interlobular bile ducts are lined by a single layer of cuboidal unit the “liver acinus” [24]. Now we know that this unit is also
cholangiocytes (Figure  1.3). They anastomose and unite larger supplied by a single terminal branch of the hepatic artery. The
septal and hilar branches. The connective tissue around the larg­ simple acinus is a parenchymal mass around a portal tract and it
est biliary branches contain peribiliary glands which also secrete is drained by more hepatic venules. The acinus is subdivided
into the biliary tract (Figure 1.4). into three zones, based on the distance from the portal vein.
Teutsch and coworkers [20, 21] analyzed serial sections of rat The distribution of these areas fits to the functional zonality of
and human livers to reconstruct the three‐dimensional structure the hepatic parenchyma. Pathological lesions (e.g. steatosis or
of hepatic tissue. Although there were differences between the necrosis) also often follow this zonal pattern, which made this
two species, the basic arrangement was similar. The reconstruc­ unit attractive. However, this zonality did not correspond per­
tion revealed primary “modules,” which constructed a more fectly to the distribution of enzyme activities and the hepatic
complex “secondary” module. The integration was based on a modules described by Teutsch [20, 21] were also not compatible
common drainage by branches of the hepatic veins and supply­ with the concept of the acinus.
ing portal veins, and the modules were covered by continuous Matsumoto and his colleagues [6] investigated the angioarchi­
vascular septa. The primary modules correspond to the two‐ tecture of the human liver on thousands of serial sections, dis­
dimensional hepatic lobules. Quite a substantial variation in the tinguishing conducting and parenchymal portions of the portal
shape and size of the modules was found, which provides venous tree. A cone‐shaped parenchymal portion (primary lob­
­morphogenetic plasticity to construct the whole organ. This ule) was defined which was supplied by a terminal portal venule.
1:  Organizational Principles of the Liver 7

The circulatory network of this unit was named the hepatic regulated by a sphincter in the terminal arterioles [29]. In addi­
microcirculatory subunit (HMS). Ekataksin and Wake [25] tion, hepatic stellate cells are the pericytes of the sinusoids and
showed that the afferent microvascular segment is also the small­ are the principal regulators of sinusoidal blood flow. They have
est unit of parenchyma drained of bile by terminal bile ducts receptors for molecular mediators regulating contraction and
[25], demonstrating that this hemodynamic segment is also the are closely associated with nerve endings, indicating neurogenic
smallest excretory unit of parenchyma. The compartment of the influence [30]. Sinusoids appear to have limited contractile abil­
hepatic parenchyma associated with an HMS contains all of ity, possibly produced by contraction of encircling stellate cells
the elementary structures of the hepatic tissue and may represent (pericytes) [29, 31]. Studies of the exteriorized liver of rodents
the elementary functional and morphological unit of the liver. It suggest that sinusoidal flow may be regulated at both inlet and
was named cholehepaton. The cholehepaton is also the smallest outlet levels [32], although other similar studies have not
bile–blood unit and conforms with the principle of countercur­ detected sphincters at either point [28]. However, sinusoidal
rent flow. The cholehepaton has no anatomical or structural flow is strongly affected by post‐sinusoidal resistance [27].
­borders, and cannot be recognized on histological sections. It is
defined by its function and mostly corresponds to Matsumoto’s
primary lobules. Six of these Matsumoto’s primary lobules con­
stitute the secondary lobule, which is almost identical with the THE CELLS OF THE LIVER
classical Kiernan’s lobule. This long and complicated detour
returned us to the classical lobule, which is widely used today to Hepatocytes
analyze reactions of the hepatic tissue.
Hepatocytes are commonly denoted as “parenchymal cells” and
However, it is worth remembering that there are other types
the other cells of the liver tissue matrix as “non‐parenchymal
of hepatic functional units and some workers contend that the
cells.” This convention is somewhat artificial since hepatocytes
hepatic tissue is an indivisible continuum that has no definable
alone are not competent to perform all ­ essential hepatic
units [26].
­functions, and several types of cells in the liver tissue matrix
function as an integrated community to carry out conjointly the
Liver hemodynamics multiplicity of hepatic functions. Functional integration of this
cellular community is accomplished by several communication
The hepatic vasculature is characterized by high capacity, high
mechanisms, including signaling networks involving numerous
compliance, and low resistance [27]. Blood vessels comprise
cytokines and chemokines, and by direct transfer of small mol­
about 22% of the liver’s mass/volume [13] and the liver contains
ecules through gap junctions [33]. Hepatocytes, responsible for
about 12% of the body’s total blood volume under physiological
most of the synthetic and many of the metabolic functions of the
conditions [27], a sizeable fraction of which can be expelled by
liver (see Chapters 23 and 24), are large polygonal cells (averag­
contraction of larger vessels by sympathetic nerve stimulation. In
ing about 25–30 μm in cross‐section and 5000–6000 μm3 in
other words, the liver is a blood reservoir. The pressure of portal
­volume [34]). They are the most numerous cells in the liver
venous blood is reduced as the major afferent vessels dichotomize
parenchyma; the adult human liver probably contains about 1011
through the parenchyma, from about 130 mmH2O in the extrahe­
hepatocytes, representing about 60% of all cells in the paren­
patic portal vein to about 60 mmH2O in the preterminal portal
chymal matrix and comprising about 80% of its mass/volume
veins of the exteriorized liver of the anesthetized rat, amounting to
[15]. Hepatocytes are shaped as complex rhomboids with
about 60% of the total transhepatic pressure gradient [13]. A
­several distinct surfaces [34]. They are polarized by molecular
­similar portal pressure gradient has been found in humans [27].
specializations of their various surface membranes in the forms
Blood flow through the liver amounts to about 1500–2000 mL
of receptors, pumps, transport c­hannels, and carrier proteins
min−1 in adult humans, about 25% of the resting cardiac output
(see Chapters 5 and 6) that comprise three functionally distinct
[27]. About 25% of the total liver blood flow is derived from the
domains (see Chapter  6): the basolateral domain, the lateral
hepatic artery at prevailing arterial pressure and oxygenation. The
domain, and the canalicular domain.
portal venous blood (about 75% of total liver blood flow) arrives at
the liver partially depleted of oxygen and at a reduced pressure as • Basolateral domain: The basolateral (or sinusoidal) domain
a consequence of having already perfused the splanchnic viscera. constitutes about 35% of the total hepatocyte surface facing the
In aggregate, sinusoids comprise about 60% of the liver’s sinusoids. The area of this surface is greatly amplified by the
vascular volume, or about 13% of the total liver mass/volume folding of the plasma membrane to form innumerable micro­
[13]. A significant decrease in blood pressure occurs in sinu­ villi that extend into the space of Disse [34]. This membrane is
soids (about 40% of the transhepatic pressure gradient), the equipped for extraction of a great variety of molecules from the
pressure declining to about 25 mmH2O in terminal hepatic veins blood and for the simultaneous secretion into the blood of other
of exteriorized liver of anesthetized rats [28]; the pressure gradi­ molecules that have been modified or synthesized by hepato­
ent in the short sinusoids is especially steep. Blood pressure in cytes. The cell matrix adhesion molecules are also present on
the inferior vena cava approximates that in the terminal hepatic this side of the hepatocyte. Although there is no structured
vein [27]. Consequently, although flow of blood through sinu­ basement membrane between the hepatocytes and sinusoidal
soids faces little resistance, it is slow and somewhat intermittent endothelial cells, the lack of integrin‐linked kinase 1 (Ilk1)
and is assisted by negative pressure produced by respiratory results in disruption of the normal structure  [35], supporting
expiration [27]. Possible mechanisms of regulating blood flow the necessity of hepatocyte–matrix interaction. About 50% of
within sinusoids are controversial. The sinusoidal circulation is the total hepatocyte surface faces adjacent hepatocytes [34].
8 THE LIVER:  THE CELLS OF THE LIVER

• Lateral domain: The plasma membranes of these intercellular carbohydrate metabolism (gluconeogenesis and glycogen stor­
surfaces are mostly flat lateral domain, which is the least com­ age by periportal hepatocytes; glycolysis by perihepatic vein
plex. This surface contains intercellular adhesion complexes ­hepatocytes). The ammonia and fatty acid metabolisms are also
(tight junctions, intermediate junctions, and desmosomes) zonally distributed. Zone 1 hepatocytes are engaged in urea pro­
that pin together the adjacent hepatocytes, form a permeabil­ duction and β‐oxidation of fatty acids, while zone 3 hepatocytes
ity barrier between the peri‐sinusoidal space of Disse and remove nitrogen by glutamine synthetase and perform lipogen­
bile canaliculi, and include gap junctions that allow commu­ esis. The metabolism of xenobiotics is more prominent in the
nication between adjacent hepatocytes by transfer of small pericentral hepatocytes. Recent research has shown that many
molecules. liver functions are dispersed heterogeneously, with dispersed
• Canalicular domain: Infoldings of the lateral surfaces create functions often acting as integrated parts of coordinate meta­
bile canaliculi, which comprise about 13% of the total hepato­ bolic systems [44].
cyte surface. This is termed the canalicular domain [34]. It is Zonation of liver functions was thought to be related to
also amplified by microvilli and modified for bile excretion. ­sinusoidal hemodynamics, which produced gradients in blood‐
The canalicular surface is confined by strong junctional com­ borne substances available to hepatocytes and other cells of the
plexes. Bile canaliculi form a belt‐like extracellular space parenchymal matrix [44]. Recent evidence gained mostly from
(about 1 μm in diameter) that is continuous along the lengths experiments with genetically engineered mice proved that the
of the hepatic plates, connecting at the portal ends with bile Wnt/β‐catenin pathway is the master regulator of hepatocytic
ducts. The molecular mechanisms responsible for the polarity zonation, but the interaction with lymphoid enhancer factor 4
of the canalicular domain are well characterized. Hepatocyte (Lef4) and HNF4α is also critical [45]. Hepatocytes and other
nuclear factor 4 (Hnf4α) is the master regulator of morpho­ cells located at afferent and efferent ends of hepatic plates are
logical and functional hepatocyte differentiation [36], but sev­ subjected to different microenvironmental conditions. Certain
eral other factors, such as liver kinase B1 (Lkb1) [37], molecules are largely extracted by the first hepatocytes that
vacuolar sorting protein 33b (Vps33b) [38], and claudin‐15 encounter the perfusing blood, lowering their concentration
[39], are also required for the polarization. downstream. For example, oxygen levels in the blood at affer­
ent and efferent ends of sinusoids differ greatly because oxygen
As befits their numerous metabolic functions, hepatocytes is efficiently extracted by hepatocytes located at the afferent
contain a complex array of mitochondria (~1700 per cell on ends of hepatic plates, exposing downstream hepatocytes to
average), peroxisomes (~370 per cell), lysosomes (~250 per relatively hypoxic conditions; the oxygen gradient alone can
cell), Golgi complexes (~50 per cell), aggregates of rough and explain much of the heterogeneity of hepatocyte function
smooth endoplasmic reticulum (~15% of cell volume), and related to position in plates [46]. Other molecules modified or
numerous microtubules/microfilaments [34]. produced by upstream hepatocytes are excreted into the sinu­
Polyploidization is another unique feature of the hepatocytes. soidal blood and may be removed by hepatocytes located fur­
This is the result of defective cytokinesis, which seems to be ther downstream. The complex interplay of metabolite
regulated by the insulin–Akt signaling pathway [40]. The extent concentration in the perfusing blood, coupled with extraction,
of ploidy increases with age. The presence of aneuploid hepato­ modification, secretion, re‐extraction, and further modification,
cytes in the normal liver is also well established [41]. The exact influence the metabolic events that occur in individual cells and
function of this process is still unknown but it is thought to help define ­unequal parenchymal territories that produce zonal vari­
in adaptation to chronic injury. ations in different physiological capabilities and pathological
Hepatic plates and adjacent sinusoids form associations that susceptibilities [44]. Disruption of metabolic zonation has
­
are structurally similar in all parts of the liver. Various liver cells severe consequences.
show numerical, structural, and functional heterogeneities Physiological turnover of hepatocytes occurs slowly. They
related to their location along the afferent–efferent axis of have a lifespan of about 400 days in an adult steady‐state hepat­
hepatic plates and sinusoids. Among the structural differences ocyte population, about 0.025% of which typically are cycling
are ploidy variations in hepatocytes; in adult mammals hepato­ [32] and the remainder rest in G0 phase. Although the hepato­
cytes located at the portal ends of hepatic plates are diploid, cytes are ready to re‐enter the cell cycle upon injury, this capac­
whereas cells of higher ploidy are located further downstream ity decreases with age. There are, however nonresolved opposing
[42]. Gap junctions containing connexin 26 are more numerous opinions about the replacement of lost hepatocytes during
on portal hepatocytes, whereas junctions containing connexin homeostatic conditions. The contribution of hematopoietic cells
32 are distributed on hepatocytes in all parts of hepatic to liver maintenance was a very popular idea 20 years ago.
plates [43]. These variations in structure and cellular composi­ Repopulation of bone marrow‐derived cells (macrophages,
tion are associated with functional differences among hepato­ myofibroblasts) in the liver is evident in recipients of liver trans­
cytes located at different points along the afferent–efferent axis plants, in which these types of liver cells are replaced with cells
of plates and sinusoids. Rappaport divided the portal‐hepatic of the host genotype [47]. In contrast, hepatocytes are not gener­
(afferent–efferent) lengths of hepatic plates into three arbitrary ated from bone marrow cells in significant numbers under either
zones (termed I, II, and III). Hepatocytes located in these zones circumstance [47].
differ in their functional capabilities and susceptibilities to path­ However, the ancient and very attractive theory of “streaming
ological damage [24], and pathways performing opposing liver” keeps returning. It was originally proposed by Zajicek and
­functions follow an inverse distribution along the portocentral colleagues [48] that periportal hepatocytes have enhanced repli­
axis. This is strikingly exemplified by regional differences in cative potential and the progeny of these cells spread under
1:  Organizational Principles of the Liver 9

homeostatic condition in a portal–central direction. Although few highly specialized endothelial cells with peculiar morphology
studies applying lineage‐tracing methodology have supported and function. They are extremely thin in normal liver, 150–170
this notion, most of the studies have ruled out this option. Most nm across, and this attenuated cytoplasm is fenestrated. The
recently, Axin2+ hepatocytes [49] abutting hepatic veins as well diameter of these “perforations” ranges from 50 to 200 nm, and
as hybrid periportal hepatocytes [50] have been reported to have they are clustered together in groups to form “sieve plates” [60].
selective growth advantage. Based on these results a bidirectional The fenestrae are dynamic structures, and their actual diameter
streaming hypothesis, “somehow like ocean water entering into is influenced by blood pressure, composition of extracellular
the delta of the Amazon River at high tide” has been proposed matrix, hormones, etc. The porosity of the sinusoidal endothe­
[51]. This issue will no doubt be resolved in the near future. lial cells is polarized, and the fenestrae are more numerous in
the centrilobular zone. The maintenance of the fenestration
requires paracrine and autorcrine signals, which are mostly pro­
Cholangiocytes vided by hepatocytes and stellate cells [61]. Surprisingly, these
Cholangiocytes, or biliary epithelial cells, comprise much less endothelial cells have mostly anaerobic metabolism, providing
than 1% of the total number of cells in the liver parenchyma, since lactate to the adjacent hepatocytes.
most are located in bile ducts in portal tracts [52]. Only the small­ High‐resolution in vivo microscopy has revealed swelling
est bile ducts penetrate the parenchymal mass in the company of and contracting of these cells as a response to vasoactive sub­
terminal portal veins, where they connect with bile canaliculi in stances, suggesting that they participate in the regulation of
hepatic plates. The points of connection of ducts with hepatic blood flow [62]. A unique functional feature of the sinusoidal
plates are defined by tubular structures called the canals of endothelial cells is their high endocytic capacity. This process is
Hering, which are composed of both cholangiocytes and hepato­ mediated by all sorts of scavenger receptors, providing the main
cytes [53]. They are also the location of liver epithelial stem cells pathway for clearance of weakened molecules from circulation
that can differentiate into both hepatocytes and cholangiocytes [63, 64] (see Chapters 26–28). Endothelial sinusoidal cells also
[54, 55]. Larger bile ducts contain cholangiocytes that rest on a express several types of PRR (e.g. mannose receptor several
basal membrane and vary in number and size in proportion to TLRs) and are able to produce inflammatory cytokines (e.g.
duct size [56]. The cholangiocytes are also polarized cells, with TNF and IL‐6) as well as playing a significant role in the innate
an apical (luminal) and a basolateral domain. Their luminal sur­ immunity. In spite of intensive studies, their role in adaptive
face membranes are expanded by microvilli and a single primary immune reactions is still controversial [52].
cilium, which is a sensor for mechanical, osmolar, and chemical The lifespan of sinusoidal endothelial cells is not known;
stress [57]. Although they contain fewer mitochondria and sparser they divide rarely and progenitor cells seem to be important in
endoplasmic reticulum than do hepatocytes, cholangiocytes in their maintenance. Two populations of progenitor cells can be
intrahepatic bile ducts, together with the network of capillaries distinguished [65]: the liver‐resident population is thought to be
that surrounds them (the peribiliary plexus), form a metabolic responsible for normal cell turnover, while bone marrow‐derived
unit that modifies the composition of canalicular bile [58]. cells help to replenish the sinusoidal endothelial cells when
The biliary tree is not just a draining pipe and 70–90% of the necessary.
bile volume is produced by the cholangiocytes. They change the
composition of the bile by secretion and absorption. The main
secretory product is bicarbonate, while they absorb bile acids,
Hepatic immune cells
glucose, and glutamate. The cholangiocytes also play an impor­ The human liver is exposed to 1.5 L of blood every minute and
tant immunoregulatory role. They are in the first line of defense a massive load of harmless dietary and commensal antigens, to
against microbial components of the biliary tract, xenobiotics, which it must remain tolerant. The predominantly tolerogenic
and foreign antigens. The cholangiocytes tackle these chal­ role of the hepatic immune system is well known [52], but the
lenges by maintaining immunotolerance. They are equipped liver is also exposed to a variety of viruses, bacteria, parasites,
with pathogen recognition receptors (PRR), all members of the and metastatic tumor cells, so needs mechanisms to override
toll‐like receptors (TLR1–10), as well as related signaling mol­ immune tolerance. In addition, the liver’s native immune sys­
ecules. Cholangiocytes produce antibacterial products (e.g. tem has a major regulatory role in the repair of the liver after
defensins, lactoferrin, lysozyme, and transport IgA) into the cell injury and loss. The liver‐centered immune system is
lumen and express MHC class I and II molecules and antigen largely segregated from the rest of the body’s immune system
presenting cells [59]. It is not surprising, therefore, that the bil­ [66, 67]. The human liver is estimated to contain about 1010
iary tree is often affected by immunological disorders, such as lymphocytes of different phenotypes, located along sinusoids
primary sclerosing cholangitis, primary biliary cholangitis, and and in portal tracts [66]. It includes a major fraction of the
graft‐versus‐host disease. The cholangiocytes are slow turnover body’s innate (native) immune capacity, as well as a small com­
cells, but under special condition they can also participate in the ponent of its acquired (adaptive) immune capacity [66, 67]. The
regeneration of liver parenchyma (see later). major components of the hepatic immune system are innate
lymphocytes, which include a variety of T cells and non‐T cells
that are able to respond rapidly to conserved ligands. These
Endothelial cells cells do not express T cell receptor (TCR) antigens. This group
Endothelial cells of sinusoids comprise about 3% of the paren­ of immune cells includes NK cells, CD56+ T cells, natural killer
chymal mass/volume [15] and probably number about 3 × 1010 T (NKT) cells, γ/δ T cells, and mucosal associated invariant
in an adult human liver. Liver sinusoidal endothelial cells are T  (MAIT) cells. The liver contains multiple types of antigen
10 THE LIVER:  ONTOGENESIS OF THE LIVER

presenting cells such as hepatic myeloid dendritic cells (DCs), largest reservoir of vitamin A, hence their former name “­fat‐
plasmacytoid DCs, CD11+ DCs, and NK1.1+ cytotoxic DCs storing cells.” When the liver is injured the stellate cells trans­
[68]. In addition to the professional lymphocytes and DCs, sev­ differentiate into myofibroblasts and are the major producer of
eral populations of hepatic resident cells (e.g. Kupffer cells, extracellular matrix (ECM). HSCs are important sources of
sinusoidal endothelial cells, stellate cells, and cholangiocytes) cytokines and growth factors. This way they have impact on
are also important and active players of the liver‐centered the proliferation, ­differentiation, and morphogenesis of the
immune system. other hepatic cell types during liver development and regen­
eration [72].
Portal fibroblasts are spindle‐shaped mesenchymal cells in
Macrophages the periportal connective tissue [73]. They are distinct from
Macrophages are myeloid cells that are widely distributed HSCs in both distribution and phenotype. They do not store
throughout the tissues of mammalian organisms. The liver har­ vitamin A but express elastin and Thy‐1. Portal fibroblasts take
bors 80% of all macrophages of the body. In addition, it is also part in physiological ECM turnover and can contribute to the
patrolled by blood monocytes [69]. Hepatic macrophages can myofibroblast population in cholestatic liver injury.
be divided into two classes based on their origin [70]: resident Bone marrow‐derived mesenchymal stem cells can also dif­
macrophages and bone marrow‐derived macrophages. Resident ferentiate into myofibroblasts. The contribution of bone marrow
macrophages of the liver, traditionally termed Kupffer cells, are cells is well established in the fibrotic processes of kidney and
established during embryonic development from the yolk sac lung but it seems to be quite limited in the liver. Epithelial–­
and persist independent of blood monocytes. These cells self‐ mesenchymal transition (EMT) is another recently described
renew during homeostatic conditions. Bone marrow‐derived mechanism. Both hepatocytes and cholangiocytes can undergo
blood‐borne monocytes give rise to monocyte‐derived hepatic such phenotypic change in tissue culture, but lineage‐tracing
macrophages that are more characteristic of liver injury. Kupffer experiments in mice provide evidence against the role of EMT
cells are not optimally suited to migration, so hepatic injuries in myofibroblast ­generation in hepatic tissue in vivo [74, 75].
massively recruit blood monocytes to the liver. These resemble
Kupffer cells phenotypically, but they remain functionally
­different. Macrophages are highly versatile cells that play a sub­
stantial role in liver homeostasis, promoting and resolving ONTOGENESIS OF THE LIVER
inflammatory processes and fibrosis [69].
Liver macrophages are avidly phagocytic through C3 and The evolutionary steps that resulted in the emergence of the
Fc  receptors, clearing the sinusoidal blood of relatively large mammalian liver with its multiple types of functioning cells are
particulate materials, including bacteria and weakened cells obscure. To the extent that the ontogenesis of the liver mirrors
(worn‐out erythrocytes, dead or damaged hepatocytes, etc.) [63, its phylogenesis, the embryonic development of the mammalian
64]. Together with sinusoidal endothelial cells they form the liver suggests the way in which the aggregation of multiple
organism’s major system for removing worn‐out cells and pro­ types of cell into the hepatic parenchyma may have evolved
teins from perfusing blood. Activated macrophages produce (for details see Chapter 2).
many chemokines and cytokines that have a fundamental role in The development sequences of the liver in fish, birds, and
the implementation of the liver’s acute‐phase reaction, coordi­ mammals are similar [74–76], but endothelial cells do not
nating the responses of all the parenchymal cells to injury [67]. appear to direct the emergence of endodermal cells from the gut
during development of the liver in zebrafish [76]. Furthermore,
endothelial cells with scavenger activity are in the gills and kid­
Myofibroblasts neys of cartilaginous and bony fish, and not in the liver as in all
Myofibroblasts are not present in the normal liver, but several terrestrial animals [77]. The location of scavenger endothelial
cell types which are present physiologically can transform or be cells in the liver reflects a late step in the evolution of the mam­
activated or transdifferentiated into the phenotype, which is the malian liver. Nevertheless, the general pattern of expression of
major source of extracellular matrix components and plays a transcription factors and genes involved in liver development is
basic role in pathological processes of the liver [71]. conserved in all these species [78], suggesting a common tran­
Hepatic stellate cells (HSCs) [72] are liver‐resident mesen­ scriptional strategy for assembling the liver. Information on
chymal cells which play an important role in liver physiology when this strategy first emerged awaits further genetic analysis
and pathology. They are found in the space of Disse, between of gut appendages in chordate ancestors of vertebrates.
the sinusoidal endothelial cells and hepatocytes. This special
location makes them able to respond to numerous kinds of
Liver parenchymal repair
injury. In addition, by encircling the sinusoids they can func­
tion as pericytes and are thought to be the most important Three distinct processes have evolved to generate new hepato­
regulator of sinusoidal diameter and blood flow. They com­ cytes needed to meet both increased physiological functional
prise about 1.5% of the parenchymal volume/mass [15] and demands and to replace hepatocytes lost to trauma and/or toxicity.
are multifunctional (see Chapters 28 and 29). The embryonic These processes comprise either a temporary reactivation of cell
origin of stellate cells is still uncertain. Most likely they origi­ cycle transit in fully differentiated, mitotically quiescent hepato­
nate from the septum transversum‐derived mesothelial cells, cytes, or the generation of entirely new hepatocyte lineages from
but other options are still open. In the healthy liver they are the adult liver stem cells (see Chapters 36 and 38).
1:  Organizational Principles of the Liver 11

The most direct and rapid parenchymal augmentation/ which are cleaved into one‐hepatocyte‐wide plates by signaling
replacement processes involve the induction of hepatocyte rep­ from and separation by endothelial and stellate cells [90, 91].
lication in the absence of a preceding increase in hepatocyte Eventually, the remaining lobes increase exclusively by the
death, often associated with increased hepatic functional enlargement of preexistent hepatic lobules, contrary to the
demand due to physiological need [78, 79]. Hyperplasia of ­physiological liver growth in young animals, when new lobules
hepatocytes by this mechanism enlarges the parenchymal mass are formed [92].
and increases hepatocyte functional capacity. This process is Although known regulatory mechanisms drive the reparative
regulated by the binding of ligands to hepatocyte nuclear recep­ process, the mechanism that “triggers” the onset of repair after
tors, nearly 50 of which have been identified [79, 80]. Nuclear loss of liver tissue is still obscure. Since the liver vasculature
receptors are transcription factors that, when bound to ligands, must accept the entire portal blood volume, it has long been
directly upregulate the combination of genes required to drive suspected that the trigger may be the massive increase in portal
hepatocytes through the cell cycle [79, 80]. Several ligands for blood flow per unit of residual mass that follows loss of liver
nuclear receptors (termed “primary hepatocyte mitogens”), tissue [93]. Increased portal blood flow and pressure cause shear
including adrenal corticoids, bile acids, sex steroids, thyroid stress in sinusoids [94], which produces a burst of nitric oxide
hormone, peroxisome proliferators, and 9‐cis,cis‐retinoic acid, and prostaglandin production by sinusoidal endothelial cells,
directly stimulate the proliferation of hepatocytes and increase possibly providing the molecular trigger [95, 96]. Alternatively
liver mass after binding to nuclear receptors [80]. Although it (or in concert), early activation of the nuclear receptor mecha­
would appear that endothelial cells would be needed to support nism of hepatocyte proliferation may function as a trigger [96],
the additional hepatocytes, no documentation of coordinate and it is possible that multiple alterations in the physiological
endothelial cell proliferation has been presented; it is, however, status of the liver remaining after tissue loss may converge to
possible that new endothelial cells could be derived from the produce a “mass action” trigger.
bone marrow. If the hepatocytes are compromised, there are alternative
Next in process complexity and speed of response is the mechanisms of liver regeneration. Enlargement or hypertrophy
replacement of the diverse liver parenchyma by the sequential of hepatocytes can compensate for the lost parenchyma [97], but
proliferation of all of the component cells (hepatocytes, chol­ this response usually provides just a transient solution.
angiocytes, endothelial cells, macrophages, stellate cells, and There has been much debate about the participation of stem or
immunocytes), and the merging of the new cells into a tissue progenitor cells in liver regeneration. A peculiar cell population,
that closely resembles the functional units of the undamaged named after the shape of their nuclei as “oval cells,” were
liver [78]. This process, which can replace up to 70% of the observed in hepatocarcinogenesis experiments in rodents. Similar
parenchymal mass in mammals, is often called “liver regenera­ cells have been described in several other species and their emer­
tion,” although this is a misnomer since in mammals the part of gence has been named “ductular reaction.” There is convincing
the liver removed surgically does not “regenerate” in the way evidence that these cells can replace the lost liver parenchyma
that body parts in certain lower animal species do. Instead, the [98] and behave as the amplification compartment of hepatic
liver, after resection, is enlarged by the expansion of remaining stem cells. Several candidates for hepatic stem cells are known,
units (lobes), in a biological process defined as “compensatory but most data indicate that the terminal segment of the biliary
hyperplasia.” In contrast to liver repair in mammals, liver repair system, the canals of Hering, harbor the adult hepatic stem cells.
after partial hepatectomy in fish most intensively involves cells Lineage‐tracing experiments in rats [54, 55, 57, 99, 100] and
at the resection margin [81, 82] and may culminate in the zebrafish [101] demonstrated the regeneration of hepatocytes
regrowth (regeneration) of the resected tissue [81]. from biliary stem cells. The application of the cre/lox lineage
The cell proliferation phase of this reparative process in tracing in mice did not support this prevailing model, because
mammals has been subjected to intensive kinetic and regulatory no biliary cell‐derived hepatocytes were observed, although the
analyses (see Chapter  45 and references therein). After tissue hepatocytic origins of biliary cells and cholangiocarcinomas
loss, residual hepatocytes are activated to proliferate within few were demonstrated [102]. However, eventually hepatic progeni­
hours. Hepatocyte proliferation begins at the portal ends of tor cells of biliary origin with liver repopulation capacity were
plates  [83], and successive waves of hepatocyte proliferation shown in mice following complete blockage of hepatocyte pro­
ultimately involve virtually all residual hepatocytes [83]. liferation [103]. At present there seem to be general agreement
Hepatocyte replacement is followed sequentially by prolifera­ [104] that both hepatocytes and cholangiocytes (or their sub­
tion of sinusoidal endothelial cells and macrophages [83, 84], populations) are able to behave as stem cells, and under specific
and the other cells of the parenchymal matrix. To the extent that conditions are capable of regenerating both epithelial cell
it has been elucidated (see Chapter 45), regulation of hepatocyte ­compartments of the liver. The capacity of these highly differen­
proliferation is regulated by a complex mixture of cytokines and tiated cells is also referred to as “plasticity” [105] but this is
growth factors [85]. Most of the regulatory molecules are pro­ mostly a debate about terminology – how we should refer to a
duced by various liver cells or are released from storage sites peculiar biological reaction. Initial observations indicate that
within the liver [85], and many are components of the acute‐ these “back‐up” stem cells, which support regenerative pro­
phase reaction [86] and other elements of the liver’s native cesses in rat and human, are organized along the branches of the
immune system [87–89]. The less completely analyzed remod­ portal vein [106, 107] and are regulated by elements of hepatic
eling phase primarily involves endothelial cells and likely the immunomodulation centered on the acute‐phase reaction [98,
other cells of the liver parenchyma. For example, proliferating 108], similar to the organization of liver architecture during
hepatocytes initially form focal multicellular clumps [90, 91], embryonic development.
12 THE LIVER:  REFERENCES

Although the subject of intensive scrutiny recently, there is 22. Bloch, E.H. The termination of hepatic arterioles and functional unit of the
no substantial evidence that hematopoietic stem cells or mesen­ liver as determined by microscopy of the living organ. Ann NY Acad Sci,
1970;170:78–87
chymal components of the liver are a significant source for the 23. Kiernan, F. The anatomy and physiology of the liver. Trans R Soc Lond,
generation of hepatocytes or biliary epithelial cells in either 1833;123:711–70.
humans or experimental animals [47]. This situation contrasts 24. Rappaport, A.M. The microcirculatory hepatic unit. Microvasc Res,
with the replenishment from hematopoietic sources of other 1973;6:212–28.
cells of the liver parenchyma [47]. 25. Ekataksin, W. and Wake, K. Liver units in three dimensions: I. Organization
of argyrophil connective tissue skeleton in porcine liver with particular refer­
ence to the “compound hepatic lobule.” Am J Anat, 1991;191(2):113–53.
26. Elias, H. and Sherrick, J.C. Morphology of the Liver. Academic Press,
New York, 1969.
REFERENCES 27. Lautt, W.W. and Greenway, C.V. Conceptual view of the hepatic vascular
bed. Hepatology, 1987;7:952–63.
1. Elias, H. A re‐examination of the mammalian liver, I. Parenchymal 28. Nakata, K., Leong, G.F., and Brauer, R.W. Direct measurement of blood
­architecture. Am J Anat, 1949;84:311–33. pressure in the minute vessels of the liver. Am J Physiol, 1960;199:1181–8.
2. Couinand, C. Liver anatomy: portal (and suprahepatic) or biliary segmentation. 29. McKuskey, R.S. Morphological mechanisms for regulating blood flow
Dig Surg, 1999;16:459–67. through hepatic sinusoids. Liver, 2000;20:3–7.
3. Majno, P., Mentha, G., Toso, C. et al. Anatomy of the liver: an outline with 30. Hellerbrand, C. Hepatic stellate cells  –  the pericytes in the liver. Pflugers
three levels of complexity: a further step towards tailored territorial liver Arch Eur J Physiol, 2013;465:775–8.
resections. J Hepatol, 2013;60:654–62. 31. Reynaert, H., Thompson, M.G., Thomas, T., and Geerts, A. Hepatic stellate
4. Elias, H. A re‐examination of the mammalian liver, II. The hepatic lobule cells: role in microcirculation and pathophysiology of portal hypertension.
and  its relation to the vascular and biliary systems. Am J Anat, 1949;84: Gut, 2002;50:571–81.
379–456. 32. Grisham, J.W. A morphologic study of deoxyribonucleic acid synthesis and
5. Rappaport, A.M., Borowy, Z.J., Lougheed, W.M. et al. Subdivision of hex­ cell proliferation in regenerating rat liver; autoradiography with thymidine
agonal liver lobules into a structural and functional unit. Role in hepatic H3. Cancer Res, 1962;22:842–49.
physiology and pathology. Anat Rec, 1954;119:11–34. 33. Kmeic, Z. Cooperation of liver cells in health and disease. Adv Anat Embryol
6. Matsumoto, T., Komori, R., Magara, M. et al. A study of the normal structure Cell Biol, 2001;161:1–151.
of the human liver, with special reference to its angioarchitecture. Jikeikai 34. Weibel, E.R., Stäubli, W., Gnägi, H.R., and Hess, F.A. Correlated
Med J, 1979;26:1–40. ­morphometric and biochemical studies on the liver cell. 1. Morphometric
7. Kotton, D.W., Fabian, A.J., and Mulligan, R.C. A novel cell population in model, stereologic methods, and normal morphometric data for rat liver.
adult liver with potent hematopoietic reconstitution activity. Blood, J Cell Biol, 1969;42:68–90.
2005;106:1574–80. 35. Gkretsi, V., Apte, U., Mars W.M. et al. Liver specific ablation of integrin‐
8. Mitra, S.K. The terminal distribution of the hepatic artery with special refer­ linked kinase in mice results in abnormal histology, enhanced cell prolifera­
ence to arterio‐portal anastomosis. J Anat, 1966;100:651–63. tion, and hepatomegaly. Hepatology 2008;48:1932–41.
9. Kardon, R.H. and Kessel, R.G. Three‐dimensional organization of the 36. Parviz, F., Matullo, C., Garrison, W.D. et  al. Hepatocyte nuclear factor
hepatic microcirculation in the rodent liver as observed in the scanning elec­ 4 alpha controls the development of a hepatic epithelium and liver morpho­
tron microscopy of corrosion casts. Gastroenterology, 1980;79:72–81. genesis. Nat Genet, 2003;34:292–6.
10. Yamamoto, K., Sherman, I., Phillips M.J. et al. Three‐dimensional observations 37. Woods, A., Heslegrave, A.J., Muckett, P.J. et al. LKB1 is required for hepatic
of the hepatic arterial terminations in rat, hamster and human liver by scanning bile acid transport and canalicular membrane integrity in mice. Biochem J,
electron microscopy of microvascular casts. Hepatology, 1985;5:452–6. 2011;434:49–60.
11. Watanabe, Y., Püschel, G.P., Gardemann, A. and Jungermann, K. Presinusoidal 38. Cullinane, A.R., Straatman‐Iwanowska, A., Zaucker, A. et  al. Mutations in
and proximal intrasinusoidal confluence of hepatic artery and portal vein in VIPAR cause an arthrogryposis, renal dysfunction and cholestasis syndrome
rat liver: functional evidence by orthograde and retrograde bivascular per­ phenotype with defects in epithelial polarization. Nat Genet, 2010;42:303–12.
fusion. Hepatology, 1994;19:1198–207. 39. Cheung, I.D., Bagnat, M., Ma, T.P. et al. Regulation of intrahepatic bile duct
12. Wisse, E., DeZanger, R.B., Charels, K. et al. The liver sieve: considerations morphogenesis by Claudin 15‐like b. Dev Biol, 2012;361, 68–78.
concerning the structure and function of endothelial fenestrae, the sinusoidal 40. Celton‐Morizur, S., Merlen, G., Conton, D. et al. The insulin/Akt pathway
wall and the space of Disse. Hepatology, 1985;5:683–92. controls a specific cell division program that leads to generation of binucle­
13. Goresky, C.A. A linear method for determining the liver sinusoidal and ated tetraploid liver cells in rodents. J Clin Invest, 2009;122:3307–15.
extravascular volumes. J Physiol, 1963;204:626–40. 41. Duncan, A.W., Hanlon Newell, A.E., Sith, L. et  al. Frequent aneuploidy
14. Grisham, J.W., Nopanitaya, W., Compagno, J., and Nagel, A.E.H. Scanning among normal human hepatocytes. Gastroenterology, 2012;142:25–8.
electron microscopy of normal rat liver: the surface structure of its cells and 42. Gupta, S. Hepatic polyploidy and liver growth control. Semin Cancer Biol,
tissue components. Am J Anat, 1975;144:295–322. 2000;10:161–71.
15. Blouin, A., Bolender, R.P., and Weibel, E.R. Distribution of organelles and 43. Traub, O., Look, J., Dermietzel, R. et  al. Comparative characterization of
membranes between hepatocytes and nonhepatocytes in the rat liver paren­ 21‐kD and 26‐kD junction proteins in murine liver and cultured hepatocytes.
chyma. A stereological study. J Cell Biol, 1977;72:441–55. J Cell Biol, 1989;108:1039–51.
16. Miller, D.L., Zanolli, C.S., and Gumucio, J.J. Quantitative morphology of 44. Jungermann, K. and Katz, N. Functional specialization of different hepato­
the sinusoids of the hepatic acinus. Gastroenterology, 1979;76:965–69. cyte populations. Physiol Rev, 1989;69:708–64.
17. Grisham, J.W. and Nopanitaya, W. Scanning electron microscopy of casts of 45. Gougelet, A., Torre, C., Veber, P. et al. T cell factor 4 and β‐catenin chromatin
hepatic microvessels, in Hepatic Circulation in Health and Disease (ed. W. occupancies pattern zonal liver metabolism. Hepatology, 2014;59:2344–57.
Laut), Raven Press, New York, 1981, pp. 87–107. 46. Jungermann, K. and Kietzmann, T. Oxygen: modulator of metabolic zona­
18. Dezső, K., Paku, S., Papp, V. et al. Architectural and immunohistochemical tion and disease of the liver. Hepatology, 2000;31:255–60.
characterization of biliary ductules in normal human liver. Stem Cells Dev, 47. Thorgeirsson, S.S. and Grisham, J.W. Hematopoietic cells as hepatocyte
2009;18:1417–22. stem cells: a critical review of the evidence. Hepatology, 2006;43:2–8.
19. Theise, N.D., Saxena, R., Portmann, B.C. et al. The canals of Hering and 48. Zajicek, G., Oren, R., and Weinreb, M. Jr. The streaming liver. Liver,
hepatic stem cells in humans. Hepatology 1999;30:1425–33. 1985;5:293–300.
20. Teutsch, H.F., Schuerfeld, D., and Groezinger, E. Three‐dimensional 49. Wang, B., Zhao, L., Fish, M. et al. Self‐renewing diploid Axin 2+ cells fuel
­reconstruction of parenchymal units in the liver of the rat. Hepatology, homeostatic renewal of the liver. Nature, 2015;524:180–5.
1999;29:494–505. 50. Font‐Burgada, J., Shalapour, S., Ramaswamy, S. et  al. Hybrid periportal
21. Teutsch, H.F. The modular microarchitecture of human liver. Hepatology, hepatocytes regenerate the injured liver without giving rise to cancer. Cell,
2005;42:317–25. 2015;162:766–79.
1:  Organizational Principles of the Liver 13

51. Allison, M.R. Hepatocytes come out of left field. Hepatology, 2016;63: 81. Zhao, R. and Duncan, S.A. Embryonic development of the liver.
1041.–2. Hepatology, 2005;41:956–67.
52. Jennem, C.N. and Kubes, P. Immune surveillance by the liver. Nat Immunol, 82. Zaret, K.S. Regulatory phases of early liver development: paradigms of
2013;14:996–1006. organogenesis. Nat Rev Genet, 2007;3:499–512.
53. Saxena, R. and Theise, N. Canals of Hering: recent insights and current 83. Sadler, K.L., Krahn, K.N., Gaur, N.A., and Ukomadu, C. Liver growth in
knowledge. Semin Liver Dis, 2004;24:43–8. the embryo and during liver regeneration in zebrafish requires the cell cycle
54. Thorgeirsson, S.S. and Grisham, J.W. Liver stem cells, in Stem Cells (ed. regulator, uhrf1. Proc Natl Acad Sci U S A, 2007;104:1570–5.
C.S. Potten), Academic Press, London, 1997, pp. 233–85. 84. Okihiro, M.S. and Hinton, D.E. Partial hepatectomy and bile duct ligation
55. Kuwahara, R., Kofman, A.V., Landis, C.S. et al. The hepatic stem cell niche: in rainbow trout (Oncorhynchus mykiss): histologic, immunohistochemical
identification fy label‐retaining assay. Hepatology, 2008;47:1994–2002. and enzyme histochemical characterization of hepatic regeneration and
56. Benedetti, A., Bassotti, C., Rapno, K. et al. A morphometric study of the epi­ hyperplasia. Toxicol Pathol, 2000;28:342–335.
thelium lining the rat intrahepatic biliary tree. J Hepatol, 1996;24:335–42. 85. Michalopoulos, G.K. Hepatostat: liver regeneration and normal tissue
57. Bing, Q., Masyuk, T.V., Muff, M.A. et al. Isolation and characterization of maintenance. Hepatology 2017;65:1384–92.
cholangiocyte primary cilia. Am J Physiol Gastrointest Liver Physiol, 86. Fulop, A.K., Pocsik, E., Brozik, E. et al. Hepatic regeneration induces tran­
2006;291:G500–9. sient acute phase reaction: systemic elevation of acute phase reactants and
58. Alpini, G., Ulrich, C., Roberts, S. et al. Molecular and functional heteroge­ soluble cytokine receptors. Cell Biol Int, 2001;25:585–92.
neity of cholangiocytes from rat liver after bile duct ligation. Am J Physiol 87. Fausto, N. Involvement of innate immune system in liver regeneration and
Gastrointest Liver Physiol, 1997;272:G289–97. injury. J Hepatol, 2006;45,347–9.
59. Zhang, H. How biliary tree maintains immune tolerance? BBA Mol Basis 88. Dong, Z., Wei, H., Sun, R., and Tian, Z. The role of innate cells in liver
Dis, 2018;1864:1367–73. injury and regeneration. Cell Mol Immunol, 2007;4:241–52.
60. Braet, F. and Wisse, E. Structural and functional aspects of liver sinusoidal 89. Preziasi, M.E. and Monga, S.P. Update on the mechanisms of liver regen­
endothelial cell fenestrae: a review. Comp Hepatol, 2002;1:1–10. eration. Semin Liver Dis, 2017;37:141–51.
61. Sorensen, K.K., Simon‐Santamaria, J., McCuskey, R.S. et al. Liver sinusoi­ 90. Martinez‐Hernandez, A. and Amenta, P. The extracellular matrix in hepatic
dal endothelial cells. Compr Physiol, 2001;5:1751–74. regeneration. FASEB J, 1995;9:1401–10.
62. McCuskey, R.S. and Reilly, F.D. Hepatic microvasculature: dynamic struc­ 91. Wack, K.E., Ross, M.A., Zegara, V. et al. Sinusoidal ultrastructure during
ture and its regulation. Semin Liver Dis, 1993;13:1–12. revascularization of regenerating liver. Hepatology, 2001;33:363–78.
63. Smedsrød, B., DeBleser, P.J., Braet, F. et al. Cell biology of endothelial and 92. Papp, V., Dezso, K., Laszlo, V. et al. Architectural changes during regenera­
Kupffer cells. Gut, 1994;35:1509–16 tive and ontogenic growth in the rat. Liver Transplant, 2009;18:177–83.
64. Smedsrød, B. Clearance function of scavenger endothelial cells. Comparative 93. Sato, Y., Kayama, S., Tsukada, K., and Hatakeyama, K. Acute portal hyper­
Hepatol, 2004;3(Suppl 1):S22. tension reflecting shear stress as a trigger to liver regeneration following
65. Wang, L., Wang, X., Wang, L. et  al. Hepatic vascular endothelial growth partial hepatectomy. Surg Today (Jpn J Surg), 1997;27:518–26.
factor regulates recruitment of rat liver sinusoidal endothelial cell progenitor 94. Schoen, J.M., Wang, H.H., Minuk, G.Y., and Lautt, W.W. Shear stress‐
cells. Gastroenterology, 2012;143:1555–63. induced nitric oxide release triggers the liver regeneration cascade. Nitric
66. Doherty, D.G. and O’Farrelly, C. Innate and adaptive lymphoid cells in the Oxide, 2001;5:453–64.
human liver. Immunol Rev, 2000;174:5–20. 95. Schoen‐Smith, J.M. and Lautt, W.W. The role of prostaglandins in trigger­
67. Parker, G.A. and Picut, C.A. Liver immunobiology. Toxicol Pathol, ing the liver regeneration cascade. Nitric Oxide, 2005;23:111–17.
2005;33:52–62. 96. Huang, W., Ma, K., Zhang, J. et al. Nuclear receptor‐dependent bile acid sign­
68. Doherty, D.G. Immunity, tolerance and autoimmunity in the liver: a compre­ aling is required for normal liver regeneration. Science, 2006;312:233–6.
hensive review. J Autoimmun, 2016;66:60–75. 97. Matot, I. and Nochmansson, N. Impaired liver regeneration after hepatec­
69. Ju, C. and Tacke, F. Hepatic macrophages in homeostasis and liver disease: tomy and bleeding is associated with shift from hepatocyte proliferation to
from pathogenesis to novel therapeutic strategies. Cell Mol Immunol, hypertrophy. FASEB J, 2017;31:5283–95.
2016;13:316–27. 98. Thorgeirsson, S.S., Factor, V.M., and Grisham, J.W. Early activation and
70. Gomez, P.E., Klapproth, K., Schulz, C. et al. Tissue‐resident macrophages expansion of hepatic stem cells, in Handbook of Stem Cells, Introduction to
originate from yolk‐sac derived erythro‐myeloid progenitors. Nature, Adult and Fetal Stem Cells, vol. 3 (eds. R. Lanza, H. Blau, D. Melton et al.),
2015;518:547–51. Academic Press, San Diego, 2004, pp. 497–512.
71. Lee, U.E. and Friedman, S.L. Mechanisms of hepatic fibrogenesis. Best 99. Evarts, R.P., Nagy, P., Marsden, E., and Thorgeirsson, S.S. A precursor‐­
Pract Res Clin Gastroenterol, 2011;25:195–206. product relationship exists between oval cells and hepatocytes in rat liver.
72. Yin, C., Evason, K. J., Asashina, K. et al. Hepatic stellate cells in liver devel­ Carcinogenesis, 1987;8:1737–40.
opment, regeneration and cancer. J Clin Invest, 2013;123:1902–10. 100. Paku, S, Schnur, J., Nagy, P et  al. Origin and structural evolution of the
73. Wells, R.G. The portal fibroblast  –  not just a poor man’s stellate cell. early proliferating oval cells in rat liver. Am J Pathol, 2001;158:1313–23.
Gastroenterology, 2014;147:41–7. 101. Choi, T.Y., Ninov, N., Stainer, D.Y. et al. Extensive conversion of hepatic
74. Elias, H. Origin and early development of the liver in various vertebrates. biliary epithelial cells to hepatocytes after near total loss of hepatocytes in
Acta Hepatol, 1955;3:1–567. zebra fish. Gastroenterology, 2014;146:776–88.
75. Le Douarin, N.M. An experimental analysis of liver development. Med Biol, 102. Schaub, J.R., Malato, Y., Gormond, C. et al. Evidence against a stem cell
1975;53:427–55. origin of new hepatocytes in a common mouse model of chronic liver
76. Field, H.A., Ober, E.A., Roeser, T., and Stainier, D.Y.R. Formation of the injury. Cell Rep, 2014;8:933–939.
digestive system in zebrafish. I. Liver morphogenesis. Dev Biol, 2003;253: 103. Lu, W.Y., Bird, T.G., Boulter, L. et al. Hepatic progenitor cells of biliary
279.–90. origin with liver repopulation capacity. Nature Cell Biol, 2015;17:971–83.
77. Seternes, T., Sørensen, K., and Smedsrød, B. Scavenger endothelial cells of 104. Michalopoulos, G.K. and Khan, Z. Liver stem cells: experimental findings
vertebrates: a nonperipheral leukocyte system for high capacity elimination and implications for human disease. Gastroenterology, 2015;149:876–82.
of waste macromolecules. Proc Natl Acad Sci U S A, 2002;99:7594–97. 105. Kopp, J.L., Grompe, M., and Sander, M. Stem cells versus plasticity in liver
78. Columbano, A. and Ledda‐Columbano, G.M. Mitogenesis by ligands of and pancreas regeneration. Nature Cell Biol, 2016;18:238–44.
nuclear receptors: an attractive model for the study of molecular mechanisms 106. Dezso, K., Papp, V., Bugyik, E. et al. Structural analysis of oval‐cell‐medi­
implicated in liver growth. Cell Death Differ, 2003;10:S19–21. ated liver regeneration in rat. Hepatology, 2012;56:1457–67.
79. Zaret, K.S. Regulatory phases of early liver development: paradigms of 107. Dezso, K., Rokusz, A., Bugyik, E. et al. Human liver regeneration in advanced
organogenesis. Nat Rev Genet, 2007;3:499–512. cirrhosis is organized by the portal tree. J Hepatol, 2017;66:778–86.
80. Nagy, L. and Schwabe, J.W. Mechanism of the nuclear receptor molecular 108. Cast, A.E., Walter, T.J., and Huppert S.S. Vascular patterning sets the stage
switch. Trends Biochem Sci, 2004;29:317–24. for macro and micro hepatic architecture. Dev Dynam, 2015;244:497–506.
Embryonic Development
2 of the Liver
Kenneth S. Zaret1, Roque Bort2, and Stephen A. Duncan3
1
Institute for Regenerative Medicine, Department of Cell and Developmental Biology, Perelman School of
Medicine, University of Pennsylvania, Smilow Center for Translational Research, Philadelphia, PA, USA
2
Instituto de Investigación Sanitaria La Fe (IIS La Fe), Unidad de Hepatología experimental, València, Spain
3
Department of Regenerative Medicine and Cell Biology, Medical University of South Carolina, Charleston, SC, USA

INTRODUCTION ACQUISITION OF HEPATIC


COMPETENCE WITHIN
The liver is one of the first organs to develop in the embryo and THE ENDODERM
it rapidly becomes one of the largest organs in the fetus. The
most essential function of the mammalian fetal liver is to pro- The liver, lung, pancreas, thyroid, and gastrointestinal tract are
vide a site for hematopoiesis. The early dependence of the fetus derived from the anterior‐ventral definitive endoderm, which is
on its own blood cell supply makes embryonic liver growth and one of the three germ layers that arise during gastrulation. Initially,
viability a sensitive phenotypic indicator for gene inactivation the endoderm is an epithelial sheet that lines the ventral surface of
studies. From an experimental perspective, the large size and the embryo. Infolding of the sheet at the anterior and posterior of the
small number of cell types in the developing liver make it easy embryo generates the foregut and hindgut (Figure  2.1). When
to study, and new insights have emerged from recent work on these morphogenetic movements reach the middle of the embryo, the
genetically modified mice, explants of embryonic tissues, pluri- gut tube closes off. During gut tube formation, different tissues
potent stem cell differentiation, and other vertebrate models are specified along the anterior–posterior and dorsal–ventral axes
such as zebrafish and frogs. Much is being learned about how of the embryo, with the liver arising from the prospective ventral
gene function is orchestrated to control tissue morphogenesis, endoderm domain of the foregut (Figure  2.1). Are there “pre‐
helping to establish liver development as a paradigm for the patterns” or local domains of endoderm that are competent to
genesis of other gut‐derived tissues. Understanding the mecha- differentiate into the liver? LeDouarin demonstrated, using heter-
nisms that govern liver development should also provide insight otypic grafts of quail embryo segments into recipient chick
into future therapies for liver diseases. Examples of such appli- embryos, that only the prospective anterior‐ventral domain of
cations include activating stem cells in the adult liver, replenish- endoderm had the capacity to develop into the liver [10, 11].
ing diseased livers with cells from pluripotent stem cells, High‐resolution lineage tracing in mouse embryos has revealed
transdifferentiating cells from different organs, and reconstitut- that distinct domains within the endoderm contribute to the devel-
ing proper liver morphology and function. oping liver [12]. The majority of cells derive from two bilateral
This review will describe the early stages of liver develop- patches within the lateral endoderm, and a smaller population
ment from the initial specification to hematopoietic cell inva- derives from a small clutch of cells at the ventral midline of the
sion, which covers the competence of progenitor cells to become anterior endoderm. Morphogenic movements ensure that these
hepatocytes, the formation of the liver bud, and the early mor- populations converge during the formation of the liver bud.
phogenesis and differentiation of the liver. The review will also More recent studies with mouse embryos and more sensitive
present the key effectors of hepatocyte maturation and discuss assays of gene expression revealed that while the prospective
the control of liver size and regeneration in the embryo and in dorsal endoderm (Figure 2.1) does not normally activate liver
the adult. The reader is referred to other reviews for summaries genes or become liver, in a tissue explant assay the dorsal endo-
of the mid‐ and late‐fetal stages of liver development [1–9]. derm cells could initiate liver gene expression when isolated

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
2:  Embryonic Development of the Liver 15

endoderm
cardiac head
mesoderm alb1–
(FGF source) –10 kb X

septum GATA FoxA


transversum
mesenchyme fore- somites
(BMP source) gut hind-
gut liver bud
ventral-anterior
endoderm alb1+
(suppression of Wnt
allows hepatic dorsal-posterior endoderm
potential) (Wnt signaling suppresses
hepatic potential)
C/EBP FoxA eY FoxA NF1
Figure 2.1  Parasagittal diagram of a mammalian embryo at the time Figure 2.2  Transcriptional competence factors in the endoderm.
of hepatic specification. Relevant tissues and signaling molecules are DNA binding sites for GATA and FOXA transcription factors are occu-
indicated. The time of development corresponds to about 8.25 days’ pied on the alb1 gene prior to alb1 expression or hepatic commitment.
gestation in the mouse and about three weeks in the human. During hepatic specification, other transcription factors bind DNA at
adjacent sites and the albumin gene becomes active.

from its adjacent mesodermal tissues [13]. The molecular


mechanism underlying these transplantation studies has been other transcription factors become occupied at the enhancer and
deciphered using Xenopus (frog) embryos, and similar mecha- alb1 becomes active (Figure 2.2).
nisms likely operate in mammalian embryos. It was found that Genetic experiments have confirmed the essential role of
due to the anterior expression of Wnt inhibitors such as Dkk, FOXA and GATA factors in hepatogenesis. A transgenic mouse
sFrp‐1 or sFrp‐5, the function of the Wnt/β‐catenin signaling with a conditional ablation of both FOXA1 and FOXA2 in the
pathway is repressed upon gastrulation in the anterior endo- foregut endoderm completely lacked the induction of hepatic
derm [14], but it is active posteriorly, where the Wnt inhibitors markers of specification, such as alb1, afp, or ttr [35]. Similarly,
are absent. Wnt downregulation in the anterior endoderm was while the liver is specified in GATA6−/− or GATA4−/− mouse
shown to be crucial for liver and pancreas specification; in fact, embryos (though subsequent development is blocked), a double
posterior endoderm that was forced to activate GSK‐3β, a Wnt knockout of both genes in zebrafish revealed a complete failure in
inhibitor, possessed hepatogenic competence [15]. Wnt signal- liver induction [36]. The role of GATA factors in controlling endo-
ing induces the transcription factor Vent in the endoderm, dermal fate appears to be conserved in humans. Several groups
which in turn represses the homeobox gene Hex that is required have shown that depletion of GATA6 affects the differentiation of
in the endoderm for liver and pancreas development [16–21]. endoderm from human induced pluripotent stem cells (iPSCs)
In summary, Wnt repression in the anterior endoderm allows [37–39]. These findings provide genetic evidence supporting the
liver and pancreas specification, whereas active Wnt signaling concept that FOXA and GATA factors serve as “pioneer factors”
in the posterior endoderm suppresses those fates. These find- in the endoderm, being among the first to bind a target gene in
ings provide a molecular basis to help explain the patterning of development and endow competence for the gene to be activated,
the endoderm. under the influence of cell‐type inductive signals.
Another approach to understanding how the endoderm gains
the competence to develop into liver is to identify the regulatory
transcription factors that directly enable the process. Regulatory
factors that are known to be important for liver differentiation FROM DEFINITIVE TO HEPATIC
are expressed in prehepatic endoderm, prior to hepatic specifi- ENDODERM
cation. Such factors could operate by helping to open up chro-
matin structure for genes that need to be transcribed during liver Once the hepatogenic competence in the definitive endoderm is
differentiation [22]. Transcription factors that are expressed in acquired, what causes the liver to be specified from the ventral
the prehepatic domain of ventral foregut endoderm, and later in foregut endoderm? The mesoderm secretes patterning signals
the liver, include FOXA1, FOXA2 [23–26], GATA4, and that instruct the differentiation of the underlying endoderm.
GATA6 [27–30]. The roles of these transcription factors can be Several lines of research conducted in chicken and mouse dem-
discerned by understanding how the factors function in a chro- onstrated that cardiac mesoderm helps instruct the underlying
matin context. DNA binding by both FOXA and GATA factors ventral foregut endoderm to become hepatic endoderm
is required for the activity of a transcriptional enhancer of alb1, (Figure  2.1). In vitro culture of mouse embryonic explants
the gene encoding serum albumin [31, 32]; alb1 is one of the together with the use of zebrafish and Xenopus embryos have
earliest genes to be activated in hepatic development [13, 33]. identified some of the relevant molecular signals.
An analysis of the DNA binding sites for FOXA2 and GATA There are over 20 genes for fibroblast growth factors (FGFs)
[34] showed that both sites are occupied on the alb1 enhancer in and 4 genes for FGF receptors [40, 41]. Each of the FGF
the endoderm (Figure  2.2), prior to alb1 transcriptional receptors has different binding specificities for FGFs, and
­
­activation or hepatic commitment [13, 31]. Once hepatic speci- the existence of multiple spliced isoforms of the receptors pro-
fication has occurred, adjacent binding sites for a variety of vides further complexity to the receptor–ligand relationships.
16 THE LIVER:  FROM HEPATIC ENDODERM TO LIVER BUD

Cardiogenic mesoderm expresses FGF1, FGF2, FGF8, and endoderm and thus the number of cells that become specified to
FGF10 in the period prior to hepatic induction in the endoderm. liver, instead of pancreas [57].
In an embryonic tissue explant system, purified FGF1 or FGF2 The role of Wnt signaling during hepatic induction appears
can efficiently activate liver gene expression in ventral foregut dynamic. The FGF‐BMP induction of the liver occurs when
explants where cardiogenic mesoderm has been removed, and Wnt signaling is being suppressed in the foregut. But shortly
an FGF antagonist can inhibit liver gene induction in foregut after the induction of the hepatic program in the endoderm, Wnt
explants where cardiogenic mesoderm has been retained [42]. signaling appears to be required for further outgrowth of the
FGF binding induces tyrosine phosphorylation activity by the endoderm into a liver bud [15]. In zebrafish, expression of
cytoplasmic domain of the receptors, resulting in activation of Wnt2b in the lateral plate mesoderm, acting through the β‐catenin
mitogen‐activated protein kinase (MAPK) signaling pathways canonical pathway, appears essential for liver specification in
within cells [41, 43]. A combination of in vivo‐genetic, whole‐ the endoderm and bud induction [58]. Differences in the devel-
embryo culture and tissue explant approaches revealed that a opment of the liver in fish versus amniotes may explain an
transient activation of the MAPK pathway by FGF signaling in ­earlier positive role for Wnt signaling in hepatic induction [5].
foregut endoderm explants is necessary for the initiation and Like most organs, the liver is asymmetrically positioned with
stabilization of the hepatic program [44]. Although the PI3K/ the developing embryo. Patients with heterotaxy syndrome
AKT pathway is activated in the endoderm shortly after the (isomerism), in which organ asymmetry is disrupted, commonly
MAPK pathway, PI3K/AKT pathway activation appears not to have defects associated with liver function including biliary
be downstream of FGF signaling and the pathways do not cross- atresia [59]. The molecular basis for the asymmetric nature of
regulate in the hepatic endoderm [44]. Ultimately, the action of the liver has only recently been explored and our understanding
FGF signaling must activate gene expression programs that remains rudimentary. However, studies in zebrafish have found
drive hepatic fate. The differentiation of human iPSCs to a that communication between EphrinB1 and EphB3 are crucial
hepatic fate has allowed researchers to identify the immediate to coordinate the movement of the hepatic endoderm with the
targets of FGF signaling [45]. Genes that are directly activated adjacent lateral plate mesoderm [60]. EphrinB1 controls hepa-
during the endoderm‐to‐hepatic transition include several tran- toblast migration while EphB3 acts in the mesoderm to repel the
scription factors, growth factors, and signaling molecules. hepatoblasts. By working together these proteins control the
Interestingly, one of the direct targets of FGF encodes NKD1, directional migration of the hepatoblasts that is required to cor-
which is a repressor of WNT signaling. When NKD1 was rectly position the liver.
depleted, differentiation of the endoderm to a hepatic fate was
inhibited [45]. These findings imply that FGF drives hepatic fate
in part by transiently suppressing canonical WNT signaling.
BMP4 is a member of the TGFβ superfamily. BMP2, BMP4, FROM HEPATIC ENDODERM
BMP5, and BMP7 are highly expressed in the septum transver- TO LIVER BUD
sum mesenchyme (STM), the latter consisting of loose mesen-
chyme cells that surround the cardiac and ventral endoderm The progress from hepatic endoderm to liver bud has been
domains [46–50]. The BMP receptors BMPRIA, BMPRII, and divided into three morphogenetic stages [17]: stage I, the forma-
ActRIIA are expressed in the endoderm [51, 52]. The mouse tion of a thickened, columnar hepatic epithelium; stage II, the
foregut explant system described above was used to reveal that formation of a pseudostratified epithelium (Figure  2.3c); and
BMP signaling from STM cells is also crucial for hepatic gene stage III, laminin breakdown and hepatic cell migration from
induction in the endoderm [53]. Thus, hepatic induction requires the epithelium into the STM. Hepatoblast migration is accom-
positive signals (FGF and BMP) from two different cell sources panied by major remodeling of the extracellular matrix sur-
(cardiac mesoderm and STM), indicating the importance of rounding the hepatic cells [61]. During this period, the entire
combinatorial signaling. The general relevance of FGFs and ventral foregut domain extends toward the midgut, bringing the
BMPs for hepatic induction has been highlighted by genetic liver region with it (Figure 2.3a,b). The mass of cells emerging
experiments in zebrafish [54]. from the endodermal epithelium and concentrating in the
The zebrafish studies of hepatic induction, which cleverly ­septum transversum is referred to as the liver bud, and the cells
traced the fate of individual endoderm cells [54], confirmed an within the liver bud are referred to as hepatoblasts. Hepatoblasts
initial observation made with populations of endoderm cells iso- will later differentiate into hepatocytes and cholangiocytes
lated from mouse embryos. That is, the foregut endoderm makes (­biliary cells).
a cell fate choice for the liver or pancreas, under the influence of At stage I of liver bud formation in mammals, endothelial
FGF and BMP signaling, as no FGF and BMP allow pancreas cells, not yet assembled into a vasculature, are adjacent to the
induction, whereas elevated levels of both signaling molecular hepatoblast epithelium [62]. Genetic depletion of endothelial
induce a liver fate [55]. Subsequent studies revealed that the cells at this stage led to the discovery that endothelial cells are
signaling network is dynamic, with intracellular FGF and BMP an essential stimulus to further liver bud development: the
responses changing in the endoderm tissue as it moves past the absence of endothelial cells halted liver bud development at
cardiac mesoderm and STM signaling centers [56]. During this stage II, with hepatoblasts remaining within the limits of the
period, the hepatogenic BMP signal in endoderm is transmitted basal epithelial membrane. A similar organogenic stimulatory
to target genes via the transcription factor SMAD4, which in role has been attributed to endothelial cells in the developing
turn recruits the p300 coactivator [57]. P300 is a histone acetyl- pancreas [63]. In the pancreas, VEGF‐A secreted by beta cells
transferase and its activity modulates liver gene induction in the attracts endothelial cells to the developing islet [64]. Endothelial
2:  Embryonic Development of the Liver 17

(a) (b)
neural
tube

dorsal
aorta
gut tube

pericardio/
peritoneal
canal
sinus
venosus
liver bud

(c)

hepatic endoderm
cells
(pseudo-stratified
epithelium)

endothelial cells and


septum transversum
mesenchyme cells

Figure 2.3  The beginning of liver development. (a) Mouse embryo, 9 days gestation. Tail is removed; arrow indicates plane of section in
(b). (b) 100× magnification of transverse section. Note the thickening of the endodermal epithelium of the gut tube at the region of the liver bud.
Boxed area is magnified to 400× in (c). Arrows depict the liver bud and other landmarks of the embryo. Photographs courtesy of J. Rossi.

cells, in turn, form a basal membrane rich in laminins, collagen the ventral foregut endoderm, contributing to the exclusion of
IV, and fibronectin required for the correct function of beta the intestinal fate [17].
cells. The nature of the signals originating from the endothelial
cells that trigger morphogenesis and cell differentiation in the
emerging liver bud is not known.
As already stated, the formation of the liver bud requires the
BEYOND THE LIVER BUD:
degradation of the basal membrane to allow migration of hepat- INTRAHEPATIC CHOLANGIOCYTE
ocytes into the STM. One of the genes controlling this step is DIFFERENTIATION
Prox1 [61]. The Prox1−/− mouse embryonic hepatic endoderm is
unable to degrade the basal epithelial membrane, resulting in The complex signals that cause the emergence of the liver bud
the clustering of hepatoblasts into a smaller liver. It is not known are followed closely by distinct signals required to grow
whether this activity is caused by a slow degradation or produc- the  bud into the liver organ. The hepatoblasts differentiate
tion of excess basal membrane components by the embryonic into hepatocytes and intrahepatic cholangiocytes at about
hepatoblasts. Expression of Prox1 is regulated in part by the ­embryonic day of gestation 13.5 (E13.5) in the mouse and
transcription factor Tbx3. As in Prox1−/− embryos, outgrowth of 7  weeks in humans [67–71]. Substantial advances in our
the liver bud is severely inhibited in Tbx3−/− embryos; however, understanding of the mechanisms that control the formation
it has been suggested that Tbx3 acts to control hepatic progeni- of the intrahepatic ducts have been provided through the
tor cell fate and proliferation through regulating expression of study of conditional knockout mouse embryos and the use of
multiple transcription factors and cell cycle regulators, rather Cre‐mediated lineage tracing [72]. The intrahepatic bile ducts
than acting solely through Prox1 [65, 66]. consist of cholangiocytes (biliary epithelial cells). They
As noted above, the homeodomain factor Hex is crucial for derive from hepatoblasts that surround branches of the portal
liver development [20] and plays an essential role in transforma- vein to form a structure called the ductal plate. Portal vein
tion from stage I to stage II [17]. Hex seems to control different mesenchyme secretes high levels of TGFβ which drives the
aspects of liver bud formation, including the adequate prolifera- differentiation of cholangiocytes and suppresses hepatocyte
tion of hepatic endoderm cells, the formation of a pseudostrati- fate. The action of TGFβ is strictly regulated and only a single
fied epithelium, and the stable maintenance of the hepatic cell layer of hepatoblasts that surround the portal vein differenti-
type [16–18]. How Hex exerts its role is not known, but it has ate to cholangiocytes [73]. The molecular effectors involved
been suggested that it represses sonic hedgehog signaling within in the fate decisions involve the ONECUT transcription
18 THE LIVER:  ROLE OF MESENCHYMAL CELLS IN HEPATOCYTE DIFFERENTIATION

Pre-natal liver development intrahepatic Post-natal


bile duct
BMP4 cholangiocytes
FGF2 HNF-6
Notch
FGF8 HNF-1b

Definitive Hepatic Bipotential Embryonic Adult


endoderm (low endoderm hepatoblasts hepatocytes hepatocytes
FoxA1/2 Wnt) Hex Hex HNF-4
Gata4/6 Prox-1 OC1(HNF-6) C/EBPs
OC2

Figure 2.4  Schema of steps of liver development and relevant signals and transcription factors that help mediate the steps. See text for details.

factors HNF6 (OC1) and OC2. Double homozygous OC1/2 bud, or within the caudal portion of the liver bud [82A]. Cell‐
mutant embryos present a disturbance of the TGFβ gradient tracing experiments will definitively assess this issue.
across the axis of the portal vein to the hepatic parenchyma.
As a result of the increase in TGFβ signaling, hepatoblasts
lying distal to the portal veins express both hepatocyte and
cholangiocyte markers [73] (Figure  2.4). Control of TGFβ ROLE OF MESENCHYMAL CELLS
signaling occurs at multiple levels, including regulation of the IN HEPATOCYTE DIFFERENTIATION
concentration of TGFβ type II receptors (TBRII) on the cell
surface of periportal hepatoblasts [74]. The Hippo‐YAP sign- Hepatocyte differentiation and liver morphogenesis is
aling pathway has also been implicated in controlling the dependent upon the cellular microenvironment. During early
TGFβ axis by directly promoting expression of TGFβ2 and stages of liver bud development, as discussed above, the
inhibiting expression of key hepatocyte transcription factors, microenvironment is provided by endothelial and mesenchy-
such as HNF4 [75]. mal cells in the STM and, after E10.5 in the mouse, when
Together with TGFβ, Notch signaling is also required for liver becomes a hematopoietic organ, hematopoietic stem
­normal biliary tract morphogenesis. Alagille syndrome, a devel- cells (HSCs) also contribute. The relevance of these mesen-
opmental disorder characterized by a paucity of intrahepatic chymal cell types is highlighted by genetic loss‐of‐function
bile ducts (IHBD), is caused predominantly by mutations in the studies in mice, where genes expressed in mesenchymal cells
Jagged1 (JAG1) gene, which encodes a ligand for Notch family surrounding the nascent liver but not in hepatoblasts are
receptors [76]. Studies using notch loss‐of‐function mice and found to be essential for correct liver formation. One of these
zebrafish models have indicated that Notch regulates bile duct cases, Lhx2, is a LIM‐homeobox gene expressed in the STM
abundance rather than cell fate specification [77–79]. However, and mesenchymal components in the adult liver (presumably
overexpression of the Notch intracellular domain, which drives stellate cells). Lhx2−/− livers display a disrupted cellular
expression of Notch target genes, increases the expression of organization with increased deposition of ECM and an altered
cholangiocyte transcription factors and represses hepatocyte gene expression pattern of the early hepatocytes [83]. The
transcription factors [80]. These findings imply that Notch mesenchymal cells are also required for the proliferation of
could directly promote cholangiocyte differentiation. The the fetal hepatocytes. When Gata4 was deleted specifically in
mechanism through which Notch controls cholangiocyte forma- the liver mesenchymal cells, embryos exhibited severe defects
tion is complicated by the fact that Notch interaction with its in liver growth, hepatocyte proliferation and survival, and
ligands Jagged 1 and Delta can have both cell‐extrinsic and fetal hematopoiesis [84].
cell‐intrinsic effects [81]. Besides the early role of individual endothelial cells in pro-
Although during development cholangiocytes derive from moting liver bud growth, blood vessels develop de novo within
the hepatoblasts within the ductal plate, it has recently been the liver bud (Figure 2.3b), forming a capillary bed that becomes
shown that hepatocytes retain the capacity to fully transdiffer- interspersed within the expanding hepatoblast population [85].
entiate into cholangiocytes in the livers of adult mice that lack These transitions establish the liver’s sinusoidal architecture,
an intrahepatic biliary system [82]. Unlike the differentiation of which is critical for organ function and sets the stage for the fetal
cholangiocytes from hepatoblasts, transdifferentiation of adult liver to support hematopoiesis. Hematopoietic cells migrate to
hepatocytes occurred independently of Notch and was driven the early liver first from the yolk sac [86] and later from the
by TGFβ signaling. These findings suggest new strategies for aorta–gonad–mesonephros region [87]. Proper intermediate fila-
the potential treatment of Alagille syndrome and cholestatic ment expression is critical for blood cell homing, as erythrocytes
liver disease. accumulate excessively in the fetal liver of keratin 8 mutants
On the other hand, a recent report suggests that the cells of [88]. Embryos deficient in the heavy metal‐responsive transcrip-
the extrahepatobiliary system (gallbladder, the hepatic, cystic tion factor MTF‐1 exhibit reduced cytokeratin expression as well
and common ducts), share a common origin with the ventral as enlarged sinusoids, and dissociated epithelial cells, but not
pancreas in a group of cells that are slightly caudal to the liver anemia [89]. The primary defect in MTF‐1‐deficient embryos
2:  Embryonic Development of the Liver 19

appears to be in a failure to control metal homeostasis and the grow in this fashion, resulting in major losses of pancreatic
oxidation–reduction state in hepatocytes at mid‐gestation. mass after partial ablation of its embryonic progenitors [107]. It
Both hematopoietic and endothelial cells provide differentiat- seems that liver, but not the pancreas, has a full regenerating
ing signals to the hepatoblasts, because heterogeneity in hepatic potential at nearly all developmental stages and while the
gene expression has been found to be related to the vascular growth limit of the liver is set by the needed organ size, in the
architecture of the fetal liver [90]. More specifically, mouse case of the pancreas it is set by the cell division number. It is
gene inactivation studies have shown that oncostatin M signal- noteworthy that even when the pancreas does not recover its
ing from hematopoietic cells to nascent hepatocytes is critical original size after 50% pancreatectomy, it does recover nearly
for liver growth [91]. Impaired hematopoietic cell proliferation 100% of its beta cell mass after four weeks [108], suggesting
in c‐myb mutant embryos [92] or impaired erythrocytic cell pro- different regenerating capabilities between the endocrine and
liferation in retinoblastoma gene (Rb) mutant embryos [93, 94] exocrine compartment. Another interesting aspect is the differ-
also result in impaired liver growth. The failure of hematopoi- ence between human and animals, since humans, unlike rodents,
etic cells to migrate to the liver in β1 integrin‐null embryos is do not have a significant increase in beta cell proliferation after
therefore also thought to contribute to the liver developmental partial pancreatectomy [109]. Identifying the genetic programs
defect in the β1 mutants [95, 96]. that cause the marked difference in hepatic and pancreatic
As the fetal liver matures and grows, a capsule forms around regeneration at the embryonic stage, when the cells are other-
it from mesothelial cells. Although the role of the mesothelium wise quite immature, could be a convenient way to reveal the
during liver development has not been appreciated, a growing basis for adult liver regeneration.
body of evidence supports a role in promoting fetal hepatocyte
proliferation. N‐myc gene expression in the liver capsule and
jumonji gene expression in the hepatic stromal cells help pro-
mote growth during mid‐gestation [97–99]. The mesothelial HEPATOCYTE DIFFERENTIATION
capsule is a rich source of growth factors that may act on the
fetal hepatocytes [100]. Moreover, deletion of the zinc finger Hepatocyte differentiation spans from liver specification in ven-
transcription factor Wt1, which is highly expressed in the liver tral foregut endoderm until the postnatal maturation of hepato-
capsule during development, inhibits proliferation of fetal hepat- cytes (Figure  2.4). Downstream of signaling molecules that
ocytes, diminishes liver growth, and affects formation of the induce liver differentiation are the transcription factors that
liver lobules [100, 101]. Although many of the genes that affect execute the liver program, including HNF1, HNF4, HNF6,
hepatoblast proliferation are probably most important during the FOXA, and C/EBP [4–6]. These liver‐specific genes activate, in
initial transition from the liver bud to the organ stage, inactiva- a cross‐regulatory fashion, each other’s promoters. By estab-
tion of these genes usually manifests itself as a hematopoietic lishing both positive and negative feedback loops, the transcrip-
defect well after the organ is formed. In these cases, a liver tion factors generate stable gene regulatory networks that ensure
­capsule develops and hematopoietic cells migrate to the liver, expression of genes that are critical for liver function [72, 110].
but the paucity of hepatoblasts leads to a defect in the hemat- In mice, HNF4 acts as a central regulator of hepatocyte gene
opoietic environment and, consequently, embryonic lethality. expression, but is dispensable for early formation of the liver
Furthermore, mutations of certain liver regulatory factors yield a [111]. However, when its role is examined during the differen-
fetal liver growth defect due to apoptosis of the hepatoblasts. tiation of human iPSCs, the requirement for HNF4 in control-
These proteins include c‐jun [102, 103], IKK2 [104], RelA ling the onset of hepatocyte gene expression is even more strict
[105], and XBP‐1 [106]. [112]. Whether the strict requirement for HNF4α in human cells
reflects differences between species or between the model sys-
tems remains to be established. Reverse transcription polymer-
ase chain reaction (RT‐PCR) analyses in the mouse revealed
EMBRYOLOGIC CONTROL OF LIVER that the expression of the primary liver‐enriched transcription
REGENERATION factors was unchanged in HNF4−/− livers at E12, except for PXR
and HNF1α [111]. However, a large number of tissue‐specific
The liver is among the few internal organs that can rapidly genes involved in the maturation of hepatoblasts to hepatocytes
regenerate after removal of tissue in the adult. Recent studies failed to be induced in HNF4−/− embryos. The extensive effects
indicate that the regenerative capacity of the liver is attained as were explained by a genome‐wide chromatin analysis, whereby
early as the hepatoblast stage in embryos. Tissue complementa- the location of HNF4 protein bound to promoter sequences of
tion experiments, where Hex−/− mouse embryonic stem cells nearly 10 000 genes was assessed simultaneously, using micro-
were injected into Hex+/+ blastocysts, discovered that at the array technology [113]. Such analysis was also performed for
onset of liver morphogenesis (E9.0), wild‐type hepatic progeni- HNF6 and HNF1α [113]. HNF1α in human hepatocytes was
tor cells can increase their proliferation rate to compensate for bound to 1.6% of the genes on the array; HNF6 bound 1.7% of
the failure of Hex−/− cells in the liver bud to survive [17]. In an the genes; and HNF4α bound to a striking 12% of the genes
independent study, when two‐thirds of liver progenitors are on the array. The genes bound by HNF4α corresponded to ~42%
genetically ablated between E9.5 and E13.5, the remaining cells of the genes bound by RNA polymerase II. That is, nearly half of
are still able to engage in a compensatory growth, compensating the active genes tested in the liver are bound by HNF4α, and
to generate a normal‐sized fetal liver within 4 days [107]. similar results were obtained by analyzing HNF4α binding to
The  pancreas, also an endoderm‐derived organ, is unable to the entire genome [114]. Subsequent studies found that HNF4
20 THE LIVER:  REFERENCES

regulates genes involved in cell junction assembly and adhesion ACKNOWLEDGMENTS


in the developing liver [115], consequently promoting epithelial
maturation of the liver parenchyma [116]. Thanks to Melanie Song for help in preparing the manuscript.
The apparently limited role of relevant transcription factors R.B. is supported by a grant from the Spanish Ministry of
such as HNF1α, in liver development, based on gene inactiva- Science (SAF‐51991R). K.Z.’s research on liver development is
tion studies in animals, contrasts sharply with ectopic and over- supported by a grant from the NIH (GM36477).
expression studies in cultured hepatic cell lines, which suggested
that HNF1α is critical for the expression of a wide variety of
liver‐specific genes [117]. Such distinctions indicate how
strongly the mechanisms of gene regulation are influenced by REFERENCES
whole‐animal physiology, beyond simple notions of gene redun-
1. Du Bois, A.M. The embryonic liver, in The Liver, Vol. 1 (ed. C.H. Rouiller),
dancy (e.g. [118, 119]). Ultimately, it is crucial to determine
Academic Press, New York, 1963, pp. 1–39.
how inductive signaling pathways and epigenetic modifications 2. Elias, H. Origin and early development of the liver in various vertebrates.
converge on regulatory transcription factor genes to coordi- Acta Hepat, 1955;3:1–57.
nately promote early liver differentiation and morphogenesis. 3. Severn, C.B. A morphological study of the development of the human liver.
Am J Anat, 1968;133:85–108.
4. Zaret, K.S. Regulatory phases of early liver development: paradigms of
organogenesis. Nat Rev Genet, 2002;3(7):499–512.
5. Zaret, K.S. Genetic programming of liver and pancreas progenitors: lessons
FUTURE AND PERSPECTIVES for stem‐cell differentiation. Nat Rev Genet, 2008;9(5):329–40.
6. Zhao, R. and Duncan, S.A. Embryonic development of the liver. Hepatology,
This review has highlighted ways that our understanding of 2005;41(5):956–67.
embryonic liver development has expanded greatly in the past 7. Si‐Tayeb, K., Lemaigre, F.P., and Duncan, S.A. Organogenesis and develop-
ment of the liver. Dev Cell, 2010;18(2):175–89.
25 years or so. The regulatory molecules that control liver
8. Ober, E.A. and Lemaigre, F.P. Development of the liver: Insights into organ
development are now being used successfully to program and tissue morphogenesis. J Hepatol, 2018;68(5):1049–62.
hepatic cells from embryonic stem cells and other sources, 9. Gordillo, M., Evans, T., and Gouon‐Evans, V. Orchestrating liver develop-
including the combination of various cell types to make orga- ment. Development, 2015;142(12):2094–108.
noids [120–122]. Given the large number of proteins that were 10. Fukuda‐Taira, S. Hepatic induction in the avian embryo: specificity of
reactive endoderm and inductive mesoderm. J Embryol Exp Morphol,
­
discovered in studies of adult livers and yet give embryonic 1981;63:111–25.
liver phenotypes when deleted in mice, there is high confidence 11. Le Douarin, N.M. et al. Origin of hemopoietic stem cells in embryonic bursa
that many of these proteins’ functions in the adult are a reca- of Fabricius and bone marrow studied through interspecific chimeras. Proc
pitulation of activities in the embryo. Thus, continued investiga- Natl Acad Sci U S A, 1975;72(7):2701–5.
tion of embryonic liver development is certain to be instructive 12. Tremblay, K.D. and Zaret, K.S. Distinct populations of endoderm cells
­converge to generate the embryonic liver bud and ventral foregut tissues.
about the function, regeneration, and repair of the adult liver, Dev Biol, 2005;280:87–99.
and seems likely to provide new sources of cells and molecules 13. Gualdi, R. et  al. Hepatic specification of the gut endoderm in vitro: cell
to combat liver disease [123]. Further refinements in the ability ­signaling and transcriptional control. Genes Dev, 1996;10:1670–82.
to inactivate genes in the early liver and better methods of 14. Schohl, A. and Fagotto, F. A role for maternal beta‐catenin in early ­mesoderm
induction in Xenopus. Embo J, 2003;22(13):3303–13.
embryo tissue culture are needed to advance the analysis of
15. McLin, V.A., Rankin, S.A., and Zorn, A.M. Repression of Wnt/beta‐catenin
liver development. Such advances may come from adaptations signaling in the anterior endoderm is essential for liver and pancreas
of methodology for other endoderm‐derived organ systems ­development. Development, 2007;134(12):2207–17.
[124]. In addition, relevant new genes will emerge from studies 16. Bort, R. et al. Hex homeobox gene‐dependent tissue positioning is required for
of model organisms where genetic screens can be coupled with organogenesis of the ventral pancreas. Development, 2004;131(4):797–806.
17. Bort, R. et al. Hex homeobox gene controls the transition of the endoderm
knowledge of genomic sequences and interrelationships of pro-
to  a pseudostratified, cell emergent epithelium for liver bud development.
tein function. New models have also emerged that allow high‐ Dev Biol, 2006;290(1):44–56.
resolution insight into the molecular basis of cell differentiation 18. Hunter, M.P. et al. The homeobox gene Hhex is essential for proper hepatoblast
[125]. The differentiation of human embryonic stem cells and differentiation and bile duct morphogenesis. Dev Biol, 2007;308(2):355–67.
induced pluripotent stem cells toward a hepatic fate appears to 19. Keng, V.W. et al. Homeobox gene Hex is essential for onset of mouse embry-
onic liver development and differentiation of the monocyte lineage. Biochem
closely mimic known developmental processes [126]. Having Biophys Res Commun, 2000;276(3):1155–61.
access to a tissue culture model of liver cell differentiation can 20. Martinez‐Barbera, J.P. et al. The homeobox gene hex is required in definitive
complement the study of traditional animal models. The differ- endodermal tissues for normal forebrain, liver and thyroid formation.
entiation of iPSCs in culture is relatively synchronous, which Development, 2000;127(11):2433–45.
allows researchers to address the molecular changes that rapidly 21. Wallace, K.N. et al. Zebrafish hhex regulates liver development and diges-
tive organ chirality. Genesis, 2001;30(3):141–3.
and dynamically occur in response to growth factor signaling 22. Zaret, K., Developmental competence of the gut endoderm: genetic potentia-
[45, 127]. Moreover, tissue culture models allow researchers to tion by GATA and HNF3/fork head proteins. Dev Biol, 1999;209:1–10.
apply high‐­throughput methods and genome‐wide assays to 23. Ang, S.L. et al. The formation and maintenance of the definitive endoderm
identify novel pathways and mechanisms that control cell fate lineage in the mouse: involvement of HNF3/forkhead proteins. Development,
1993;119(4):1301–15.
[127–129]. Considering that our mechanistic understanding of
24. Monaghan, A.P. et al. Postimplantation expression patterns indicate a role
liver development has emerged so recently, the prospects are for the mouse forkhead/HNF‐3α, β, and γ genes in determination of the
bright for much deeper knowledge and new applications for definitive endoderm, chordamesoderm and neuroectoderm. Development,
liver therapies in the future. 1993;119:567–78.
2:  Embryonic Development of the Liver 21

25. Ruiz i Altaba, A. et  al. Sequential expression of HNF‐3α and HNF‐3β by 52. Roelen, B.A. et al. Differential expression of BMP receptors in early mouse
embryonic organizing centers: the dorsal lip/node, notochord, and floor development. Int J Dev Biol, 1997;41(4):541–9.
plate. Mech Dev, 1993;44:91–108. 53. Rossi, J.M. et  al. Distinct mesodermal signals, including BMP’s from the
26. Sasaki, H. and Hogan, B.L.M. Differential expression of multiple fork head septum transversum mesenchyme, are required in combination for hepa-
related genes during gastrulation and pattern formation in the mouse embryo. togenesis from the endoderm. Genes Dev, 2001;15:1998–2009.
Development, 1993;118:47–59. 54. Shin, D. et al. Bmp and Fgf signaling are essential for liver specification in
27. Arceci, R. et  al. Mouse GATA‐4: a retinoic acid‐inducible GATA‐binding zebrafish. Development, 2007;134(11):2041–50.
transcription factor expressed in endodermally derived tissues and heart. Mol 55. Deutsch, G. et al. A bipotential precursor population for pancreas and liver
Cell Biol, 1993;13:2235–46. within the embryonic endoderm. Development, 2001;128:871–81.
28. Gao, X. et al. Distinct functions are implicated for the GATA‐4, ‐5, and ‐6 56. Wandzioch, E. and Zaret, K.S. Dynamic signaling network for the specifica-
transcription factors in the regulation of intestine epithelia cell differentia- tion of embryonic pancreas and liver progenitors. Science, 2009. 324(5935),
tion. Mol Cell Biol, 1998;18:2901–11. 1707–10.
29. Laverriere, A.C. et al. GATA‐4/5/6, a subfamily of three transcription factors 57. Xu, C.R. et al. Chromatin “prepattern” and histone modifiers in a fate choice
transcribed in developing heart and gut. J Biol Chem, 1994;269:23177–84. for liver and pancreas. Science, 2011;332(6032):963–6.
30. Suzuki, E. et al. The human GATA‐6 gene: structure, chromosomal location, 58. Ober, E.A. et  al. Mesodermal Wnt2b signalling positively regulates liver
and regulation of expression by tissue‐specific and mitogen‐responsive sig- specification. Nature, 2006;442(7103):688–91.
nals. Genomics, 1996;38:283–90. 59. Lakshminarayanan, B. and Davenport, M. Biliary atresia: A comprehensive
31. Bossard, P. and Zaret, K.S. GATA transcription factors as potentiators of gut review. J Autoimmun, 2016;73:1–9.
endoderm differentiation. Development, 1998;125:4909–17. 60. Cayuso, J. et  al. EphrinB1/EphB3b coordinate bidirectional epithelial‐­
32. Liu, J.K., DiPersio, C.M., and Zaret, K.S. Extracellular signals that regulate mesenchymal interactions controlling liver morphogenesis and laterality.
liver transcription factors during hepatic differentiation in vitro. Mol Cell Dev Cell, 2016;39(3):316–28.
Biol, 1991;11(2):773–84. 61. Sosa‐Pineda, B., Wigle, J.T., and Oliver, G. Hepatocyte migration during
33. Cascio, S. and Zaret, K.S. Hepatocyte differentiation initiates during endo- liver development requires Prox1. Nat Genet, 2000;25(3):254–5.
dermal‐mesenchymal interactions prior to liver formation. Development, 62. Matsumoto, K. et  al. Liver organogenesis promoted by endothelial cells
1991;113:217–25. prior to vascular function. Science, 2001;294:559–63.
34. Mueller, P.R. and Wold, B. In vivo footprinting of a muscle specific enhancer 63. Lammert, E., Cleaver, O., and Melton, D. Induction of pancreatic differentia-
by ligation mediated PCR. Science, 1989;246:780–6. tion by signals from blood vessels. Science, 2001;294(5542):564–7.
35. Lee, C.S. et  al. The initiation of liver development is dependent on Foxa 64. Nikolova, G. et  al. The vascular basement membrane: a niche for insulin
transcription factors. Nature, 2005;435(7044):944–7. gene expression and beta cell proliferation. Dev Cell, 2006;10(3):397–405.
36. Holtzinger, A. and Evans, T. Gata4 regulates the formation of multiple 65. Suzuki, A. et  al. Tbx3 controls the fate of hepatic progenitor cells in liver
organs. Development, 2005;132(17):4005–14. ­development by suppressing p19ARF expression. Development, 2008;135(9):
37. Fisher, J.B. et al. GATA6 is essential for endoderm formation from human 1589–95.
pluripotent stem cells. Biol Open, 2017;6(7):1084–95. 66. Ludtke, T.H. et  al. Tbx3 promotes liver bud expansion during mouse
38. Tiyaboonchai, A. et al. GATA6 plays an important role in the induction of ­development by suppression of cholangiocyte differentiation. Hepatology,
human definitive endoderm, development of the pancreas, and functionality 2009;49(3):969–78.
of pancreatic beta cells. Stem Cell Rep, 2017;8(3):589–604. 67. Germain, L., Blouin, M.J., and Marceau, N. Biliary epithelial and hepato-
39. Shi, Z.D. et al. Genome editing in hPSCs reveals GATA6 haploinsufficiency cytic cell lineage relationships in embryonic rat liver as determined by the
and a genetic interaction with GATA4 in human pancreatic development. differential expression of cytokeratins, α‐fetoprotein, albumin, and cell
Cell Stem Cell, 2017;20(5):675–688.e6. ­surface‐exposed components. Cancer Res, 1988;48(17):4909–18.
40. McKeehan, W.L., Wang, F., and Kan, M. The heparan sulfate‐fibroblast 68. Shiojiri, N. Enzymo‐ and immunocytochemical analyses of the differentia-
growth factor family: diversity of structure and function. Prog Nucl Acid Res tion of liver cells in the prenatal mouse. J Embryol Exp Morphol,
Mol Biol, 1998;59:135–76. 1981;62:139–52.
41. Szebenyi, G. and Fallon, J.F. Fibroblast growth factors as multifunctional 69. Shiojiri, N. Analysis of differentiation of hepatocytes and bile duct cells in
signaling factors. Int Rev Cytol, 1999;185:45–106. developing mouse liver by albumin immunofluorescence. Dev Growth
42. Jung, J. et al. Initiation of mammalian liver development from endoderm by Differ, 1984;26:555–61.
fibroblasts growth factors. Science, 1999;284:1998–2003. 70. Shiojiri, N. Development and differentiation of bile ducts in the mammalian
43. Wang, J.‐K., Gao, G., and Goldfarb, M. Fibroblast growth factor receptors have liver. Microsc Res Technol, 1997;39(4):328–35.
different signaling and mitogenic potentials. Mol Cell Biol, 1994;14:181–8. 71. Antoniou, A. et al. Intrahepatic bile ducts develop according to a new mode
44. Calmont, A. et  al. An FGF response pathway that mediates hepatic gene of tubulogenesis regulated by the transcription factor SOX9.
induction in embryonic endoderm cells. Dev Cell, 2006;11(3):339–48. Gastroenterology, 2009;136(7):2325–33.
45. Twaroski, K. et al. FGF2 mediates hepatic progenitor cell formation during 72. Gerard, C., Tys, J., and Lemaigre, F.P. Gene regulatory networks in differen-
human pluripotent stem cell differentiation by inducing the WNT antagonist tiation and direct reprogramming of hepatic cells. Semin Cell Dev Biol,
NKD1. Genes Dev, 2015;29(23):2463–74. 2017;66:43–50.
46. Dudley, A.M., Rougeulle, C., and Winston, F. The Spt components of SAGA 73. Clotman, F. et al. Control of liver cell fate decision by a gradient of TGF beta
facilitate TBP binding to a promoter at a post‐activator‐binding step in vivo. signaling modulated by Onecut transcription factors. Genes Dev, 2005;19(16):
Genes Dev, 1999;13(22):2940–5. 1849–54.
47. Lyons, K.M., Pelton, R.W., and Hogan, B.L. Patterns of expression of 74. Takayama, K. et al. CCAAT/enhancer binding protein‐mediated regulation
murine Vgr‐1 and BMP‐2a RNA suggest that transforming growth factor‐ of TGFbeta receptor 2 expression determines the hepatoblast fate decision.
beta‐like genes coordinately regulate aspects of embryonic development. Development, 2014;141(1):91–100.
Genes Dev, 1989;3(11):1657–68. 75. Lee, D.H. et al. LATS‐YAP/TAZ controls lineage specification by regulating
48. Solloway, M.J. and Robertson, E.J. Early embryonic lethality in Bmp5;Bmp7 TGFbeta signaling and Hnf4alpha expression during liver development. Nat
double mutant mice suggests functional redundancy within the 60A sub- Commun, 2016;7:11961.
group. Development, 1999;126(8):1753–68. 76. Oda, T. et  al. Mutations in the human Jagged1 gene are responsible for
49. Winnier, G. et al. Bone morphongenetic protein‐4 is required for mesoderm Alagille syndrome. Nat Genet, 1997;16(3):235–42.
formation and patterning in the mouse. Genes Dev, 1995;9:2105–16. 77. Kodama, Y. et al. The role of notch signaling in the development of intrahe-
50. Zhang, H. and Bradley, A. Mice deficient for BMP2 are nonviable and patic bile ducts. Gastroenterology, 2004;127(6):1775–86.
have  defects in amnion/chorion and cardiac development. Development, 78. Lorent, K. et  al. Inhibition of Jagged‐mediated Notch signaling disrupts
1996;122(10):2977–86. zebrafish biliary development and generates multi‐organ defects compatible
51. Mishina, Y. et al. Bmpr encodes a type I bone morphogenetic protein recep- with an Alagille syndrome phenocopy. Development, 2004;131(22):5753–66.
tor that is essential for gastrulation during mouse embryogenesis. Genes 79. Lozier, J., McCright, B., and Gridley, T. Notch signaling regulates bile duct
Dev, 1995;9(24):3027–37. morphogenesis in mice. PLoS ONE, 2008;3(3):e1851.
22 THE LIVER:  REFERENCES

80. Tanimizu, N. and Miyajima, A. Notch signaling controls hepatoblast 107. Stanger, B.Z., Tanaka, A.J., and Melton, D.A. Organ size is limited by the
­differentiation by altering the expression of liver‐enriched transcription number of embryonic progenitor cells in the pancreas but not the liver.
­factors. J Cell Sci, 2004;117(15):3165–74. Nature, 2007;445(7130):886–91.
81. Kaylan, K.B. et al. Combinatorial microenvironmental regulation of liver 108. Lee, C.S. et al. Regeneration of pancreatic islets after partial pancreatec-
progenitor differentiation by Notch ligands, TGFbeta, and extracellular tomy in mice does not involve the reactivation of neurogenin‐3. Diabetes,
matrix. Sci Rep, 2016;6:23490. 2006;55(2):269–72.
82. Schaub, J.R. et al. De novo formation of the biliary system by TGFbeta‐ 109. Menge, B.A. et  al. Partial pancreatectomy in adult humans does not
mediated hepatocyte transdifferentiation. Nature, 2018;557(7704):247–51. ­provoke beta‐cell regeneration. Diabetes, 2008;57(1):142–9.
82. A. Spence, J.R. et al. Sox17 regulates organ lineage segregation of ventral 110. Kyrmizi, I. et  al. Plasticity and expanding complexity of the hepatic
foregut progenitor cells. Dev Cell, 2009;17(1):62–74. transcription factor network during liver development. Genes Dev,
­
83. Wandzioch, E. et al. Lhx2−/− mice develop liver fibrosis. Proc Natl Acad 2006;20(16):2293–305.
Sci U S A, 2004;101(47):16549–54. 111. Li, J., Ning, G., and Duncan, S.A. Mammalian hepatocyte differentiation
84. Delgado, I. et  al. GATA4 loss in the septum transversum mesenchyme requires the transcription factor HNF‐4alpha. Genes Dev, 2000;14(4):
­promotes liver fibrosis in mice. Hepatology, 2014;59(6):2358–70. 464–74.
85. Medlock, E.S. and Haar, J.L. The liver hemopoietic environment: I. 112. DeLaForest, A. et  al. HNF4A is essential for specification of hepatic
Developing hepatocytes and their role in fetal hemopoiesis. Anat Rec, progenitors from human pluripotent stem cells. Development,
­
1983;207(1):31–41. 2011;138(19):4143–53.
86. Johnson, G.R. and Moore, M.A. Role of stem cell migration in initiation of 113. Odom, D.T. et al. Control of pancreas and liver gene expression by HNF
mouse foetal liver haemopoiesis. Nature, 1975;258:726–8. transcription factors. Science, 2004;303(5662):1378–81.
87. Medvinsky, A. and Dzierzak, E. Definitive hematopoiesis is autonomously 114. Hoffman, B.G. et  al. Locus co‐occupancy, nucleosome positioning, and
initiated by the AGM region. Cell, 1996;86(6):897–906. H3K4me1 regulate the functionality of FOXA2‐, HNF4A‐, and PDX1‐
88. Baribault, H. et  al. Mid‐gestational lethality in mice lacking keratin 8. bound loci in islets and liver. Genome Res, 2010;20(8):1037–51.
Genes Dev, 1993;7:1191–202. 115. Battle, M.A. et al. Hepatocyte nuclear factor 4alpha orchestrates expression
89. Günes, C. et al. Embryonic lethality and liver degeneration in mice lacking the of cell adhesion proteins during the epithelial transformation of the devel-
metal‐responsive transcriptional activator MTF‐1. EMBO J, 1998;17:2846–54. oping liver. Proc Natl Acad Sci U S A, 2006;103(22):8419–24.
90. Gaasbeek Janzen, J.W. et al. Gene expression in derivatives of embryonic 116. Parviz, F. et al. Hepatocyte nuclear factor 4alpha controls the development
foregut during prenatal development of the rat. J Histochem Cytochem, of a hepatic epithelium and liver morphogenesis. Nat Genet, 2003;34(3):
1988;36(10):1223–30. 292–6.
91. Kamiya, A. et al. Fetal liver development requires a paracrine action of oncos- 117. Tronche, F. et al. Hepatocyte nuclear factor 1(HNF1) and liver gene expres-
tatin M through the gp130 signal transducer. Embo J, 1999;18(8):2127–36. sion, in Liver Gene Expression (eds. F. Tronche and M. Yaniv), R.G. Landes
92. Mucenski, M.L. et  al. A functional c‐myb gene is required for normal Company, 1994, pp. 155–82.
murine fetal hepatic hematopoiesis. Cell, 1991;65(4):677–89. 118. Barbacci, E. et al. Variant hepatocyte nuclear factor 1 is required for vis-
93. Jacks, T. et  al. Effects of an Rb mutation in the mouse. Nature, ceral endoderm specification. Development, 1999;126(21):4795–805.
1992;359:295–300. 119. Cereghini, S. et al. Expression patterns of vHNF1 and HNF1 homeopro-
94. Lee, E.Y. et al. Mice deficient for Rb are nonviable and show defects in teins in early postimplantation embryos suggest distinct and sequential
neurogenesis and haematopoiesis. Nature, 1992;359(6393):288–94. developmental roles. Development, 1992;116:783–97.
95. Hirsch, E. et al. Impaired migration but not differentiation of haematopoietic 120. Gouon‐Evans, V. et  al. BMP‐4 is required for hepatic specification of
stem cells in the absence of beta1 integrins. Nature, 1996;380(6570):171–5. mouse embryonic stem cell‐derived definitive endoderm. Nat Biotechnol,
96. Houssaint, E., Differentiation of the mouse hepatic primordium. I. An 2006;24(11):1402–11.
­analysis of tissue interactions in hepatocyte differentiation. Cell Differ, 121. Loh, K.M. et  al. Efficient endoderm induction from human pluripotent
1980;9:269–79. stem cells by logically directing signals controlling lineage bifurcations.
97. Giroux, S. and Charron, J. Defective development of the embryonic liver in Cell Stem Cell, 2014;14(2):237–52.
N‐myc‐deficient mice. Dev Biol, 1998;195:16–28. 122. Takebe, T. et  al. Vascularized and functional human liver from an iPSC‐
98. Motoyama, J. et  al. Organogenesis of the liver, thymus and spleen is derived organ bud transplant. Nature, 2013;499(7459):481–4.
affected in jumonji mutant mice. Mech Dev, 1997;66:27–37. 123. Goldman, O. and Gouon‐Evans, V. Human pluripotent stem cells: myths and
99. Sawai, S. et al. Defects of embryonic organogenesis resulting from targeted future realities for liver cell therapy. Cell Stem Cell, 2016;18(6):703–6.
disruption of the N‐myc gene in the mouse. Development, 1993;117:1445–55. 124. Wells, J.M. and Melton, D.A. Vertebrate endoderm development. Annu Rev
100. Onitsuka, I., Tanaka, M., and Miyajima, A. Characterization and functional Cell Dev Biol, 1999;15:393–410.
analyses of hepatic mesothelial cells in mouse liver development. 125. Vallier, L. Heps with pep: direct reprogramming into human hepatocytes.
Gastroenterology, 2010;138(4):1525–35, 1535.e1–6. Cell Stem Cell, 2014;14(3):267–9.
101. Ijpenberg, A. et al. Wt1 and retinoic acid signaling are essential for stellate 126. Yiangou, L. et  al. Human pluripotent stem cell‐derived endoderm for
cell development and liver morphogenesis. Dev Biol, 2007;312(1):157–70. modeling development and clinical applications. Cell Stem Cell,
­
102. Eferl, R. et al. Functions of c‐Jun in liver and heart development. J Cell 2018;22(4):485–99.
Biol, 1999;145(5):1049–61. 127. Bertero, A. et al. Activin/nodal signaling and NANOG orchestrate human
103. Hilberg, F. et al. c‐Jun is essential for normal mouse development and hepa- embryonic stem cell fate decisions by controlling the H3K4me3 chromatin
togenesis. Nature, 1993;365:179–81. mark. Genes Dev, 2015;29(7):702–17.
104. Li, Q. et  al. Severe liver degeneration in mice lacking the 1κB kinase 2 128. Jing, R., Duncan, C.B., and Duncan, S.A. A small‐molecule screen reveals
gene. Science, 1999;284:321–5. that HSP90beta promotes the conversion of induced pluripotent stem cell‐
105. Beg, A.A. et al. Embryonic lethality and liver degeneration in mice lacking derived endoderm to a hepatic fate and regulates HNF4A turnover.
the RelA component of NF‐KB. Nature, 1995;376:167–70. Development, 2017;144(10):1764–74.
106. Reimold, A.M. et al. An essential role in liver development for transcription 129. Camp, J.G. et al. Multilineage communication regulates human liver bud
factor XBP‐1. Genes Dev, 2000;14(2):152–7. development from pluripotency. Nature, 2017;546(7659):533–8.
PART TWO:
THE CELLS
SECTION A:
CELL BIOLOGY
OF THE LIVER
Cytoskeletal Motors: Structure
3 and Function in Hepatocytes
Mukesh Kumar1, Arnab Gupta2, and Roop Mallik1
1
Department of Biological Sciences, Tata Institute of Fundamental Research, Navy Nagar, Colaba, Mumbai, India
2
Department of Biological Sciences, Indian Institute of Science Education and Research Kolkata, Mohanpur, India

INTRODUCTION HEPATOCYTE POLARITY


AND THE ACTIN NETWORK
The hepatocyte is the major epithelial cell of the liver, making
up to 70–80% of liver mass. Hepatocytes produce and secrete Although hepatocyte polarity is well‐described at an anatomical
bile, extract specific molecules from blood, and secrete oth­ level, little is known about the molecules and the mechanisms
ers into blood flow. This “sieving” function makes the liver a that initiate, establish, and maintain this polarization. During
key homeostatic organ and requires intracellular transport of mouse development, a liver bud composed of nonpolarized hepa­
a vast variety of lipids/proteins towards multiple distinctly toblasts is detected as early as embryonic day 9.5. Studies on
polarized surfaces in hepatocytes. In order to do this, hepatocytes mouse liver development by Feracci et al. have demonstrated that
must partition as well as connect two different environments – initiation of hepatocyte polarization occurs very early in develop­
the blood and the bile. To serve this function in the liver, ment (day 15). From embryonic day 17 onwards, hepatoblasts
hepatocytes are arranged in cords. An individual cell is aggregate together resembling acini. The acini exhibit a simple
polygonal and faces at least two blood sinusoids (the basal polar phenotype, with their apical surfaces facing a central lumen [3].
domain). A branched network of grooves between adjacent Gradually, into the postnatal stages, the simple polarity changes
cells forms the bile canaliculus, representing the apical to polygonal hepatic polarity, a phenomenon adeptly emulated in
domain. Such a polygonal shape means that hepatocytes do WIF‐B cells. This entire process of ­hepatocyte as well as WIF‐B
not have a single basolateral‐to‐apical axis, and the transcy­ polarization is heavily dependent on the spatiotemporal location
totic pathways between basolateral and apical membranes are and function of the cytoskeletal ­proteins. The anatomy of micro­
more complex than other single apical–basolateral axis polar­ filament in the polarized hepatocyte was studied in great detail by
ized epithelial cells, such as enterocytes, renal epithelial Ishii et  al. [4]. WIF‐B cells maintain PM proteins that are
cells, etc. [1]. Most newly synthesized membrane and secre­ restricted to either the apical (bile  canalicular) or basolateral
tory proteins exiting the trans‐Golgi network (TGN) are domain, with tight junctions demarcating the apical–basolateral
directed via the basolateral membrane, with less TGN‐exit boundaries. Phalloidin staining reveals a thin but highly intense
traffic directed towards the apical side (Figure  3.1a). ringed network of actin f­ ilaments around the bile canaliculus/api­
Similarly, endocytosed cargos including plasma membrane cal membrane (BC), along with a sparse network close to the
(PM) fragments internalized at the basolateral (blood) side basolateral membrane (Figure 3.1a). Densely concentrated actin
are transported to lysosomes, the apical PM, or bile via tran­ filaments were identified around the bile canaliculi in the form of
scytotic pathways. Since these multiple and distinct sorting microvillous core filaments and pericanalicular web filaments.
steps of the cargos in the biosynthetic and endocytic path­ The microvillous core filaments exhibited growth at their apical
ways rely on microtubules, microfilaments (actin), and their ends. In contrast, filaments of the pericanalicular web, running
respective motor proteins, the cytoskeletal system functions parallel to the cell surface, showed no fixed polarities.
as the primary coordinator between the apical and basolateral Interestingly, adjacent ­ filament pairs often showed opposite
regions in hepatocytes [2]. polarities, an alignment necessary for filament sliding. A group

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
28 THE LIVER:  HEPATOCYTE POLARITY AND THE ACTIN NETWORK

(a) (b)
Bile
canaliculus
Cargo binding domain Myosin V

SAC

?
Golgi
Coiled coil domain
?

? Light chain binding domain


+ ? Nucleus

Basola + Motor domain


te
endos ral
ome
MyoVb with ATP7B as cargo
Cortical actin

Figure 3.1  Actin‐dependent trafficking in hepatocytes and the myosin‐V motor. (a) Most of the apically targeted proteins (copper transporting
ATPase, ATP7B in this case) travels to the destination via basolateral endosomes. The microtubule‐based motor proteins that drive the cargo from
the TGN to the basolateral endosomes and subsequently to the subapical compartment (SAC) has not been determined (shown as interrogative
clause). Once the cargo‐carrying vesicles reach the SAC, myosin Vb drives its recycling between the SAC and the bile canaliculus/apical mem­
brane. The myosin Vb remains anchored to the actin microfilaments that form a dense network beneath the apical membrane. (b) The general
structure of myosin V that is conserved in all the myosin V proteins (MyoVa, Vb, and Vc).

of sporadic actin f­ ilaments with no uniform polarity acts as a link proteins in Rab11a‐containing endosomes and did not distribute
between the canalicular membrane and coated vesicles [4]. Such to the PM [10]. The study demonstrated that polarization of
spatial organization of actin filaments is implicated in mainte­ hepatocytes requires recruitment of Rab11a and myosin Vb to
nance of microvilli length, contraction of the canalicular walls intracellular membranes that contain apical ABC transporters,
facilitating bile secretion, and transport of coated vesicles in the permitting their targeting to the PM.
apical membrane (bile canalicular membrane). Treatment of Stable MYO5 knockdown (MY05B‐KD) in CaCo‐2 cells
hepatocytes with cytochalasin D, which targets actin, completely induced loss of microvilli, alterations in tight junction proteins,
inhibited self‐assembly of rat hepatocytes into spheroids. Thus, and disruption of polarized trafficking [9, 11, 12]. Due to short
hepatocytes require an intact actin network to self‐assemble effi­ life expectancy, liver phenotypes usually go undetected in
ciently into functional tissue‐like structures [5]. MVID patients [13]. However, MVID patients developed a
Of all the actin‐related motors, most work has been done on cholestatic liver disease similar to progressive familial intrahe­
myosin V (Figure  3.1b). Myosin Vb is most abundant in the patic cholestasis (PFIC) and benign recurrent cholestasis (BRC)
liver as well as the WIF‐B cells [6]. It is an unconventional myo­ [14]. PFIC and BRC have been linked to mutations in ABCB11
sin regulating vesicle trafficking by binding Rab11a [7, 8]. (encoding bile salt export pump, BSEP) and ATP8B1 (encoding
Myosin Vb inhibition induces failure to attain complete polarity FIC1 protein) genes, respectively [15, 16]. Interestingly, in
as well as trafficking defects of various apically targeted cargos, neither of those two genes were disease‐causing mutations
­
marked by accumulation of metabolites leading to a diseased detected in any of these patients. Immunohistochemistry stud­
state. Understanding human genetic diseases linked to muta­ ies ­suggested mislocalization of BSEP in cytoplasmic vesicular
tions in myosin Vb and its effector proteins has been greatly staining as compared to controls, where the protein primarily
helpful in comprehending liver polarity and polarized protein localized at the apical membrane. This phenotype is a conse­
trafficking. Mutations in myosin Vb were found to cause micro­ quence of impairment of the MY05B/RAB11A apical recycling
villus inclusion disease (MVID), which is characterized by a endosome pathway in hepatocytes [14]. Utilizing polarized
lack of apical microvilli and intracellular structures containing WIF‐B cells, it was demonstrated that the copper‐exporting
microvilli in the intestine, leading to a severe form of congenital ATPase ATP7B (also known as the Wilson disease protein) is a
diarrhea [9]. Years before detailed mechanistic understanding of major cargo of myosin Vb in the liver. Mutations in ATP7B
MVID phenotypes, Wakabayashi et  al., while investigating imparts a nontrafficking phenotype to the protein, causing a
mechanisms of apical targeting, serendipitously discovered that ­disorder of liver copper accumulation termed Wilson disease [6].
knocking down of Rab11 before its polarization prevented cana­ Interestingly, a similar copper accumulation phenotype was
licular formation in WIF‐B9 cells. Polarization could also be demonstrated in polarized WIF‐B cells when myosin Vb was
arrested by overexpression of the Rab11a‐GDP locked form or acutely knocked down using a DN mutant. Detailed micro­
the myosin Vb tail dominant negative (DN) construct. In WIF‐ scopic studies revealed that in the absence of functional myosin
B9 cells, which lack bile canaliculi, apical ABC (ATP‐binding Vb, ATP7B‐containing vesicles accumulated beneath the
cassette) transporters colocalize with transcytotic membrane ­apical membrane and failed to fuse with it. This study further
3:  Cytoskeletal Motors: Structure and Function in Hepatocytes 29

delineated the physical and functional localizations of micro­ a sedimentation coefficient of 6S and an Mr of 110 000–120 000
tubule and cortical actin with respect to apical trafficking. [23]. Later, it was shown that the denaturation of the 6S protein
Apically targeted cargos exit the TGN and traffic basolaterally produced α and β monomeric subunits (Mr = 55 000). These two
before getting redirected apically in vesicles utilizing the micro­ very similar subunits were found to exist within MTs in a 1 : 1
tubular network [17]. The vesicles are then transferred to the molar ratio, forming αβ heterodimers [24]. The term “tubulin”
myosin Vb motor anchored at the cortical actin that is located was coined by Mohri in 1968 [25]. The ability of tubulin purified
beneath the apical membrane (Figure  3.1a). This transfer of from brain homogenate to polymerize in vitro in the presence of
­cargos occurs in specialized endosomal compartments called GTP, Mg2+, and EGTA greatly improved our understanding of MT
subapical compartments (SACs) lying in close proximity to the assembly and dynamics [26]. A third type of tubulin, γ‐tubulin,
microtubule organizing center (MTOC). Studies on the role of was discovered and was found to be associated with the minus end
myosin Vb in trafficking of apically targeted proteins reveals a of MTs near the centrosome. Although γ‐tubulin is not a major
fascinating spatiotemporal correlation with the polarization component of MTs, it plays an i­ mportant role in the assembly and
­status of WIF‐B cells. In completely polarized WIF‐B cells, expression of MT arrays in eukaryotic cells [27].
myosin Vb is primarily localized as a ring‐like pattern around MTs undergo rapid stochastic growth and shrinkage
the F‐actin ring at the bile canaliculus (Figure 3.1a). However, (Figure 3.2b), known as “dynamic instability” [28, 29]. This phe­
in pre‐polarized cells, myosin Vb, as well as the myosin Vb tail nomenon depends on various intrinsic and extrinsic factors,
DN protein, localize to a tight juxtanuclear intracellular site including the concentration of tubulin. The tubulin subunits asso­
described as the “apical compartment” [18]. This apical compart­ ciate/dissociate at different rates at the plus and minus ends of
ment is an intracellular membranous structure clustered near protofilaments. The critical concentration is the minimum con­
the minus ends of microtubules that are labeled with γ‐tubulin, centration of tubulin dimers above which there is a net growth of
where apically destined cargos accumulate selectively. tubulin polymer. Nucleotides at the N‐site are not available for
exchange, unlike nucleotides at the E‐site which can exchange
with the solution [30]. Preferential polymerization of tubulin
dimers at the β‐tubulin end defines the MT polarity. The rapidly
MICROTUBULES: HISTORY, STRUCTURE polymerizing β‐end is called the plus end whereas the slow‐
AND DYNAMICS growing end with exposed α‐tubulin is called the minus end.
GTP hydrolysis within the polymer is slower than GTP addition
Microtubules (MTs) are tracks for transport of intracellular cargo at the plus end. Resultant accumulation of a “GTP cap” at the
by microtubule‐dependent motors such as dynein and kinesin plus end favors dimer addition [28, 30]. As the tubulin dimer
(Figure  3.2a). MTs were observed as rod‐like structures in concentration decreases below a critical value, dimer addition at
­spermatozoid of Sphagnum and ciliated epithelia [19, 20], and the plus end is reduced in comparison to GTP hydrolysis. This
were hypothesized to power the beating movements of cilia. leads to a rapid shrinkage of MTs and is termed catastrophe.
Improvement in structural preservation techniques [21] showed Catastrophe can be rescued by the addition of GTP‐tubulin [28].
that MTs contain multiple (for example 13) subunits arranged in a The competition between polymerization and catastrophe can be
circular fashion [22]. Ledbetter and Porter named the structure regulated by MT‐associated proteins and MT‐severing proteins
“microtubule,” because the protofilaments were organized around inside cells [31, 32]. Although dynamic, the MTs still collec­
a hollow center and therefore appeared tubular. Taylor and cow­ tively maintain a stiff intracellular skeleton to impart shape to a
orkers characterized a single [3H]‐colchicine binding protein with cell and support intracellular cargo transport.

(a) (c) (d)


Light chain
+ IC
Cargo Tail
domain
LIC

Linker Heavy
Depolymerization chain
Stalk
phase
Polymerization domain
phase Head/motor
ule
tub

domain
cro

β
Mi

– MT binding Motor
domain domain
α Dynein
(b) Kinesin

Figure 3.2  Microtubules and associated motor proteins. (a) Microtubules normally extend from the minus end of the microtubule (anchored at the
microtubule organizing center (MTOC)) to the plus end (fast‐growing). (b) Microtubules consisting of αβ‐tubulin dimers (shown) remain in a
dynamic mode, alternating between polymerization and depolymerization phases. The molecular structures of (c) dynein (retrograde motor) and (d)
conventional kinesin (anterograde motor) are shown. IC, intermediate chain; LIC, light intermediate chain.
30 THE LIVER:  MICROTUBULE ORGANIZATION, POLARIZATION, AND CARGO TRAFFICKING IN HEPATOCYTES

MICROTUBULE MOTORS Kinesin


Since the discovery of conventional kinesin in 1985 by Brady and
MTs also serve as tracks against which molecular motors can
Vale [36, 53], a large number of ATPases have been categorized
generate mechanical force (Figure 3.2a). Motors utilize ATP as
into the kinesin superfamily [53, 54]. Kinesins are classified into
fuel to generate force and movement for intracellular transport,
14 classes (kinesins 1–14), all these motors contain an ATP‐bind­
cell division etc. Gibbons and Rowe were the first to report MT
ing domain with significant homology [55]. Kinesins are protein
motor‐based ATPase activity when they purified the dynein
complexes of approximately 380 kDa, consisting of two heavy
ATPase from Tetrahymena cilia [33]. The term dynein was
chains (each 120 kDa) and two light chains (each 64 kDa) [56]. As
derived from the Greek word dyne, meaning “force,” as this
shown in Figure 3.2d, the N‐terminal globular head domains of
motor was thought to be responsible for beating of cilia.
the heavy chains are the “motor domains” that bind to the MT [57]
Discovery of dynein in the flagella of sea urchin spermatozoa
and also hydrolyze ATP to provide energy for movement [58].
confirmed that dynein was responsible for ciliary and flagellar
The processive motion of conventional kinesins is explained by
beating [34]. Although the MT motor proteins were suspected to
their “hand‐over‐hand” walking [59]. The flexible stalk domains
power mitosis and organelle transport, molecular characteriza­
help kinesin to dimerize, and the neck linker acts as a lever to
tion was difficult due to their low cytosolic concentration. The
facilitate steps [60]. The C‐terminal tail domain, in association
use of extracts from squid giant axon paved the way for studying
with the light chain, forms the cargo‐binding domain of kinesin
intracellular transport in vitro [35, 36]. This was followed by the
(Figure  3.2d). The kinesin tail domain forms complexes with
biochemical purification of two organelle‐associated cytoplas­
adaptor proteins such as Miro/Milton, syntabulin, DENN/MADD,
mic motors, namely dynein and kinesin, that usually transport
etc. [55]. These combinations achieve a high level of specificity in
cargo in opposite directions along MTs [37, 38].
cargo recognition. Spatiotemporal regulation of cargo delivery is
achieved by the phosphorylation of kinesin, Rab GTPase activity
Cytoplasmic dynein and Ca2+ signaling [55]. Most kinesins have a motor domain at the
Dynein (Figure  3.2c) is a multi‐subunit protein complex of N‐terminal region (known as N‐kinesins) that drive cargos
~2000 kDa consisting of two heavy chains (DHC; each ~530 towards the plus end of the microtubule (anterograde motors).
kDa), a variable number of light chains (LCs) and intermedi­ Others kinesins (known as C‐kinesins, e.g. kinesin‐14) have the
ate chains (ICs). The ICs bind to DHC and multiple LCs bind motor domain at the C‐terminal, while a few kinesins have a cata­
to ICs at separate sites [39]. DHC is a single polypeptide and lytic domain in the middle (M‐kinesin, e.g. kinesin‐13). C‐kinesin
an unusual member of the hexameric AAA+ superfamily drives minus end‐directed motility of cargos, while M‐kinesin
(ATPase associated with various cellular activities). The two depolymerizes MTs [55, 61]. Similar to dynein, the net direction
motor domains of DHC interact with MTs. Each motor of kinesin‐driven cargo movement depends on the MT orientation
domain contains six AAA+ sub‐domains, two of which within a particular cell type.
(AAA1 and AAA3) can bind and hydrolyze ATP [40, 41].
Force is generated by a shift in the linker domain that joins the
dynein tail with AAA1 [42]. Intermediate chains are thought
to bind dynein to its many cargos, such as vesicles, Golgi
MICROTUBULE ORGANIZATION,
body, lipid droplets, kinetochores, and mRNA [43]. Recent POLARIZATION, AND CARGO
studies suggest that different adaptor proteins also mediate TRAFFICKING IN HEPATOCYTES
the cargo binding and motor activities of dynein. Jun N‐termi­
nal kinase interacting proteins (JIP), Rab7‐interacting lysoso­ Like other epithelial cells, hepatocytes must exchange macro­
mal proteins RILP/p150Glued and sorting nexins are well‐known molecules between the external environment and their interior
adaptor proteins for dynein [39, 42, 44, 45]. Intriguingly, milieu. To achieve this, hepatocytes are polarized with distinct
although the dynein motor generates a smaller force than apical (or bile canalicular) and basolateral (or sinusoidal)
kinesin, it can work in large teams by virtue of an in‐built gear domains segregated by tight junctions. The lateral surfaces of
mechanism and the ability to catch‐bond to MTs against high hepatocytes form cell–cell contacts while the basal surface inter­
opposing load [46, 47]. Taking these findings further, choles­ acts with the extracellular matrix. Unlike typical epithelial cells,
terol‐enriched lipid rafts on an endosome/phagosome mem­ hepatocytes have a multipolar organization. Each hepatocyte
brane were shown to serve as a platform where many dynein shares a boundary with multiple narrow lumina, bile canaliculi,
motors can cluster together. This geometric clustering allowed and basal domains of neighboring cells that face the endothelial
a large team of dyneins to simultaneously contact a single lining. A hepatocyte sandwiched in this manner has to sustain
MT, and generate a collective force that rapidly transports the two countercurrent flow systems: the synthesis/secretion of bile
phagosome towards degradative lysosomes [48, 49]. Because and uptake/secretion of blood components at the apical and
cytoplasmic dynein transports cargos towards the minus end basolateral surfaces, respectively [1]. The two surface domains
of MTs, it is known as a retrograde motor [50]. The orienta­ exhibit distinct protein and lipid compositions [62], and suste­
tion of MTs differs in epithelial cells of different origins. nance of this complex polarity likely requires vesicular transport
Therefore, the dynein‐driven motion may localize cargos along MTs and actin inside hepatocytes.
towards the apical or the basolateral membrane, depending on Novikoff et al. imaged the MT organization in cultured ­primary
cell type. In hepatocytes, dynein likely drives cargos towards rat hepatocytes by immunofluorescence labeling of tubulin [63].
the perinuclear region [51, 52]. A “starburst” pattern, commonly identified as radially organized
3:  Cytoskeletal Motors: Structure and Function in Hepatocytes 31

Golgi

Bile
canaliculi sER

LDs

Kinesin-1
Sinusoid
Microtubule

Nucleus
Cortical actin

Figure 3.3  Microtubule organization in hepatocytes. Microtubules extend from pericanalicular region towards the sinusoidal region. Lipid drop­
lets (LDs) are driven by conventional kinesin towards peripherally located smooth ER (sER) near the plus ends of microtubules (shown). Endocytic
vesicles are driven dominantly by dynein and exocytic vesicles by kinesin motors (motors are not shown).

MTs (e.g. in melanocytes), was observed with one end of MTs vesicles around the Golgi [71]. Exposure of hepatocytes to col­
focused at a centrally located centrosome, and the other ends chicine reduces endosome–lysosome fusion, suggesting that
emanating out like an umbrella to line the cortical region of the dynein transports endocytic vesicles to the apical surface [72].
hepatocyte. In agreement with this geometry, the microtubule Fluorescently labeled early endocytic vesicles isolated from rat
organizing center (MTOC) in polarized hepatocytes is localized liver are driven along MTs by kinesin‐1 and KIFC2 motors [73].
adjacent to a perinuclear region and MTs extend from canalicu­ In contrast to the early endocytic vesicles, dynein and KIF3A
lar to sinusoidal region [64]. Cytoplasmic dynein is associated drive motion of late endosomes which do not exhibit significant
with, and may drive ligand‐containing endosomes towards the fission [74].
central centrosomal region of hepatocytes [51, 65]. A simplified
view of the microtubule orientation in relation to the apical and
basolateral membrane in hepatocytes is shown in Figure 3.3. As
a consequence of this geometry, MT‐based motion towards the MICROTUBULE MOTORS
apical PM in hepatocytes may require a plus end‐directed motor IN TRIGLYCERIDE SECRETION
(e.g. kinesin‐1). The MT orientation and directionality of motors
is also relevant to lipid trafficking. For example, the smooth The liver can be thought of as the “lipid manager” of the body.
endoplasmic reticulum (sER) (where lipids are packaged for Most of this lipid is stored away in intracellular organelles
secretion) appears located towards the ­periphery of hepatocytes called lipid droplets (LDs). LDs are loaded with triglyceride
and may be supplied with lipids via kinesin‐driven transport of (TG) molecules that consist of fatty acid chains esterified to
lipid droplets [66]. glycerol. Our view that LDs are coalesced fat sitting idle in cells
The liver is an organ of immense metabolic importance. It is has changed drastically. Proteomic and imaging studies hinted
involved in the production of bile acid, cholesterol, plasma pro­ that LDs interact with mitochondria, peroxisomes, and ER
tein, assembly/secretion of triglyceride‐laden very low‐density depending on the cell/tissue types and metabolic requirements
lipoprotein particles (VLDL), elimination of toxic substances, [75–77]. This interaction of LDs with a variety of cellular com­
and processing of hormones and cytokines [1]. All these pro­ partments requires long‐distance transport of the LDs within
cesses involve directed intracellular transport between distinct cells. Live imaging reveals that LDs move in a directed manner
subcellular compartments and membranes of a hepatocyte. The in many cell types ranging from algae to hepatocytes [66, 78].
ER–Golgi network is the central pathway in protein sorting and Cytoplasmic dynein and kinesin are physically associated with
endosomal targeting. Nascent proteins arising in the ER are LDs to drive intracellular motion of LDs [66, 79].
transported to the trans‐Golgi network (TGN) via the Golgi cis­ The TG in LDs can be utilized to generate energy in h­ epatocytes,
ternae. The proteins sequestered at TGN are actively sorted and or can be secreted in the form of VLDL. The bulk of TG (~70%)
packaged into vesicles to be further transported to the apical in VLDL is derived from cytosolic LDs [80]. The earliest report
membrane or to endosomal compartments [67]. MT motors are suggesting that MTs are involved in lipoprotein secretion from the
essential for the generation and transport of TGN‐derived liver was published in 1973 [81]. Furthermore, Reaven et al. dem­
­transport carriers [68, 69]. Disruption of MTs results in the mis­ onstrated that depolymerization of microtubules drastically
sorting of apical proteins [70]. A function‐blocking antibody decreased VLDL lipidation in hepatocytes [82]. More recently, we
against kinesin also inhibits exit of an apical reporter protein, found that LDs purified from rat liver show rapid plus end‐directed
NTRp75, from the TGN and leads to accumulation of small motion along MTs [83]. We also showed that kinesin‐1‐driven LD
32 THE LIVER:  CONCLUSIONS

transport supplies triglyceride to the smooth ER in hepatocytes for of MTs within cells. The most common and evolutionarily con­
assembling VLDL, and that knockdown of kinesin‐1 in McA‐ served modification on tubulin is acetylation. The ε‐amino
RH7777 cells inhibits the characteristic peripheral localization of group of lysine‐40 (K40) residue of α‐tubulin is predominantly
LDs [66, 84]. Intriguingly, kinesin‐1 is recruited to LDs in a com­ acetylated with αTAT1 (α‐tubulin acetyltransferase‐1) [92, 93]
plex with ADP ribosylation factor 1 (ARF1), a small GTPase and and reversed by deacetylases: Sirt2 (sirtuin type 2) [94] or his­
a key regulator of lipolysis [66]. Because ARF1 induces ­membrane tone deacetylase (HDAC6) [95]. The acetylation‐induced weak
curvature, it is possible that the additional force from kinesin lateral interactions of protofilaments increase MT flexibility and
causes vesiculation from ARF1‐rich regions on the LD membrane. make MTs resistant to mild cold, nocodazole, and colchicine
ARF1 and kinesin were activated on LDs in an insulin‐dependent exposure [96, 97]. Besides acetylation, the K40 site is also sus­
manner, making this pathway responsive to the metabolic state of ceptible to trimethylation [98]. In mitotic cells, α‐tubulin of the
an animal. In the fed state kinesin‐1 was more active on LDs, central spindle is trimethylated at K40, and enriched at the plus
transporting them to the peripherally localized sER in hepato­ ends of MTs at metaphase. Despite numerous reports on the
cytes, where sER‐resident lipases could catabolize the TG to gen­ phosphorylation of α‐ and β‐tubulin at serine, threonine, and
erate diacylglycerol [85]. Because it has two fatty acid chains, tyrosine residues, the physiological significance of these modi­
diacylglycerol may equilibrate across the sER membrane, where­ fications remains to be understood. The unstructured C‐terminal
after it is reconverted into TG in the sER lumen and supplied to tail (CTT) domains of α‐ and β‐tubulin are frequent targets of
synthesize mature VLDL particles [86]. tyrosination, Δ2 modification (penultimate glutamate is also
The catabolism of LD‐TG for VLDL lipidation is extremely removed from detyrosinated tubulin) and detyrosination.
efficient [80] and may require a synergy between kinesin, ARF1, α‐Tubulin undergoes tyrosination–detyrosination cycles via the
and specific LD/sER‐associated lipids in the fed state. In con­ activity of tubulin tyrosinase ligase and carboxypeptidase. An
trast, in the fasted state, lowered insulin‐signaling reduces kine­ in vitro study suggests that PTM of tubulin regulates the MT
sin‐1 on LDs to limit sER–LD contact and tempers TG supply depolymerization rate. Moreover, modifications in CTT of tubu­
for VLDL lipidation [66]. Tuning kinesin‐dependent transport lin are recognized by molecular motors that also govern the
of LDs in hepatocytes may therefore enable the liver to protec­ motor velocity and its processivity [99]. In vivo, detyrosinated
tively sequester massive amounts of TG after fasting (fasting‐ MTs persist for hours, but tyrosinated MTs turn over in minutes
induced steatosis), thus preventing lipotoxic effects of TG on [100]. The KIF2C kinesin that depolymerizes MTs has prefer­
peripheral tissues. The cellular mechanisms that control TG ential activity for tyrosinated MTs both in vivo and in vitro. The
secretion from the liver, and therefore systemic TG homeostasis binding of KIF5 kinesin to MTs is also regulated by multiple
across daily metabolic cycles, can help understand TG imbal­ PTMs. The presence of tyrosinated tubulin also favors interac­
ance leading to lipodystrophies such as fatty liver and diabetes tion of the MTs with plus end‐tracking proteins (+TIPs) that
[87]. Indeed, we have seen significant changes in lipid/protein have conserved CAP‐Gly domains [101]. Some tubulin modifi­
trafficking to LDs across feeding–fasting cycles [66, 88]. It is cation, such as glutamylation, succination, and O‐Glc‐N‐acylation
surprising that relatively little work has been done employing have been recently identified, but their detailed physiological
animal models and the simple feeding–fasting response. We impact is as yet unexplored.
also hope that disease‐relevant mutations in animal models can Alcoholic liver may develop as a result of abnormal PTMs in
answer how motor proteins transport LDs within hepatocytes to hepatocytes. Acetylation of MTs in hepatocytes increases tubu­
control trafficking in response to metabolic states. lin stability [102]. Treatment of hepatoma‐derived WIF‐B cells
MT‐dependent transport is also essential for hepatitis C virus with a drug that specifically depolymerizes dynamic MTs
(HCV) replication in host hepatocytes. HCV replicates its impairs MT‐dependent trafficking along three distinct cellular
genome at the ER‐derived membranous web, and assembles new pathways [103, 104]. Tubulin is hyperacetylated in ethanol‐fed
infectious particles using ER resident structural proteins [89]. rats, which increases MT stability and might also alter intracel­
Newly synthesized viral proteins (e.g. HCV‐core) must transfer lular motility [105]. Acetylated MTs are required for adipogen­
from ER to LDs for HCV to propagate. This transfer and HCV esis, and this acetylation is controlled by AMPK‐mediated
assembly was diminished upon kinesin‐1 knockdown [66]. phosphorylation of αTAT1 [106]. AMPK, a master regulator of
HCV‐core induces LD redistribution in an MT‐ and dynein‐ lipid homeostasis, also promotes relocation of LDs and mito­
dependent manner [90]. HCV also cleaves the Rab‐interacting chondria on detyrosinated MTs in VERO cells of nonhepatic
lysosomal protein to redirect Rab7‐containing vesicles from origin to promote fatty acid oxidation [107]. Despite its poten­
dynein‐ to kinesin‐dependent transport, and in this manner pro­ tial physiological consequences, the effect of alcohol‐induced
motes virion secretion [91]. MT‐dependent transport, therefore, acetylation on LD dynamics in hepatocytes is not known.
impacts multiple aspects of the viral lifecycle in hepatocytes, and
is a rich area for further scientific investigations.
CONCLUSIONS
MICROTUBULE MODIFICATIONS Motor‐dependent motion along actin and MTs sustains the
AND LIVER PATHOLOGY intracellular organization of many cell types, and is particularly
relevant to polarized epithelial cells such as hepatocytes.
MTs and their associated functions are tuned by the post‐trans­ In  hepatocytes, the MTs are organized in a radial fashion
lational modification (PTM) of tubulin. PTMs and various ­originating from the centrosome and spreading out towards the
microtubule‐interacting proteins contribute to the heterogeneity peripheral sinusoidal regions [108, 109]. This arrangement of
3:  Cytoskeletal Motors: Structure and Function in Hepatocytes 33

MTs, along with the directionality of cytoplasmic dynein and 16. Klomp, L.W.J., Vargas, J.C., Van Mil, S.W.C. et al. Characterization of muta­
kinesin motors, defines endo/exocytosis and protein trafficking tions in ATP8B1 associated with hereditary cholestasis. Hepatology,
2004;40(1):27–38.
in hepatocytes [63]. As an example, kinesin‐dependent lipid 17. Nyasae, L.K., Schell, M.J., and Hubbard, A.L. Copper directs ATP7B to
droplet transport maintains lipid homeostasis across metabolic the  apical domain of hepatic cells via basolateral endosomes. Traffic,
transitions. Viruses (such as HCV) may also require MT‐ 2014;15(12):1344–65.
dependent transport for assembly in hepatocytes and secretion 18. Tuma, P.L., Nyasae, L.K., and Hubbard, A.L. Nonpolarized cells selectively
sort apical proteins from cell surface to a novel compartment, but lack apical
to the blood plasma in complex with VLDL particles [90, 110].
retention mechanisms. Mol Biol Cell, 2002;13(10):3400–15.
PTM of tubulin affects the dynamics and organization of MTs, 19. Manton, I. and Clarke, B. An electron microscope study of the spermatozoid
and also motor‐driven transport on MTs [99]. PTMs regulate of sphagnum. J Exp Bot, 1952;3(3):265–75.
intracellular cargo transport and organelle–organelle contacts, 20. Fawcett, D.W. and Porter, K.R. A study of the fine structure of ciliated
and aberrant PTMs might hinder substrate delivery leading to ­epithelia. J Morphol, 1954;94(2):221–81.
21. Sabitini, D.D., Bensch, K., and Barrnett, R.J. Cytochemistry and electron
physiological anomalies such as fatty liver disease. In conclu­
microscopy. The preservation of cellular ultrastructure and enzymatic
sion, the MT cytoskeleton in hepatocytes provides a rich and ­activity by aldehyde fixation. J Cell Biol, 1963;17:19–58.
very disease‐relevant context for further investigations. Of par­ 22. Ledbetter, M.C. and Porter, K.R. Morphology of microtubules of plant cell.
ticular interest in this context is the recent finding that MT Science, 1964;144(3620):872–4.
motors control lipid flux from the liver across metabolic cycles 23. Borisy, G.G. and Taylor, E.W. The mechanism of action of colchicine.
Binding of colchicine‐3H to cellular protein. J Cell Biol, 1967;34:525–33.
[66]. How cargos are handed over between the actin and MT 24. Bryan, J. and Wilson, L. Are cytoplasmic microtubules heteropolymers?
systems in hepatocytes is also poorly understood and warrants Proc Natl Acad Sci U S A, 1971;68(8):1762–6.
investigation. 25. Mohri, H. Amino‐acid composition of “tubulin” constituting microtubules of
sperm flagella. Nature, 1968;217:1053–4.
26. Borisy, G.G., Olmsted, J.B., and Klugman. R.A. In vitro aggregation of cyto­
plasmic microtubule subunits. Proc Natl Acad Sci U S A, 1972;69(10):2890–4.
27. Oakley, B.R. γ‐Tubulin: the microtubule organizer? Trends Cell Biol,
REFERENCES 1992;2(1):1–5.
28. Mitchison, T. and Kirschner, M. Dynamic instability of microtubule growth.
1. Treyer, A. and Müsch, A. Hepatocyte polarity. Compr Physiol, Nature, 1984;312(5991):237–42.
2013;3(1):243–87. 29. Matov, A., Applegate, K., Kumar, P. et al. Analysis of microtubule dynamic
2. Musch, A. Microtubule organization and function in epithelial cells. Traffic, instability using a plus‐end growth marker. Nat Methods, 2010;7(9):761–8.
2004;5:1–9. 30. Downing, K.H. and Nogales, E. Tubulin and microtubule structure. Curr
3. Feracci, H., Connolly, T.P., Margolis, R.N., and Hubbard, A.L. The estab­ Opin Cell Biol, 1998;10:16–22.
lishment of hepatocyte cell surface polarity during fetal liver development. 31. McNally, F.J. and Vale, R.D. Identification of katanin, an ATPase that severs
Dev Biol, 1987;123(1):73–84. and disassembles stable microtubules. Cell, 1993;75(3):419–29.
4. Ishii, M., Washioka, H., Tonosaki, A., and Toyota, T. Regional orientation of 32. Dehmelt, L. and Halpain, S. The MAP2/Tau family of microtubule‐­
actin filaments in the pericanalicular cytoplasm of rat hepatocytes. associated proteins. Genome Biol, 2005;6(1):204.
Gastroenterology, 1991;101(6):1663–72. 33. Gibbons, I.R. and Rowe, A.J. Dynein: a protein with adenosine ­triphosphatase
5. Tzanakakis, E.S., Hansen, L.K., and Hu, W.S. The role of actin filaments and activity from cilia. Science, 1965;149(3682):424–6.
microtubules in hepatocyte spheroid self‐assembly. Cell Motil Cytoskeleton, 34. Gibbons, B.H. and Gibbons, I.R. Flagellar movement and adenosine triphos­
2001;48(3):175–89. phatase activity in sea urchin sperm extracted with triton X‐100. J Cell Biol,
6. Gupta, A., Schell, M.J., Bhattacharjee, A., Lutsenko, S., and Hubbard, A.L. 1972;54(1):75–97.
Myosin Vb mediates Cu+ export in polarized hepatocytes. J Cell Sci, 35. Allen, R.D., Metuzals, J., Tasaki, I., Brady, S.T., and Gilbert, S.P. Fast axonal
2016;129(6):1179–89. transport in squid giant axon. Science, 1982;218(4577):1127–9.
7. Hales, C.M., Griner, R., Hobdy‐Henderson, K.C. et  al. Identification and 36. Brady, S.T., Lasek, R.J., and Allen, R.D. Fast axonal transport in extruded
Characterization of a Family of Rab11‐interacting Proteins. J Biol Chem, axoplasm from squid giant axon. Science, 1982;218(4577):1129–31.
2001;276(42):39067–75. 37. Schroer, T.A., Schnapp, B.J., Reese, T.S., and Sheetz, M.P. The role of kine­
8. Lapierre, L.A., Kumar, R., Hales, C.M. et al. Myosin vb is associated with sin and other soluble factors in organelle movement along microtubules.
plasma membrane recycling systems. Mol Biol Cell, 2001;12(6):1843–57. J Cell Biol, 1988;107(5):1785–92.
9. Müller, T., Hess, M.W., Schiefermeier, N. et  al. MY05B mutations cause 38. Schnapp, B.J. and Reese, T.S. Dynein is the motor for retrograde axonal
microvillus inclusion disease and disrupt epithelial cell polarity. Nat Genet, transport of organelles. Proc Natl Acad Sci U S A, 1989;86(5):1548–52.
2008;40(10):1163–5. 39. Allan, V.J. Cytoplasmic dynein. Biochem Soc Trans, 2011;39(5):1169–78.
10. Wakabayashi, Y., Dutt, P., Lippincott‐Schwartz, J., and Arias, I.M. Rab11a 40. Gibbons, I.R., Lee‐Eiford, A., Mocz, G. et al. Photosensitized cleavage of
and myosin Vb are required for bile canalicular formation in WIF‐B9 cells. dynein heavy chains. Cleavage at the “V1 site” by irradiation at 365 nm in
Proc Natl Acad Sci U S A, 2005;102(42):15087–92. the presence of ATP and vanadate. J Biol Chem, 1987;262(6):2780–6.
11. Ruemmele, F.M., Müller, T., Schiefermeier, N. et  al. Loss‐of‐function of 41. Reck‐Peterson, S.L. and Vale, R.D. Molecular dissection of the roles of
MY05B is the main cause of microvillus inclusion disease: 15 Novel muta­ nucleotide binding and hydrolysis in dynein’s AAA domains in Saccharomyces
tions and a CaCo‐2 RNAi cell model. Hum Mutat, 2010;31(5):544–51. cerevisiae. Proc Natl Acad Sci U S A, 2004;101(6):1491–5.
12. Knowles, B.C., Roland, J.T., Krishnan, M. et al. Myosin Vb uncoupling from 42. Kardon, J.R. and Vale, R.D. Regulators of the cytoplasmic dynein motor. Nat
RAB8A and RAB11A elicits microvillus inclusion disease. J Clin Invest, Rev Mol Cell Biol, 2009;10:854–65.
2014;124(7):2947–62. 43. Vallee, R.B., Williams, J.C., Varma, D., and Barnhart, L.E. Dynein: an
13. Kravtsov, D., Mashukova, A., Forteza, R. et al. Myosin 5b loss of function ancient motor protein involved in multiple modes of transport. J Neurobiol,
leads to defects in polarized signaling: implication for microvillus inclusion 2004;58:189–200.
disease pathogenesis and treatment. Am J Physiol Liver Physiol, 44. Johansson, M., Rocha, N., Zwart, W. et al. Activation of endosomal dynein
2014;307(10):G992–1001. motors by stepwise assembly of Rab7‐RILP‐p150Glued, ORP1L, and the
14. Girard, M., Lacaille, F., Verkarre, V. et al. MY05B and bile salt export pump receptor βIII spectrin. J Cell Biol, 2007;176(4):459–71.
contribute to cholestatic liver disorder in microvillous inclusion disease. 45. Fu, M.M. and Holzbaur, E.L. Integrated regulation of motor‐driven organelle
Hepatology, 2014;60(1):301–10. transport by scaffolding proteins. Trends Cell Biol, 2014;24:564–74.
15. Strautnieks, S.S., Bull, L.N., Knisely, A.S. et al. A gene encoding a liver‐­ 46. Mallik, R., Carter, B.C., Lex, S.A., King, S.J., and Gross, S.P. Cytoplasmic
specific ABC transporter is mutated in progressive familial intrahepatic dynein functions as a gear in response to load. Nature, 2004;427(6975):
­cholestasis. Nat Genet, 1998;20(3):233–8. 649–52.
34 THE LIVER:  REFERENCES

47. Rai, A.K., Rai, A., Ramaiya, A.J., Jha, R., and Mallik, R. Molecular adapta­ 73. Bananis, E., Murray, J.W., Stockert, R.J., Satir, P., and Wolkoff, A.W.
tions allow dynein to generate large collective forces inside cells. Cell, Regulation of early endocytic vesicle motility and fission in a reconstituted
2013;152(1–2):172–82. system. J Cell Sci, 2003;116(Pt 13):2749–61.
48. Anishkin, A. and Kung, C. Stiffened lipid platforms at molecular force foci. 74. Bananis, E., Nath, S., Gordon, K. et al. Microtubule‐dependent movement of
Proc Natl Acad Sci U S A, 2013;110(13):4886–92. late endocytic vesicles in vitro: requirements for dynein and kinesin. Mol
49. Rai, A., Pathak, D., Thakur, S. et al. Dynein clusters into lipid microdomains Biol Cell, 2004;15(8):3688–97.
on phagosomes to drive rapid transport toward lysosomes. Cell, 75. Brasaemle, D.L., Dolios, G., Shapiro, L., and Wang, R. Proteomic analysis
2016;164(4):722–34. of proteins associated with lipid droplets of basal and lipolytically stimulated
50. Hirokawa, N. Kinesin and dynein superfamily proteins and the mechanism 3T3‐L1 adipocytes. J Biol Chem, 2004;279(45):46835–42.
of organelle transport. Science, 1998;279:519–26. 76. Yang, L., Ding, Y., Chen, Y. et al. The proteomics of lipid droplets: structure,
51. Oda, H., Stockert, R.J., Collins, C. et  al. Interaction of the microtubule dynamics, and functions of the organelle conserved from bacteria to humans.
cytoskeleton with endocytic vesicles and cytoplasmic dynein in cultured rat J Lipid Res, 2012;53(7):1245–53.
hepatocytes. J Biol Chem, 1995;270(25):15242–9. 77. Gao, Q. and Goodman, J.M. The lipid droplet – a well‐connected organelle.
52. Wang, Y. Cytoplasmic dynein participates in apically targeted stimulated Front Cell Dev Biol, 2015;3.
secretory traffic in primary rabbit lacrimal acinar epithelial cells. J Cell Sci, 78. Guimaraes, S.C., Schuster, M., Bielska, E. et al. Peroxisomes, lipid droplets,
2003;116(10):2051–65. and endoplasmic reticulum “hitchhike” on motile early endosomes. J Cell
53. Vale, R.D., Reese, T.S., and Sheetz, M.P. Identification of a novel force‐ Biol, 2015;211(5):945–54.
generating protein, kinesin, involved in microtubule‐based motility. Cell, 79. Boström, P., Rutberg, M., Ericsson, J. et al. Cytosolic lipid droplets increase
1985;42(1):39–50. in size by microtubule‐dependent complex formation. Arterioscler Thromb
54. Brady, S.T., Lasek, R.J., and Allen, R.D. Fast axonal transport in extruded Vasc Biol, 2005;25(9):1945–51.
axoplasm from squid giant axon. Science 1982;218(4577):1129–31. 80. Wiggins, D. and Gibbons, G.F. The lipolysis/esterification cycle of hepatic tria­
55. Hirokawa, N., Noda, Y., Tanaka, Y., and Niwa, S. Kinesin superfamily cylglycerol. Its role in the secretion of very‐low‐density lipoprotein and its
motor  proteins and intracellular transport. Nat Rev Mol Cell Biol, response to hormones and sulphonylureas. Biochem J, 1992;284(Pt 2):457–62.
2009;10:682–96. 81. Orci, L., Marchand, Y.L.E., Singh, A., Assimacopoulos‐Jennnet, F., Rouiller,
56. Kuznetsov, S.A., Vaisberg, E.A., Shanina, N.A. et al. The quaternary struc­ C., and Jeanrenaud, B. Role of microtubules in lipoprotein secretion by the
ture of bovine brain kinesin. EMBO J, 1988;7(2):353–6. liver. Nature, 1973;244:30.
57. Bloom, G.S., Wagner, M.C., Pfister, K.K., and Brady, S.T. Native structure 82. Reaven, E.P. and Reaven, G.M. Evidence that microtubules play a permis­
and physical properties of bovine brain kinesin and identification of the ATP‐ sive role in hepatocyte very low density lipoprotein secretion. J Cell Biol,
binding subunit polypeptide. Biochemistry, 1988;27(9):3409–16. 1980;84(1):28–39.
58. Asenjo, A.B., Krohn, N., and Sosa H. Configuration of the two kinesin motor 83. Barak, P., Rai, A., Rai, P., and Mallik, R. Quantitative optical trapping on
domains during ATP hydrolysis. Nat Struct Biol, 2003;10(10):836–42. single organelles in cell extract. Nat Methods, 2013;10(1):68–70.
59. Yildiz, A., Tomishige, M., Vale, R.D., and Selvin, P.R. Kinesin walks hand‐ 84. Schulze, R.J. and McNiven, M.A. Fasting inhibits the recruitment of kine­
over‐hand. Science, 2004;303(5658):676 LP‐678. sin‐1 to lipid droplets and stalls hepatic triglyceride secretion. Hepatology,
60. Yun, M., Bronner, C.E., Park, C.G. et al. Rotation of the stalk/neck and one 2019;69(1):444–6.
head in a new crystal structure of the kinesin motor protein, Ncd. EMBO J, 85. Gilham, D., Alam, M., Gao, W., Vance, D.E., and Lehner, R. Triacylglycerol
2003;22(20):5382–9. hydrolase is localized to the endoplasmic reticulum by an unusual retrieval
61. Noda, Y., Sato‐Yoshitake, R., Kondo, S., Nangaku, M., and Hirokawa, N. sequence where it participates in VLDL assembly without utilizing VLDL
KIF2 is a new microtubule‐based anterograde motor that transports membra­ lipids as substrates. Mol Biol Cell, 2005;16(2):984–96.
nous organelles distinct from those carried by kinesin heavy chain or 86. Lankester, D.L., Brown A.M., and Zammit, V.A. Use of cytosolic triacylg­
KIF3A/B. J Cell Biol, 1995;129(1):157–67. lycerol hydrolysis products and of exogenous fatty acid for the synthesis
62. Rosario, J., Sutherland, E., Zaccaro, L., and Simon, F.R. Ethinylestradiol of  triacylglycerol secreted by cultured rat hepatocytes. J Lipid Res,
administration selectively alters liver sinusoidal membrane lipid fluidity and 1998;39(9):1889–95.
protein composition. Biochemistry, 1988;27(11):3939–46. 87. Anderwald, C., Bernroider, E., Krssak, M. et al. Effects of insulin treatment
63. Novikoff, P.M., Cammer, M., Tao, L. et al. Three‐dimensional organization in type 2 diabetic patients on intracellular lipid content in liver and skeletal
of rat hepatocyte cytoskeleton: relation to the asialoglycoprotein endocytosis muscle. Diabetes, 2002;51(10):3025–32.
pathway. J Cell Sci, 1996;109(Pt 1):21–32. 88. Sadh, K., Rai, P., and Mallik, R. Feeding‐fasting dependent recruitment of
64. Murray, J.W. and Wolkoff, A.W. Roles of the cytoskeleton and motor pro­ membrane microdomain proteins to lipid droplets purified from the liver.
teins in endocytic sorting. Adv Drug Deliv Rev, 2003;55:1385–403. PLoS One, 2017;12(8).
65. Goltz, J.S., Wolkoff, A.W., Novikoff, P.M., Stockert, R.J., and Satir P. A role 89. Scheel, T.K.H. and Rice, C.M. Understanding the hepatitis C virus life cycle
for microtubules in sorting endocytic vesicles in rat hepatocytes. Proc Natl paves the way for highly effective therapies. Nat Med, 2013;19:837–49.
Acad Sci U S A, 1992;89(15):7026–30. 90. Boulant, S., Douglas, M.W., Moody, L. et al. Hepatitis C virus core protein
66. Rai, P., Kumar, M., Sharma, G. et al. Kinesin‐dependent mechanism for con­ induces lipid droplet redistribution in a microtubule‐ and dynein‐dependent
trolling triglyceride secretion from the liver. Proc Natl Acad Sci U S A, manner. Traffic, 2008;9(8):1268–82.
2017;114(49):12958–63. 91. Wozniak, A.L., Long, A., Jones‐Jamtgaard, K.N., and Weinman, S.A.
67. Allan, V.J., Thompson, H.M., and McNiven, M.A. Motoring around the Hepatitis C virus promotes virion secretion through cleavage of the Rab7
Golgi. Nat Cell Biol, 2002;4:E236–42. adaptor protein RILP. Proc Natl Acad Sci U S A, 2016;201607277.
68. Van der Sluijs, P., Bennett, M.K., Antony, C., Simons, K., and Kreis, T.E. 92. LeDizet, M. and Piperno, G. Identification of an acetylation site of
Binding of exocytic vesicles from MDCK cells to microtubules in vitro. J Chlamydomonas alpha‐tubulin. Proc Natl Acad Sci U S A, 1987;84(16):5720–4.
Cell Sci, 1990;95(Pt 4):545–53. 93. Shida, T., Cueva, J.G., Xu, Z., Goodman, M.B., and Nachury, M.V. The
69. Fath, K.R., Trimbur, G.M., and Burgess, D.R. Molecular motors and a spec­ major alpha‐tubulin K40 acetyltransferase TAT1 promotes rapid ciliogenesis
trin matrix associate with Golgi membranes in vitro. J Cell Biol, 1997;139(5): and efficient mechanosensation. Proc Natl Acad Sci U S A, 2010;107(50):
1169–81. 21517–22.
70. Kreitzer, G., Schmoranzer, J., Low, S.H. et al. Three‐dimensional analysis of 94. North, B.J., Marshall, B.L., Borra, M.T., Denu, J.M., and Verdin, E. The
post‐Golgi carrier exocytosis in epithelial cells. Nat Cell Biol, 2003;5:126–36. human Sir2 ortholog, SIRT2, is an NAD+‐dependent tubulin deacetylase.
71. Kreitzer, G., Marmorstein, A., Okamoto, P., Vallee, R., and Rodriguez‐ Mol Cell, 2003;11(2):437–44.
Boulan, E. Kinesin and dynamin are required for post‐Golgi transport of a 95. Hubbert, C., Guardiola, A., Shao, R. et  al. HDAC6 is a microtubule‐­
plasma‐membrane protein. Nat Cell Biol, 2000;2(2):125–7. associated deacetylase. Nature, 2002;417(6887):455–8.
72. Berg, T., Kindberg, G.M., Ford, T., and Blomhoff, R. Intracellular transport 96. Xu, Z., Schaedel, L., Portran, D. et  al. Microtubules acquire resistance
of asialoglycoproteins in rat hepatocytes. Evidence for two subpopulations from  mechanical breakage through intralumenal acetylation. Science,
of lysosomes. Exp Cell Res, 1985;161(2):285–96. 2017;356(6335):328–32.
3:  Cytoskeletal Motors: Structure and Function in Hepatocytes 35

97. LeDizet, M. and Piperno G. Cytoplasmic microtubules containing acety­ 104. Phung‐Koskas, T., Pilon, A., Poüs, C. et  al. STAT5B‐mediated growth
lated  ??‐tubulin in Chlamydomonas reinhardtii: spatial arrangement and ­hormone signaling is organized by highly dynamic microtubules in hepatic
properties. J Cell Biol, 1986;103(1):13–22. cells. J Biol Chem, 2005;280(2):1123–31.
98. Park, I.Y., Powell, R.T., Tripathi, D.N. et al. Dual chromatin and cytoskel­ 105. Kannarkat, G.T., Tuma, D.J., and Tuma, P.L. Microtubules are more stable
etal remodeling by SETD2. Cell, 2016;166(4):950–62. and more highly acetylated in ethanol‐treated hepatic cells. J Hepatol,
99. Sirajuddin, M., Rice, L.M., and Vale, R.D. Regulation of microtubule 2006;44(5):963–70.
motors by tubulin isotypes and post‐translational modifications. Nat Cell 106. Mackeh, R., Lorin, S., Ratier, A. et  al. Reactive oxygen species, amp‐
Biol, 2014;16:335. Activated protein kinase, and the transcription cofactor p300 regulate α‐
100. Kreitzer, G., Liao, G., and Gundersen, G.G. Detyrosination of tubulin Tubulin acetyltransferase‐1 (αtat‐1/mec‐17)‐dependent microtubule
regulates the interaction of intermediate filaments with microtubules in hyperacetylation during cell stress. J Biol Chem, 2014;289(17):11816–28.
vivo via a kinesin‐dependent mechanism. Mol Biol Cell, 1999;10(April): 107. Herms, A., Bosch, M., Reddy, B.J.N. et al. AMPK activation promotes lipid
1105–18. droplet dispersion on detyrosinated microtubules to increase mitochondrial
101. Peris, L., Thery, M., Fauré, J. et al. Tubulin tyrosination is a major factor fatty acid oxidation. Nat Commun, 2015;6:7176.
affecting the recruitment of CAP‐Gly proteins at microtubule plus ends. J 108. Ihrke, G., Neufeld, E.B., Meads T. et al. WIF‐B cells: An in vitro model for
Cell Biol, 2006;174(6):839–49. studies of hepatocyte polarity. J Cell Biol, 1993;123(6 II):1761–75.
102. Groebner, J.L. and Tuma PL. The altered hepatic tubulin code in alcoholic 109. Meads, T. and Schroer, T.A. Polarity and nucleation of microtubules in
liver disease. Biomolecules, 2015;5:2140–59. polarized epithelial cells. Cell Motil Cytoskeleton, 1995;32(4):273–88.
103. Poüs, C., Chabin, K., Drechou, A. et al. Functional specialization of stable 110. Coller, K.E., Heaton, N.S., Berger, K.L. et al. Molecular determinants
and dynamic microtubules in protein traffic in WIF‐B cells. J Cell Biol, and dynamics of hepatitis C virus secretion. PLOS Pathog, 2012;8(1):
1998;142(1):153–65. e1002466.
Hepatocyte Surface Polarity
4 Anne Müsch1 and Irwin M. Arias2
1

2
Department of Developmental & Molecular Biology, Albert Einstein College of Medicine, Bronx, NY, USA
National Institutes of Health, Bethesda, MD, USA

INTRODUCTION elements, including extracellular matrix (ECM), adherens and


tight junctions, intracellular protein trafficking machinery,
With exception of erythrocytes and possibly a few others, all cytoskeleton, and energy production. Inheritable and acquired
mammalian cells are polarized. Epithelial cell polarity, as pre- defects in these elements that result in depolarization or failure
sent in the liver, gastrointestinal tract, and kidney, developed to polarize are presented in Chapter 29.
during the Precambrian Period and is essential for the formation
of multicellular organs and species. Selective absorption and
secretion requires cellular polarization. Hepatocyte polarization
is neglected, unique, complex, and, as shown in recent studies, THE UNIQUE HEPATOCYTE POLARITY
critical in many inheritable and acquired liver diseases. PHENOTYPE
Surprisingly, only one current text in hepatology or pathol-
ogy [1] mentions hepatocellular polarity. Neither the authors The liver presents a remarkable example of how cell shape
nor a large number of hepatologists or pathologists recall seeing serves function. The two liver epithelial cell types – hepatocytes
a pathology report which comments on the polarization status of and bile duct cells  –  adopt radically different polarity pheno-
hepatocytes! A possible explanation is that hepatology devel- types that serve their distinct physiological roles. Bile duct cells,
oped in the late nineteenth century, when efforts were always which form simple conduits for bile, organize like other tubule‐
made to associate pathology with clinical signs, symptoms, and forming epithelial cells with a single luminal domain perpen-
outcome. The major tool in this effort was hematoxylin and dicular to cell–cell adhesion domains and opposite a basal
eosin (H&E)‐stained sections of liver. The H&E stain does not surface; they are monopolar. Hepatocytes, which mediate exten-
reveal the small bile canaliculus, which is the apical polarized sive bidirectional molecular exchange with the blood, maximize
domain of the hepatocyte. The bile canaliculus is functionally contact with blood vessels by establishing a second basal sur-
sealed by tight junctions and, with its microvilli, constitutes face in place of the biliary cell’s apical surface. The hepatocyte
~13% of total hepatocyte plasma membrane [1]. Discovery of apical domains assemble instead at cell–cell contact sites,
the bile canaliculus did not occur until transmission and scan- interrupting cell–cell adhesion domains (Figure  4.1). Each
­
ning electron microscopy were introduced in the 1940s. Despite hepatocyte forms two or more lumina with several of its neigh-
the availability of numerous bile canalicular proteins and asso- bors, yielding the branched luminal network known as bile
ciated antibodies, such as 5′ nucleotidase, cCAM 105, ABC B1, canaliculi. Hepatocytes are therefore multipolar. While the
­
ABC B11, and others, they are rarely used in routine studies of monopolar organization of bile duct cells is similar to that of
liver histopathology. Probably because polarity is not described most epithelia, the hepatocyte polarity phenotype is unique.
by pathologists, it is also neglected by clinicians. Immunohistochemistry, electron microscopy, and gene
This chapter will consider mechanisms responsible for hepa- expression analysis collectively indicate that hepatocytes form
tocellular polarity. Canalicular network formation and hepato- functional cell–cell adherens and tight junctions, and that they
cyte polarity require coordinated expression of several key express and localize the constituents of evolutionary conserved

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
4:  Hepatocyte Surface Polarity 37

(a)
Basal
Hepatocytes
Lateral
Luminal/apical

Bile ducts

(b) (A) (B)

HA 321
HA4

10

(C) (D) Sinusoidal


front
BC

Golgi
complex
BC
High
concentrations
of endo/exocytotic BC
organelles BC

Figure 4.1  (a) The two polarity phenotypes in the liver – hepatocytes and bile duct epithelia – have three distinct membrane domains: basal
domains facing the extracellular matrix and underlying blood vessels; lateral domains engaged in cell–cell adhesion; and luminal domains, which
face the outside world (they are situated at the cell apex in monopolar cells and hence also called apical domains). Hepatocytes have multiple lumi-
nal surfaces that interrupt cell–cell adhesion domains and most feature two basal surfaces facing the space of Disse on either side. Bile duct epithe-
lial cells each have a single luminal and basal domain flanked by lateral surfaces. Adapted from Müsch, Curr Opin Cell Biol, 2018;54:18–23 with
permission of Elsevier. (b) The architecture of the liver is unique. A: A scanning electron micrograph of a portion of a liver lobule is shown. A
continuous network of bile canaliculi runs along the exposed cell surfaces of the liver plate. B: The two distinct PM domains are visualized by
immunofluorescence detection of the basolateral PM protein, HA321/BEN and the apical PM protein, HA4/cell‐CAM105/ectoATPase. An electron
micrograph of a hepatocyte (C) and a corresponding schematic drawing (D) highlight the “active zones” in vesicular trafficking. The major sorting
organelles (TGN and endosomes) and transport vesicles are concentrated in small “clear” zones that are probably the most active in vesicle traffic.
These zones are located between the Golgi and the apical PM and near the basolateral PM (the shaded regions in D). Reproduced from Braiterman
and Hubbard, The Liver, 5th edn, Fig. 6.1.
38 THE LIVER:  EXPERIMENTAL SYSTEMS FOR THE STUDY OF EPITHELIAL POLARIZATION IN THE LIVER

polarity complexes, which establish epithelial surface identities, branching morphogenesis in the lung [13]. Future work is
in a similar manner to that of monopolar epithelial cells. These expected to determine whether a taurocholate AMPK signaling
observations make it likely that the distinct polarity phenotypes cascades also governs bile duct branching.
are caused by different regulation of the same key polarization
mechanisms. Identification of regulatory mechanisms that
define the two polarity phenotypes is only beginning to emerge.
For a comprehensive listing of polarity‐associated proteins,
EXPERIMENTAL SYSTEMS FOR THE
their functions, and expression in hepatocytes, see Braiterman STUDY OF EPITHELIAL POLARIZATION
and Hubbard, The Liver, 5th edn, Table 6.2. IN THE LIVER
Due to lack of a readily available stable culture system for pri-
mary hepatocytes, much of what has been learned about hepato-
ESTABLISHMENT OF CONNECTED BILE cyte polarization comes from polarized hepatocytic cell lines, in
particular from Hep G2 and WIF‐B cells, which are derived
DUCT AND BILE CANALICULAR
from cancers [14–17]. Both form spherical lumina between
NETWORKS DURING LIVER neighboring cells but do not develop a canalicular network. The
DEVELOPMENT recently established Can10 line, which, like WIF‐B, is derived
from a rat hepatoma, is the first hepatocytic cell line in which
The two liver epithelial cells originate from common precur- individual lumina connect to form canaliculi; however, it has not
sors, called hepatoblasts. Hepatoblasts delaminate from a mon- yet been extensively characterized [18]. HepG2 and WIF‐B
olayered epithelial tube, the foregut, proliferate and invade the cells recapitulate some critical hallmarks of the hepatocytic
surrounding mesenchyme [2, 3] (see Chapter  2 for more polarity phenotype observed in primary cultures or in vivo, such
details). Hepatoblasts around the portal vein first form a mon- as the gradual, nonsynchronous formation of luminal surfaces at
olayered epithelium, the ductal plate  [4]. Two prominent cell–cell contact sites, which contrast with the rapid and syn-
features distinguish this from the adjacent hepatoblasts: a
­ chronous polarization of monopolar epithelial cells in culture,
laminin‐positive basement membrane and strong E‐cadherin the asymmetric cell divisions subsequently discussed, and the
labeling at cell–cell contact sites. Parts of the ductal plate indirect targeting of apical single membrane‐spanning and GPI‐
develop into bile ducts, while the remnants have been proposed anchored proteins. HepG2 polarization is sensitive to oncostatin
to become hepatic stem cells [5, 6]. Bile ducts develop when M, a cytokine critical for hepatocyte differentiation in vivo [19];
the ductal plate cells signal adjacent hepatoblasts to adopt a however, neither cell line responds to taurocholic acid, the pre-
similar biliary fate. The two cell layers establish a basal lamina sumptive in vivo trigger of bile canaliculi formation. This is a
encircling them, and then generate a lumen between them. serious limitation to the study of signaling mechanisms leading
Hepatoblasts outside the vicinity of portal veins develop into to hepatocyte lumen formation. Also, the mechanism and regu-
hepatocytes. Hepatocytes acquire polarity (i.e. form bile cana- lation of ABC transporter trafficking in WIF‐B and HepG2 cells
liculi) only in late gestation, and the bile canalicular network differ in several respects from that observed in hepatocytes [20].
continues to elongate postnatally [7, 8]. The hepatocyte polari- Primary hepatocytes, isolated by liver perfusion with colla-
zation spurt coincides with the onset of bile acid synthesis genase, re‐establish polarity and the dense canalicular network
[8–10]. Remarkably, the major bile acid taurocholate stimu- found in intact liver when cultured in collagen sandwiches.
lates hepatocyte polarization in culture, suggesting that tauro- These cultures not only repolarize in a predictable manner, but
cholate serves as hepatocyte polarization trigger [11]. they also maintain function and gene expression profiles of
For bile excreted into hepatocyte bile canaliculi to reach the mature hepatocytes for up to two weeks [21–24]. This culture
bile duct, the two epithelial tissues must link their respective system, although less amenable to experimental manipulation
luminal networks into a contiguous entity. This is accomplished than cell lines, provides information directly related to hepato-
when major bile ducts branch into smaller ductules, which con- cellular processes as confirmed by subsequent gene knockout
nect to the bile canaliculi. Techniques that combined thick sec- studies subsequently described [25].
tion immunofluorescence on cleared tissue with carbon ink In addition to models for hepatocyte polarity, a few cell lines
injections into the common bile duct revealed that branching of derived from normal biliary cells have also been established
smaller ductules from the larger ducts happened in parallel with [26, 27]. These lines polarize with monopolar organization in
the spurt in bile canaliculi formation and with the first appear- 2D culture on permeable filter substrates where they acquire
ance of bile in the intestine at E18 in the mouse [12]. high trans‐epithelial resistance, a measure of epithelial barrier
Pharmacological inhibition of the hepatocyte bile acid trans- function [28]. The Normal Rat Cholangiocyte line (NRC) devel-
porter Mrp2, which pumps bile acids into the bile canaliculi, oped by LaRusso’s group [27] develops a primary cilium, which
prevented bile duct branching. Bile acids that reach bile ducts is a sign of biliary differentiation [29], and recapitulates the in
from these newly formed bile canaliculi might thus constitute vivo regulation and the cell surface appearance of biliary water
signals for ductal branching into narrower ductules, which then and ion channels [30–32].
connect to the bile canaliculi. Taurocholate signaling for bile In development and during recovery from injury, hepatocytes
canaliculi formation involves activation of the kinase liver and biliary cells evolve from common progenitors. These
kinase B1 (LKB1) and its effector AMP‐regulated kinase include stem cells and transit‐amplifying cells, called hepato-
(AMPK) [11], a signaling cascade that has also been linked to blasts [33]. The liver stem cell reservoir is believed to reside at
4:  Hepatocyte Surface Polarity 39

the canal of Hering (the junction between bile ductules and Crb
hepatocytes) and in peribiliary glands and the gallbladder and is
present throughout life [5, 34]. Hepatoblasts, on the other hand, ls P
are abundant during liver development but in the adult only Pa Par6 Cdc42
detectable in response to liver injury. Both types of progenitors Patj aPKC
have been isolated and expanded in culture in attempts to dif- TJ TJ
ferentiate them into the two epithelial lineages [7, 35–37]. Few P P P Par3
Par1 Lgl
of these studies have convincingly demonstrated hepatocytic or
biliary polarity. In those instances where proper polarization
was achieved, the progenitors gave rise to only one but not both Crb
lineages [38, 39]. Thus, in vitro recapitulation of the branching Cdc42 Lgl

T
Par6
of the two liver epithelial cell types still eludes the field. A pat- Scrib
ented human cancer cell line with hepatoblast characteristics,
Dlg
HepaRG [40], can be polarized with either hepatocyte polarity
when cultured in spheroids [41], or as monolayered hollow P
cysts (i.e. with bile duct polarity) in 3D matrigel culture [42]. Par3 Par1

T
Because of its human origin and relatively high degree of func-
tional differentiation when cultured in spheroids, these cells
have become popular for toxicology studies, although they have
yet to be exploited for polarity studies.
Two technical advances have opened the door to elucidating
Figure 4.2  Polarity complexes organize two distinct membrane
polarity mechanisms even in the intact liver. The first pertains to
domains (apical and basolateral) through self‐recruitment, positive
our ability to express or ablate proteins in hepatocytes by tail feedback, and mutual inhibition. The transmembrane protein Crumbs
vein injection of adenoviruses [43, 44] or of even naked DNA (Crb) engages in lateral homotypic interactions via its extracellular
via hydrodynamic injections [45, 46]. This “poor man’s genet- domain and becomes phosphorylated by aPKC at its cytoplasmic
ics” approach allows acute in vivo manipulation of hepatocytes domain. Both events stabilize the protein, which determines the apical
with relative ease when compared to the generation of knockout domain. Crb interacts with the Pals/Patj complex, which links the apical
and transgenic animals. The second advance is in intravital domain to tight junctions. aPKC, which is activated by scaffolding with
active Cdc42 and Par6, also directs its substrate Par3 to tight junctions
imaging, which now allows live cell imaging of cellular pro- and prevents basolateral surface determinants Lgl and Par1 from attach-
cesses in liver lobes [47]. Such analysis has recently shown that ing to the apical domain. Conversely, Lgl, which genetically interacts
hepatocytes lacking the kinase LKB1 (discussed in more detail with Scribble and Dlg, promotes Crb endocytosis from the basolateral
below) have leaky tight junctions [48]. domain and inhibits the Cdc42/Par6 complex to operate at the lateral
These novel techniques add to the traditional biochemical, membrane. Par1‐mediated phosphorylation of Par3 removes Par3 from
histochemical, and electron microscopic studies of the liver and the lateral domain.
complement dynamic live cell imaging performed in cell lines.
between the apical and lateral complexes and in mammalian
cells is critical for the formation of tight junctions. Available
evidence suggests that the Par, Crumbs, and Scribble complexes
POLARIZATION MECHANISMS are present in both monopolar epithelia and in hepatocytes.
Thus, in the adult liver, immunohistochemical analysis has
determined that aPKC iota and zeta and Par3 are colocalized
Polarity complexes with tight junction markers, specifically ZO‐1, occludin, and
Cell surface polarity is established when signals generated by claudin‐3 at the boundary between bile canaliculi and sinusoidal
local cues or by stochastic fluctuation become amplified through membranes, suggesting that they function in apical junctional
feedback mechanisms to yield a robust segregation of distinct complexes as in other epithelial cells [52]. The basolateral
membrane domains. Work in invertebrates combined with theo- polarity determinants Scribble and Lgl‐2 have been localized to
retical modeling identified a set of core signaling mechanisms the sinusoidal domains in the polarized hepatocytic WIF‐B cells
that could generate cell‐autonomous polarity. They include sev- and Scribble has been shown to mediate similar protein–protein
eral conserved epithelial signaling complexes operating through interactions in the hepatocytic cell line WIF‐B and in monopo-
a combination of positive feedback and mutual antagonistic lar (kidney‐derived) MDCK cells [53].
signaling to stake out distinct surface domains (reviewed in
[49–51] and Figure 4.2). The transmembrane protein Crumbs, Cell matrix adhesion
in conjunction with a Cdc42–aPKC–Par6–Par3 signaling net-
work generates apical polarity in part by clustering and stabiliz- In the context of the tissue, these “polarity complexes” are
ing Crumbs; a Dlg–Lgl–Scribble network promotes basolateral likely positioned by external cues, which include diffusible sig-
surface identity by preventing Crumbs accumulation and by nals and signals originating from cell–ECM and cell–cell adhe-
promoting the accumulation of lateral membrane proteins, sion. Thus, during bile duct formation signals from the portal
including E‐cadherin. The third complex centered around PALs/ mesenchyme mediate the deposition of a basement membrane
Patj and the aPKC–Par3 proteins operates at the intersection between mesenchymal and ductal cells [54, 55]. The basement
40 THE LIVER:  POLARIZATION MECHANISMS

membrane is a two‐dimensional sheet of matrix molecules junctions are not just organizers of cell structure but serve as
secreted by epithelial cells and the underlying connective tis- signaling platforms regulating gene expression, differentiation,
sue. It is organized into sheets primarily by laminin and fibrous proliferation, and morphogenesis. It is this signaling role that
collagen IV networks that are connected with each other via makes them well suited to translate an external cue (i.e. cell–cell
nidogen and the proteoglycan perlecan. It captures growth fac- contact) into a polarization sequence.
tors and cytokines and activates cell surface receptors, mostly Consistent with this idea, hepatocyte luminal domains form at
integrins. ECM–cell signaling triggered by the basement mem- sites of cell–cell contact. Similarly, in the experimental model of
brane in the ductal cells cues apical surface formation on the a single nonpolarized MDCK cell, which first establishes polar-
membrane domain opposite the basement membrane. ity after cell division yields two daughters, the new cell–cell con-
Consistent with a role for laminin in bile duct formation, 3D tact site serves as patch for the establishment of a lumen between
polarization of hepatoblast‐derived biliary cells into a hollow the two cells (Figure 4.3). Lumen formation in this experimental
monolayered cyst in vitro required culture in laminin‐rich system ensues when endocytosed apical proteins, which consti-
matrices and was dependent on the laminin receptor β1‐integrin tutively cycle between endosomes and the entire (nonpolarized)
[38, 56]. Remarkably, hepatocytes differ from other epithelial plasma membrane in single cells, are specifically targeted to and
cells in that they do not assemble a basal lamina [57]. While accumulate at the new cell–cell contact site. In this manner, api-
collagen IV and laminin expression is stimulated during hepa- cal proteins are progressively cleared from the plasma membrane
toblast differentiation along the biliary lineage, expression of outside the cell–cell contact region and are transported to the
these ECM proteins is not stimulated during hepatocyte differ- nascent luminal domain [66, 67]. This constitutes a pathway
entiation [58]. Mature hepatocytes completely lack expression called basolateral‐to‐apical transcytosis. Once MDCK cells have
of the critical basal lamina constituents laminin and nidogen established their first luminal surface, however, new cell–cell
[59]. Whether this unique epithelial feature contributes to the contacts generated after additional cell divisions no longer
unique hepatocyte polarity phenotype has not been directly ­support de novo lumen formation, resulting in the characteristic
tested in vivo but several pieces of circumstantial evidence sup- monopolar organization with a single luminal domain.
port such a scenario: (1) Fibrosis, which results in extensive Hepatocytes differ from monopolar cells in that each cell–cell
matrix deposition between hepatocytes and endothelial cells contact may trigger a new lumen. We hypothesize that two prin-
results in loss of hepatocyte polarity [60, 61]; (2) lack of ECM cipal mechanisms are responsible for this difference: the nature
deposition proved critical for hepatocytic polarity in an experi- of cell–cell adhesion signaling and apical ­protein trafficking.
mental model in which monopolar and hepatocytic polarity can
be switched. This model relies on the kidney‐derived cell line
MDCK, which exhibits monopolar organization but switches to (a) Bas Bas Lat Bas
hepatocytic organization when induced to overexpress the
kinase Par1b [62]. Hepatocytic polarization upon Par1b over- Ap
Bas Ap
expression coincided with reduced ECM deposition and focal
adhesion; importantly, this hepatocytic polarity phenotype was
reversed when the cells were plated on a collagen IV matrix
(b)
[63]. Further investigation revealed reduced RhoA activity Monopolar Multipolar
downstream of the lack of ECM signaling as the critical polar-
ity cue: RhoA depletion was sufficient to induce hepatocytic
polarity in MDCK cells, while pharmacological Rho activation
in the hepatocytic cell line WIF‐B promoted their monopolar
organization [64, 65]. These findings point to RhoA signaling
downstream of cell adhesion signaling as a putative key regula-
tor of the polarity phenotypes.

Cell–cell junctions
In polarized epithelia the contacting membranes form three Figure 4.3  Model for lumen formation in monopolar (ductal cells)
types of intercellular junctions: tight junctions, anchoring junc- versus multipolar (hepatocytes) epithelial cells. (a) Both monopolar and
hepatocytic cells initiate de novo lumen formation at cell–cell adhesion
tions, and gap junctions. Tight junctions provide a barrier for
domains. In monopolar MDCK cell doublets embedded in a 3D matrix
paracellar flow of macromolecules and solutes, thus enforcing the cell–cell contact domain serves as targeting patch for recycling api-
their vectorial transport across epithelial cells; they also restrict cal membrane cargo, which prior to cell–cell contact formation recycles
the diffusion of membrane proteins between the apical and in a nonpolarized manner. When the apical targeting patch is sealed
basolateral domains, thereby maintaining surface polarity. from the surrounding lateral membrane by tight junctions, water trans-
Anchoring junctions, which include adherens junctions and port and the resulting turgor can inflate a lumen. This mechanism is also
­desmosomes, couple cytoskeletal elements to the plasma mem- hypothesized for hepatocytes, which similarly initiate lumina at cell–
cell contact sites. (b) Monopolar and multipolar epithelia differ in the
brane, providing mechanical integrity and allowing mechanical formation of secondary luminal domains: Hepatocytes can form lumina
coupling of cells. Gap junctions mediate cell–cell communica- at each of their cell–cell contact sites. Monopolar epithelia, by contrast,
tion by permitting the passage of small molecular weight solutes do not form additional lumina at new cell–cell contact sites. Instead,
(up to 1 kDa) directly between neighboring cells. Intercellular cell divisions enlarge the existing lumen.
4:  Hepatocyte Surface Polarity 41

As determined in MDCK cells, basolateral‐to‐apical transcy- therefore, that a hepatocytic cell line (HepG2) which failed to
tosis is the mechanism by which apical domains form de novo target E‐cadherin and β‐catenin to the cell surface still established
when new cell–cell junctions become a target site for apical‐ functional tight junctions and bile canaliculi‐like luminal domains
directed cargo. Basolateral‐to‐apical transcytosis is also the (albeit with delayed kinetics), suggesting that E‐cadherin is not
mechanism by which newly synthesized single membrane‐ absolutely essential for the establishment of hepatic polarity [86].
spanning and glycosylphosphatidylinositol (GPI)‐anchored api- Conversely, increasing E‐cadherin levels in this cell line led to the
cal proteins are targeted to the luminal domain in hepatocytes formation of a horizontal tight junction belt characteristic of
[68]. By contrast, MDCK cells and other monopolar epithelia monopolar cells [87]. In MDCK cells, substitution of E‐cadherin
differ from hepatocytes in that they predominantly target newly for an adhesion‐deficient mutant promoted lumen establishment
synthesized apical proteins directly to the apical surface [69]. at cell–cell contact sites as in hepatocytes, at least transiently [88].
Taken together, these two observations suggest that transcytotic Collectively, these data suggest a model by which E‐cadherin‐
apical targeting in hepatocytes ensures that there is an abundant mediated adhesion signaling promotes monopolar organization
supply of apical proteins to create new apical lumina at each and antagonizes hepatocytic polarity. Consistent with this model,
new hepatocyte contact domain. In contrast, direct apical target- immunohistochemistry of liver tissue sections shows significantly
ing to an existing apical domain in monopolar epithelia, com- higher E‐cadherin staining of bile ducts when compared with
bined with the inherent low rate of endocytic recycling at the adjacent hepatocytes [4]. This might be due to lower E‐cadherin
apical surface [70, 71] means apical cargo is unavailable for mRNA levels in hepatocytes, as suggested by recent RNA
endocytic targeting to new cell–cell contact sites. This could sequencing data [89].
prevent formation of new apical domains at cell–cell junctions Even if lumen formation in hepatocyte does not require
in monopolar epithelia such as bile ducts. E‐cadherin adhesion, it likely depends on other cell–cell adhe-
Cell–cell adhesion is initiated by nectins, a family of four sion molecules. This can be concluded from cell culture studies
IgG‐like, Ca2+‐independent adhesion molecules, which activate in which mechanical cell compaction [90] or spheroid forma-
GTP exchange factors for the small GTPases Cdc42, Rac1, and tion [91], both conditions that maximize cell–cell contacts,
Rap1. Signaling by these GTPases then promotes establish- ­significantly stimulated lumen formation.
ment of Ca2+‐dependent adhesion by cadherins, predominantly
E‐cadherin, which in mature adherence junctions clusters in a
zonula adherens adjacent to tight junctions and is linked to a
The taurocholate signaling mechanism
circumferential actin belt [72, 73]. Hepatocytes also express N‐ Analysis of taurocholate signaling, the putative hepatocyte
cadherin [74, 75], which belongs to the same class as E‐cad- polarization trigger, in primary cultures established that tauro-
herin but in other epithelial cells replaces E‐cadherin during cholate increased hepatocyte cAMP levels, which led to activa-
epithelial‐to‐mesenchymal transitions (EMT) [76]. How N‐cadherin‐ tion of EPAC, a GEF for small GTPases of the Rap family. Rap
expressing hepatocytes avoid the fate of EMT is not known. signaling in turn stimulated activity of AMP‐regulated kinase
Junctional adhesion molecules (JAMs), of which JAM‐A is the (AMPK) via its main activating kinase in the liver, LKB1 [11]
best studied, are, together with claudins and occludin, the adhe- (see Chapter 38 for further details).
sion molecules of tight junctions. They initially mingle with LKB1 was discovered as a tumor suppressor mutated in the
nectin and E‐cadherin at cell–cell contacts in immature junc- human Peutz–Jegher syndrome, which is characterized by
tions prior to their segregation into separate adherens and tight hamartomas and polyps in the gastrointestinal tract and by an
junctions [77, 78]. Of these TJ proteins, JAM‐A is absolutely elevated cancer risk. LKB1 activity depends on association with
essential for lumen formation in both hepatocytic (WIF‐B) [79] a pseudokinase and a scaffolding protein [92]. In some tissues,
and monopolar (MDCK) cell lines [80], whereas claudin com- various growth factors activate LKB1, suggesting that hepato-
position appears to contribute to the decision to polarize with cyte polarity and other functions may be regulated by hormones
monopolar or heptocytic polarity. Claudins are the multi‐­ and/or growth factors. Notably, in single intestinal cells LKB1
membrane‐spanning proteins whose extracellular loops are activation induces cell‐autonomous cell surface polarization,
thought to establish the barrier for paracellular flow. The cell hinting that LKB1 might initiate signaling by the self‐organiz-
type‐specific combination in which its 23 members are ing polarity complexes mentioned above [93]. In addition, in
expressed determines the “tightness” of the permeability barrier MDCK cells AMPK was required for tight junction formation
as well as its ion selectivity. Curiously, siRNA‐mediated deple- downstream of cell–cell adhesion [94, 95]. Tight junctional pro-
tion of claudin‐2 from WIF‐B cells [81] and of claudin‐3 from teins appear to be the main AMPK targets in its role as a polarity
the related Can10 cells [82] resulted in a switch from hepato- kinase. They include the substrate Gα‐interacting vesicle‐asso-
cytic to monopolar organization, suggesting specific signaling ciated protein, (aka Girdin), which when phosphorylated by
roles for claudins beyond regulating paracellular flux. AMPK re‐enforces tight junctions under energy stress [96], and
E‐cadherin is considered a key trigger of epithelial polarization cingulin [97], the suggested AMPK substrate responsible for
in monopolar cells [83]. This was established in so‐called leaky tight junctions observed by intravital imaging in LKB1
“Ca‐switch assays,” which exploit the Ca‐dependence of E‐cadherin knockout livers in vivo [48]. In addition to tight junction defects,
adhesion. In Ca‐free medium monopolar epithelial cells such as LKB1 knockout livers also presented with defective targeting of
MDCK cells do not polarize even when plated at confluence bile acid transporters to the luminal surface, which were cor-
(i.e. in the presence of only nectin‐mediated cell–cell adhesion). rected by cAMP activation [25]. As can be expected from the
Re‐addition of Ca triggers rapid, synchronized tight junction polarity defects, LKB1 conditional liver knockout mice suffered
establishment and lumen formation [84, 85]. It was unexpected, cholestasis and liver injury [98].
42 THE LIVER:  POLARIZED PROTEIN TRAFFICKING

POLARIZATION REQUIRES ENERGY prerequisite for heptocytic polarization in cell culture models.
Thus, a RhoA–HNF4a axis might link hepatocyte polarity to
Hepatocyte polarization and all of its components are energy‐ functionality.
dependent and have been increasingly related to two protein
kinase, LKB1 and AMPK (see Chapter 38 for further details).
AMPK and LKB1 contribute to hepatocyte polarization by con-
trolling ATP synthesis and energy metabolism. AMPK, a serine POLARIZED PROTEIN TRAFFICKING
threonine kinase containing a catalytic α subunit and regulatory
β and γ subunits, controls energy metabolism within cells by The vectorial activities of epithelia require that each transporter/
sensing the cellular AMP to ATP ratio. Activation of AMPK by channel has a clearly defined apical–basal polarity, which
phosphorylation of the α subunit Thr172 decreases energy con- results from sorting events in the biosynthetic and recycling
sumption and increases energy production during cellular stress, routes at the trans‐Golgi network (TGN), common recycling
such as hypoxia, glucose deprivation, and ischemia, and has an endosomes (CRE), and apical recycling endosomes (ARE)
important role in hepatic metabolism through effects on glu- [107] (Figures 4.4 and 4.5). Biosynthetic protein trafficking itin-
cose, lipid and protein homeostasis and mitochondrial biogene- eraries in hepatocytes have been established by combining an in
sis. Long‐term effects involve regulation of the glycolytic and vivo pulse‐chase protocol with cell fractionation [108]. This
lipogenic pathways. protocol, which exploits the observation that 35S‐methionine,
In collagen sandwich cultured hepatocytes, the stress of iso- when injected into the tail vein first reaches the liver and is
lation resulted in depolarization, ATP depletion, and mitochon- mostly incorporated into newly synthesized hepatocyte pro-
drial fragmentation [99]. Mitochondrial fusion occurred within teins, is in fact the only in vivo approach to protein trafficking
2 days associated with increased ATP synthesis from oxidative for any mammalian epithelium. It reveals that hepatocytes target
phosphorylation and canalicular network formation. Subsequent polytopic membrane proteins, such as ABC transporters, from
AMPK activation upregulated glucose uptake, glycolysis, and a the TGN directly to the bile canalicular domain, some of which
further increase in ATP. These studies reveal that, after stress, first accumulate at a Rab11a‐positive ARE pool, from which
hepatocytes preferentially restore polarity even at low ATP lev- they then cycle to the canalicular membrane [109]. In contrast,
els, suggesting that polarity is a prime requirement for cellular all single membrane‐spanning and GPI‐anchored bile canalicu-
activity [100]. lar membrane proteins are targeted from the TGN to the basolat-
eral plasma membrane, from where they transcytose in
endosomes through the cell to the canalicular membrane [108,
110, 111]. This is different from all monopolar epithelia studied
to date, which utilize the basolateral‐to‐apical transcytotic route
HEPATOCYTE POLARITY IS LINKED TO for apical targeting to a lesser extent than hepatocytes. The most
THE EXPRESSION OF DIFFERENTIATION poignant difference is that unlike monopolar epithelia, hepato-
MARKERS cytes lack a pathway that targets soluble proteins from the TGN
to the bile canalicular domain via vesicular transport. All solu-
When hepatocytes are cultured on rigid 2D matrices they rap- ble cargo that passes through the secretory pathway is (predomi-
idly lose expression of differentiation markers such as albumin nantly) directed into the space of Disse [112, 113].
or various transporters. The degree of de‐differentiation is pro- The mechanisms that segregate apical and basolateral pro-
portional to the ECM concentration regardless of its nature. teins and mediate their surface domain‐specific targeting from
Increasing matrix concentrations also cause increasing cell the TGN have been largely elucidated in the epithelial model
spreading which is incompatible with hepatocyte polarization cell line MDCK. Work over three decades by many groups
[101–103]. Conversely, hepatocytes cultured on soft matrices, using apical and basolateral model proteins yielded the follow-
particularly in sandwich configuration, maintain expression of ing model of polarized protein delivery in the direct biosynthetic
differentiation markers; they also acquire polarity [21, 22]. pathway (Figure 4.4): Sorting machinery at the TGN recognizes
This correlation between increased expression of markers for apical and basolateral sorting signals and packages apical and
hepatocyte function and better polarization was also observed basolateral proteins into separate TGN‐derived transport carri-
when hepatocytes or hepatic cell lines such as HepaRG were ers [107]. Sorting signals in basolateral proteins have affinities
cultured in the absence of exogenous matrix but with extensive for clathrin adaptors, which mediate their clathrin‐dependent
cell–cell contact in spheroids [91]. Its crux is cytoskeletal basolateral surface targeting; in the absence of clathrin basolat-
­tension, which activates a mechano‐transduction cascade in eral proteins still reach the plasma membrane, but they are mis‐
which integrin signaling leads to FAK→RhoA/RhoK→ERK1/2 targeted to the apical domain [114]. Based on their interaction
activation, which promotes proliferation and induces an EMT with different adaptors, basolateral cargoes take at least two dif-
[104, 105]. Importantly, activation of RhoA/RhoK in this path- ferent routes out of the TGN: If not captured by TGN‐localized
way inhibits expression of the hepatocytic master transcription clathrin adaptors such as AP1A, they are targeted to transfer-
factor HNF4a, which controls a hepatocyte‐specific transcrip- rin‐positive recycling endosomes where basolateral proteins
tional network that includes both functional proteins and polar- interact with the endosomal clathrin adaptor AP1B for basolat-
ity determinants [106]. Forced HNF4a expression on rigid eral surface delivery. AP1A‐interacting cargo, by contrast, is
matrix remedied loss of several, although not all, functional delivered to the basolateral surface without traversing recycling
markers. As previously discussed, low RhoA activity is also a compartments. While some basolateral proteins are known to
4:  Hepatocyte Surface Polarity 43

AP pathways
(a) (b)
BL pathways
AP

MyoVb
Rab11a
AEE
ARE

AP1B (i)
MyoVb
CRE Rab11a

BC
BC ARE
TGN
AP1A
AP4
TGN
(ii)
trans-

TGN

BL BL

Figure 4.4  Targeting pathways from the TGN to the apical/canalicular and basolateral surfaces in MDCK and hepatocytes. (a) Known pathways
in MDCK cells. Apical (AP) pathways (green arrows): depending on their presence in detergent‐insoluble microdomains, raft‐dependent and raft‐
independent apical routes have been distinguished. Several membrane and secretory proteins have been shown to traverse the apical recycling
endosome (ARE) (in a MyoVb and Rab11‐dependent manner) or the apical early endosome (AEE) before reaching the apical domain. Basolateral
(BL) pathways (blue arrows): basolateral protein exiting from the TGN is clathrin mediated. Multiple pathways have been distinguished. Some
basolateral proteins reach the basolateral domain without intermediate (dependent on TGN‐localized clathrin adaptors AP1A and AP4), others
traverse the common recycling endosome (CRE), from where their basolateral targeting is mediated by clathrin adaptor AP1B. (b) Known and
hypothetical pathways in hepatic cells. Two classes of apical proteins have been distinguished: polytopic membrane proteins, such as bile acid
transporters, travel from the TGN to the apical domain directly (at least some via a Rab11 compartment); GPI‐anchored and single membrane‐span-
ning membrane proteins first reach the basolateral domain from the TGN. It is not resolved whether (i) apical and basolateral proteins are sorted
into distinct TGN‐derived transport carriers that both reach the basolateral domain, or (ii) whether apical and basolateral proteins travel in common
carriers to the basolateral domain. Since hepatocytes lack AP1B, it is unlikely that hepatocytes target their biosynthetic cargo via the CRE. Adapted
from Treyer and Müsch, Compr Physiol, 2013;3:243–87 with permission of John Wiley & Sons.

enter both pathways, others are predominantly targeted via one account for heptocyte TGN‐to‐basolateral targeting of apical
or the other pathway [115, 116]. Apical sorting signals, on the proteins that in MDCK cells are targeted to the apical surface
other hand, are pleomorphic, constituted by N‐glycans, GPI directly. So far, only one contributing mechanism has been
anchors, and specialized transmembrane protein domains with clearly identified: the lack in hepatocytes of myelin and lym-
affinity for lipid rafts. Lipid rafts are formed by sphingolipids phocyte 1 (MAL1), a proteolipid tetraspanning raft‐associated
and cholesterol, which cluster apical cargo and also establish membrane protein, which in MDCK cells is critical for TGN‐
membrane domain boundaries that segregate apical and basolat- derived apical transport of proteins that depend on association
eral cargo domains [117]. Apical proteins are targeted from the with lipid rafts for their polarized targeting [128, 129].
TGN in microtubule (MT)‐dependent tubular carriers, indepen- Expression of MAL1 in the hepatocytic line WIF‐B (which
dently of clathrin [69]. Apical and basolateral transport carriers ­normally lack it) promoted vectorial delivery of GPI‐apical pro-
also differ in their fission and fusion machineries. Thus, while teins and single‐pass transmembrane proteins from the TGN to
basolateral carrier scission depends on the activities of protein the apical surface [130]. It was also suggested (for the HepG2
kinase D and on CtBP1‐S/BARS‐induced activation of cell line) that the transcytotic mechanism is regulated by the
lysophosphatidic acid acyltransferase δ [118–120], apical carri- levels of surface E‐cadherin [86], although no mechanism for
ers require the scission factor dynamin 2, which functions in such regulation has been proposed.
conjunction with branched actin filaments that assemble around Hepatocytes also lack the basolateral cargo adaptor AP1B. In
the apical carriers at the scission site [121, 122]. Basolateral all other epithelia that naturally fail to express AP1B, AP1B‐
fusion is mediated by SNARE pairings containing VAMP3 and dependent basolateral proteins, such as transferrin receptor,
syntaxin 4, while apical carriers fuse via tetanus‐resistant v‐ localize at the apical domain [116]. This is because their endo-
SNAREs (VAMP7,8) that pair with syntaxin 3 [44, 123–127]. somal sorting in both the exocytic and the recycling pathways
Hepatocytes can be expected to utilize some of the same tar- are compromised. However, such polarity reversal is not
geting mechanisms, yet there also have to be differences to observed in hepatocytes or hepatocytic cell lines. This implies
44 THE LIVER:  CYTOSKELETAL MICROFILAMENT AND MICROTUBULAR SYSTEMS IN TRAFFICKING

epithelial cells accumulate subapical membranes that appear to


BL be bottlenecks for protein delivery to the apical surface. They
are enriched in Rab GTPases, including Rab11a and its effector,
the actin‐associated molecular motor myosin Vb, as well as the
LYS Rab11 adaptor proteins Fip1 and Fip2 [140–143]. Inhibition of
AEE Rab11a or myosin Vb indeed prevented polarization of WIF‐B
cells and primary hepatocytes [144] and, when introduced into
MyoVb
Rab11a polarized cells, prompted depolarization and internalization of
BC apical proteins [145]. These observations highlight the impor-
ARE SAC tance of apical membrane traffic for the establishment and
? maintenance of polarity.
The Rab11‐positive endosomes also serve as “holding cells”
for hepatocyte ABC transporters, the majority of which are
LYS BEE BEE
maintained intracellularly rather than at the bile canalicular
membrane [25, 146]. This intracellular pool is mobilized by bile
acids circulating in the enterohepatic circuit, and by postprandi-
ally secreted peptide hormones, which increase cAMP produc-
tion in hepatocytes to cope with the increased demand for bile
acid secretion. Taurocholate and cAMP activate distinct signal-
BL proteins AP proteins ing pathways to mobilize ABC transporters to the canalicular
Figure 4.5  Targeting pathways from the apical and basolateral plasma domain. Regulation of these two responses differs. Increase in
membranes in hepatocytes. Pathways for apical (AP) proteins (green cAMP concentration results in PKA‐mediated stimulation of
arrows): Upon endocytosis from the canalicular domain proteins are phosphoinositide 3‐kinase but not taurocholate‐stimulated incor-
sorted in the apical early endosome (AEE) into either the late endoso- poration of ABCB11 into the canalicular membrane [147–150].
mal/lysosomal pathway or for recycling to the apical recycling endo- Bulk endocytosis from the apical surface of polarized epithe-
some (ARE). Apical proteins that have reached the basolateral surface lia occurs at much lower rate than endocytosis from the basolat-
from the TGN are internalized via clathrin‐coated vesicles (non‐raft) or
in a clathrin‐independent, flotillin‐dependent manner (raft) and reach
eral domain [70]. This is likely due to the extensive direct and
the subapical endosome (SAC) via basolateral early endosomes (BEEs). indirect linkage of apical membrane proteins to actin filaments,
From the SAC they are targeted to the apical surface via the ARE. which makes it hard for the endocytic machinery to deform the
Basolateral (BL) resident proteins (blue arrows) undergo fast recycling membrane. On the other hand, a branched actin network at the
from the BEE directly or via recycling endosomes. Whether they min- base of a nascent endocytic vesicle promotes clathrin‐mediated
gle with apical‐directed cargo in a common recycling endosome (CRE) endocytosis when linked to the scission factor dynamin via the
as described for MDCK cells is still debated. While the SAC has been
actin and dynamin‐binding protein cortactin [151, 152]. Apical
described by some as the equivalent of the MDCK CRE, others have
suggested that BL and AP proteins are segregated before apical‐directed bile acid transporters BSEP, MDR1, and MDR2 contain in their
cargo reaches the SAC. Similar to the AEE, the BEE also sorts proteins cytoplasmic domains a binding site for HCLS1‐associated pro-
for degradation to the lysosome (LYS). Adapted from Treyer and tein X‐1 (HAX‐1) [153], which also binds cortactin and hence
Müsch, Compr Physiol, 2013;3:243–87 with permission of John could recruit these ABC transporters into productive endocytic
Wiley & Sons. pits. Consistent with this hypothesis, HAX‐1 depletion increased
BSEP accumulation at the apical domain [153]. Together, their
that hepatocytes organize their endosomal recycling itineraries facilitated endocytosis and regulated re‐insertion from an endo-
differently from MDCK cells (Figure 4.5). In the MDCK model, cytic pool allows for the highly dynamic behavior of ABC trans-
endocytosed cargo can either undergo fast recycling to their porters at the bile canalicular domain.
membrane of origin from early endosomes or, alternatively,
recycle from a CRE, where cargo from both surfaces mix and is
sorted into distinct apical‐ or basolateral‐directed recycling car-
riers [131–133]. The CRE is also where basolateral‐to‐apical
CYTOSKELETAL MICROFILAMENT
transyctotic cargo is segregated from basolateral recycling cargo AND MICROTUBULAR SYSTEMS IN
[132, 134]. Apical carriers exiting the CRE either mature into or TRAFFICKING(SEE CHAPTER 3)
fuse with the AREs beneath the apical surface, while basolateral
cargo returns to the surface directly [135, 136]. Studies in the Proper endosomal trafficking and recycling of proteins to all
hepatocytic cell lines WIF‐B and HepG2 as well as electron plasma membrane domains requires an intact actin and microtu-
microscopy 3D reconstruction of the hepatocyte endosomal sys- bular cytoskeletal system. In particular, plus‐end dynamic
tem have indeed suggested differences in endosome organiza- microtubules marked by CLIP170 and EB1 mediate trafficking
tion compared to MDCK cells. In particular, a subapical of secreted and canalicular proteins. Newly synthesized
compartment (SAC) has been described, which according to dif- ABCB11, the canalicular bile acid transporter, and other canali-
ferent schools either mediates apical and basolateral cargo sort- cular ABC transporters traffic in post Golgi vesicles from the
ing equivalent to the MDCK CRE or represents a unique TGN along microtubules. However, dynamic microtubules do
compartment of the basolateral‐to‐apical transcytotic pathway not attach to the canalicular membrane and their cargo
[16, 111, 137–139]. Regardless of the model and cell type, all endosomes are transferred to the pericanalicular actin system
4:  Hepatocyte Surface Polarity 45

(Figure 4.5). The complex mechanism for cargo transfer is not (a) Monopolar Hepatocytic
known; however, microtubules become associated with actin
through a pericanalicular actin‐binding complex containing IQ
Gap, APC, Hax‐1, and cortactin proteins. Live cell imaging
studies reveal that selective plasma membrane localization of
transporter proteins is predominantly due to the localization of
specific docking proteins. In polarized WIF‐B cells, ABCB11
and ABCB1 traffic along microtubules throughout the cell but
only attach to specific sites on the canalicular membrane. The
docking site has been proposed to be syntaxin 3 that facilitates
fusion of protein‐sorting vesicles with the inner leaflet of the
(b)
canalicular membrane from which they diffuse until reaching
the tight junction which limits presence to the canalicular
domain. The actin‐binding protein radixin links some cargo
molecules, such as ABCC2 and ABCB11, to the pericanalicular
actin system. Radixin knockout mice manifest impaired ABCC2
and ABCB11 localization to the canalicular domain which
becomes progressively devoid of microvilli, resulting in hepato-
cyte injury.
Formin controls the assembly and disassembly of short actin Figure 4.6  Metaphase spindle orientation (a) and division outcome
filaments which are involved in endosomal transport. In the (b) in cultured monopolar and heptocytic cells. Monopolar epithelia
orient their metaphase spindle parallel to their basal surface when
HepG2 hepatoma cell line INF2, CDC42 and the transmem- astral microtubules bind to cortical cues (green) below the apical sur-
brane protein MAL2 are required for transcytotic trafficking of face (red) (a). The cleavage furrow, which is established perpendicular
canalicular membrane proteins. to the spindle pole axis, bisects the apical surface and yields two iden-
tical daughters, each attached to the substratum (b). In hepatocytes the
spindle attaches adjacent to two different lumina. The cleavage furrow
therefore does not bisect the luminal domains. In cultured WIF‐B and
HEPATOCYTE POLARITY AND CELL HepG2 cells, a stereotypic spindle tilt with respect to the substrate con-
tacting basal domain causes one daughter cells to be removed from the
DIVISION substratum. Whether this is relevant in vivo remains to be determined.
Adapted from [64].
Although mature polarized hepatocytes are quiescent, they can,
under conditions of severe injury, re‐enter the cell cycle and pro- adjacent to their luminal surfaces [64, 161] (Figure 4.6). Lumen
liferate [154]. Hepatocytes also actively divide during postnatal geometry and the presence of at least two luminal domains per
development while their bile canaliclar network matures. How hepatocyte likely cause each astral microtubule set to attach
proliferating epithelial cells orient their mitotic spindle and next to a different lumen, thereby preventing bisection of either
the cleavage furrow, which always forms perpendicular to the lumen. By contrast, in monopolar cells, the two astral microtu-
­spindle pole axis, is of critical importance for both tissue and bule attachment sites flank the same (single) luminal domain,
cellular organization [155]. Monopolar epithelial cells orient resulting in its bisection. In an additional difference to monopo-
their metaphase spindle parallel to the basement membrane lar cells, hepatocytic cell lines can accommodate the metaphase
[156, 157]. This ensures that the cleavage furrow bisects their spindle only in a quasi‐diagonal position, resulting in a stereo-
luminal domain, yielding two identical daughters that both typic spindle tilt with respect to their basal domains, which
remain in the epithelial plane, thereby ensuring that the epithe- yields a division out of the monolayer and results in the bilayer-
lium remains monolayered. Critical to this outcome are cortical ing observed in HepG2 and WIF‐B cell cultures [65]. Live cell
cues positioned at equal distance from the basal surface at oppo- imaging of mammalian liver tissue [48, 162] should make it
site lateral domains. The cues consist of an evolutionarily con- feasible to test the relevance of this division mode in vivo.
served protein complex in which the α subunit of a trimeric
G‐protein provides the cortical anchor and the minus end‐
directed microtubule motor dynein binds the plus ends of astral
microtubules. In metaphase these complexes on opposite mem- LIVER DISEASE AND POLARITY
brane domains capture one of each set of astral spindle microtu-
bules, thereby aligning the metaphase spindle parallel to the As detailed in Chapter 29, mutations in MYO5B encoding myo-
basal domain [158, 159]. Unlike monopolar cells, hepatocytes sin Vb and Rab11b, Rab25, and Rab8, cause microvillus inclu-
rarely bisect their luminal domain in cytokinesis [160]. This sion disease (MVID), in which malabsorption results from the
preserves canalicular lumen organization and prevents the gen- absence of the intestinal brush border. Other patients manifest
eration of acini. Remarkably, molecular analysis in the WIF‐B cholestasis and progressive liver disease. Mouse Rab8 condi-
and HepG2 culture models suggested that the different cell divi- tional knockouts mimic MVID. Mutations affecting syntaxin 3,
sion outcomes in monopolar and hepatocytic cells occur even an apical membrane SNARE (family of membrane proteins that
though both cell types use similar spindle orientation mecha- ensures fusion between opposing membranes), suggesting that
nism, that is, both place their cortical spindle captures cues myosin 5b, Rab8, and syntaxin 3 may be involved in the same
46 THE LIVER: REFERENCES

trafficking pathway. Discovery of loss‐of‐function mutations in 22. Dunn, J.C. et  al. Hepatocyte function and extracellular matrix geometry:
genes encoding recycling endosome‐associated proteins such as long‐term culture in a sandwich configuration. FASEB J, 1989;3(2):174–7.
23. LeCluyse, E.L., Audus, K.L., and Hochman, J.H. Formation of extensive
myosin 5b in MVID, VPS33B and VIPAR in arthrogryposis, canalicular networks by rat hepatocytes cultured in collagen‐sandwich con-
renal dysfunction, and cholestasis syndrome (ARC), also sup- figuration. Am J Physiol, 1994;266(6 Pt 1):C1764–74.
ports the importance of the RE in establishment and mainte- 24. Michalopoulos, G.K. et al. Comparative analysis of mitogenic and morpho-
nance of hepatocyte polarity. genic effects of HGF and EGF on rat and human hepatocytes maintained in
collagen gels. J Cell Physiol, 1993;156(3):443–52.
25. Homolya, L. et al. LKB1/AMPK and PKA control ABCB11 trafficking and
polarization in hepatocytes. PLoS One, 2014;9(3):e91921.
26. Grubman, S.A. et al. Regulation of intracellular pH by immortalized human
REFERENCES intrahepatic biliary epithelial cell lines. Am J Physiol, 1994;266(6 Pt
1):G1060–70.
1. Arias, I.M. et al. The Liver: Biology and Pathobiology. 5th edn (eds. I.M. 27. Vroman, B. and LaRusso, N.F. Development and characterization of polar-
Arias et al.), Wiley‐Blackwell, Chichester, 2009. ized primary cultures of rat intrahepatic bile duct epithelial cells. Lab Invest,
2. Si‐Tayeb, K., Lemaigre, F.P., and Duncan, S.A. Organogenesis and develop- 1996;74(1):303–13.
ment of the liver. Dev Cell, 2010;18(2):175–89. 28. Salter, K.D. et al. Modified culture conditions enhance expression of differ-
3. Gordillo, M., Evans, T., and Gouon‐Evans, V. Orchestrating liver develop- entiated phenotypic properties of normal rat cholangiocytes. Lab Invest,
ment. Development, 2015;142(12):2094–108. 2000;80(11):1775–8.
4. Antoniou, A. et al. Intrahepatic bile ducts develop according to a new mode 29. Gradilone, S.A. et al. Cholangiocyte cilia express TRPV4 and detect changes
of tubulogenesis regulated by the transcription factor SOX9. in luminal tonicity inducing bicarbonate secretion. Proc Natl Acad Sci U S A,
Gastroenterology, 2009;136(7):2325–33. 2007;104(48):19138–43.
5. Zhang, L. et  al. The stem cell niche of human livers: symmetry between 30. Lazaridis, K.N. et  al. Kinetic and molecular identification of sodium‐
development and regeneration. Hepatology, 2008;48(5):1598–607. dependent glucose transporter in normal rat cholangiocytes. Am J Physiol,
6. Carpentier, R. et al. Embryonic ductal plate cells give rise to cholangiocytes, 1997;272(5 Pt 1):G1168–74.
periportal hepatocytes, and adult liver progenitor cells. Gastroenterology, 31. Spirli, C. et al. Functional polarity of Na+/H+ and Cl−/HCO3− exchangers in a
2011;141(4):1432–8, 1438.e1–4. rat cholangiocyte cell line. Am J Physiol, 1998;275(6 Pt 1):G1236–45.
7. Wauthier, E. et al. Hepatic stem cells and hepatoblasts: identification, isola- 32. Zsembery, A. et al. Purinergic regulation of acid/base transport in human and
tion, and ex vivo maintenance. Methods Cell Biol, 2008;86:137–225. rat biliary epithelial cell lines. Hepatology, 1998;28(4):914–20.
8. Gallin, W.J. Development and maintenance of bile canaliculi in vitro and in 33. Miyajima, A., Tanaka, M., and Itoh, T. Stem/progenitor cells in liver devel-
vivo. Microsc Res Tech, 1997;39(5):406–12. opment, homeostasis, regeneration, and reprogramming. Cell Stem Cell,
9. Kanamura, S., Kanai, K., and Watanabe, J. Fine structure and function of hepat- 2014;14(5):561–74.
ocytes during development. J Electron Microsc Tech, 1990;14(2):92–105. 34. Reid, L.M. Stem/progenitor cells and reprogramming (plasticity) mecha-
10. Little, J.M. et al. Taurocholate pool size and distribution in the fetal rat. J nisms in liver, biliary tree, and pancreas. Hepatology, 2016;64(1):4–7.
Clin Invest, 1979;63(5):1042–9. 35. Tanimizu, N. et al. Long‐term culture of hepatic progenitors derived from
11. Fu, D. et al. Bile acid stimulates hepatocyte polarization through a cAMP‐ mouse Dlk+ hepatoblasts. J Cell Sci, 2004;117(Pt 26):6425–34.
Epac‐MEK‐LKB1‐AMPK pathway. Proc Natl Acad Sci U S A, 36. Wang, Y. et al. Paracrine signals from mesenchymal cell populations govern
2011;108(4):1403–8. the expansion and differentiation of human hepatic stem cells to adult liver
12. Tanimizu, N. et al. Intrahepatic bile ducts are developed through formation fates. Hepatology, 2010;52(4):1443–54.
of homogeneous continuous luminal network and its dynamic rearrangement 37. Huch, M. et al. Long‐term culture of genome‐stable bipotent stem cells from
in mice. Hepatology, 2016;64(1):175–88. adult human liver. Cell, 2015;160(1–2):299–312.
13. Lo, B. et  al. Lkb1 regulates organogenesis and early oncogenesis 38. Tanimizu, N., Miyajima, A., and Mostov, K.E. Liver progenitor cells develop
along  AMPK‐dependent and ‐independent pathways. J Cell Biol, cholangiocyte‐type epithelial polarity in three‐dimensional culture. Mol Biol
2012;199(7):1117–30. Cell, 2007;18(4):1472–9.
14. Darlington, G.J., Kelly, J.H., and Buffone, G.J. Growth and hepatospecific 39. Tanimizu, N. et al. Hepatic biliary epithelial cells acquire epithelial integrity
gene expression of human hepatoma cells in a defined medium. In Vitro Cell but lose plasticity to differentiate into hepatocytes in vitro during develop-
Dev Biol, 1987;23(5):349–54. ment. J Cell Sci, 2013;126(Pt 22):5239–46.
15. Decaens, C. et al. Establishment of hepatic cell polarity in the rat hepatoma‐ 40. Guillouzo, A. et al. The human hepatoma HepaRG cells: a highly differenti-
human fibroblast hybrid WIF‐B9. A biphasic phenomenon going from a sim- ated model for studies of liver metabolism and toxicity of xenobiotics. Chem
ple epithelial polarized phenotype to an hepatic polarized one. J Cell Sci, Biol Interact, 2007;168(1):66–73.
1996;109(Pt 6):1623–35. 41. Gunness, P. et al. 3D organotypic cultures of human HepaRG cells: a tool for
16. Ihrke, G. et  al. Apical plasma membrane proteins and endolyn‐78 travel in vitro toxicity studies. Toxicol Sci, 2013;133(1):67–78.
through a subapical compartment in polarized WIF‐B hepatocytes. J Cell 42. Dianat, N. et  al. Generation of functional cholangiocyte‐like cells from
Biol, 1998;141(1):115–33. human pluripotent stem cells and HepaRG cells. Hepatology,
17. Kelly, J.H. and Darlington, G.J. Modulation of the liver specific phenotype 2014;60(2):700–14.
in the human hepatoblastoma line Hep G2. In Vitro Cell Dev Biol, 43. Nakatani, T. et al. Assessment of efficiency and safety of adenovirus medi-
1989;25(2):217–22. ated gene transfer into normal and damaged murine livers. Gut,
18. Peng, X. et al. How to induce non‐polarized cells of hepatic origin to express 2000;47(4):563–70.
typical hepatocyte polarity: generation of new highly polarized cell models 44. Shayakhmetov, D.M. et al. Analysis of adenovirus sequestration in the liver,
with developed and functional bile canaliculi. Cell Tissue Res, transduction of hepatic cells, and innate toxicity after injection of fiber‐mod-
2006;323(2):233–43. ified vectors. J Virol, 2004;78(10):5368–81.
19. van der Wouden, J.M., van IJzendoorn, S.C., and Hoekstra, D. Oncostatin M 45. Zhang, G., Budker, V., and Wolff, J.A. High levels of foreign gene expression
regulates membrane traffic and stimulates bile canalicular membrane bio- in hepatocytes after tail vein injections of naked plasmid DNA. Hum Gene
genesis in HepG2 cells. EMBO J, 2002;21(23):6409–18. Ther, 1999;10(10):1735–7.
20. Wakabayashi, Y., Lippincott‐Schwartz, J., and Arias, I.M. Intracellular traf- 46. Eggenhofer, E. et al. High volume naked DNA tail‐vein injection restores
ficking of bile salt export pump (ABCB11) in polarized hepatic cells: consti- liver function in Fah‐knock out mice. J Gastroenterol Hepatol,
tutive cycling between the canalicular membrane and rab11‐positive 2010;25(5):1002–8.
endosomes. Mol Biol Cell, 2004;15(7):3485–96. 47. Marques, P.E. et al. Imaging liver biology in vivo using conventional confo-
21. Dunn, J.C., Tompkins, R.G., and Yarmush, M.L. Long‐term in vitro function cal microscopy. Nat Protoc, 2015;10(2):258–68.
of adult hepatocytes in a collagen sandwich configuration. Biotechnol Prog, 48. Porat‐Shliom, N. et al. Liver kinase B1 regulates hepatocellular tight junc-
1991;7(3):237–45. tion distribution and function in vivo. Hepatology, 2016;64(4):1317–29.
4:  Hepatocyte Surface Polarity 47

49. Tepass, U. The apical polarity protein network in Drosophila epithelial cells: 78. Asakura, T. et al. Similar and differential behaviour between the nectin‐afa-
regulation of polarity, junctions, morphogenesis, cell growth, and survival. din‐ponsin and cadherin‐catenin systems during the formation and disrup-
Annu Rev Cell Dev Biol, 2012;28:655–85. tion of the polarized junctional alignment in epithelial cells. Genes Cells,
50. Thompson, B.J. Cell polarity: models and mechanisms from yeast, worms 1999;4(10):573–81.
and flies. Development, 2013;140(1):13–21. 79. Braiterman, L.T. et al. JAM‐A is both essential and inhibitory to develop-
51. Rodriguez‐Boulan, E. and Macara, I.G. Organization and execution of the ment of hepatic polarity in WIF‐B cells. Am J Physiol Gastrointest Liver
epithelial polarity programme. Nat Rev Mol Cell Biol, 2014;15(4):225–42. Physiol, 2008;294(2):G576–88.
52. Takaki, Y. et al. Dynamic changes in protein components of the tight junction 80. Rehder, D. et al. Junctional adhesion molecule‐a participates in the forma-
during liver regeneration. Cell Tissue Res, 2001;305(3):399–409. tion of apico‐basal polarity through different domains. Exp Cell Res,
53. Kallay, L.M. et  al. Scribble associates with two polarity proteins, Lg12 2006;312(17):3389–403.
and  Vang12, via distinct molecular domains. J Cell Biochem, 81. Son, S. et al. Knockdown of tight junction protein claudin‐2 prevents bile
2006;99(2):647–64. canalicular formation in WIF‐B9 cells. Histochem Cell Biol,
54. Terada, T. and Nakanuma, Y. Expression of tenascin, type IV collagen and 2009;131(3):411–24.
laminin during human intrahepatic bile duct development and in intrahepatic 82. Grosse, B. et al. Build them up and break them down: tight junctions of cell
cholangiocarcinoma. Histopathology, 1994;25(2):143–50. lines expressing typical hepatocyte polarity with a varied repertoire of clau-
55. Shiojiri, N. and Sugiyama, Y. Immunolocalization of extracellular matrix dins. Tissue Barriers, 2013;1(4):e25210.
components and integrins during mouse liver development. Hepatology, 83. Takeichi, M. Cadherin cell adhesion receptors as a morphogenetic regula-
2004;40(2):346–55. tor. Science, 1991;251(5000):1451–5.
56. Tanimizu, N. et  al. alpha1‐ and alpha5‐containing laminins regulate the 84. Gonzalez‐Mariscal, L., Chavez de Ramirez, B., and Cereijido, M. Tight
development of bile ducts via beta1 integrin signals. J Biol Chem, junction formation in cultured epithelial cells (MDCK). J Membr Biol,
2012;287(34):28586–97. 1985;86(2):113–25.
57. Schaffner, F. and Poper, H. Capillarization of hepatic sinusoids in man. 85. Vega‐Salas, D.E., Salas, P.J. and Rodriguez‐Boulan, E. Exocytosis of vacu-
Gastroenterology, 1963;44:239–42. olar apical compartment (VAC): a cell‐cell contact controlled mechanism
58. Yang, L. et al. A single‐cell transcriptomic analysis reveals precise pathways for the establishment of the apical plasma membrane domain in epithelial
and regulatory mechanisms underlying hepatoblast differentiation. cells. J Cell Biol, 1988;107(5):1717–28.
Hepatology, 2017;66(5):1387–401. 86. Theard, D. et al. Cell polarity development and protein trafficking in hepat-
59. Martinez‐Hernandez, A. and Amenta, P.S. The hepatic extracellular matrix. ocytes lacking E‐cadherin/beta‐catenin‐based adherens junctions. Mol Biol
I. Components and distribution in normal liver. Virchows Arch A Pathol Anat Cell, 2007;18(6):2313–21.
Histopathol, 1993;423(1):1–11. 87. Konopka, G. et al. Junctional adhesion molecule‐A is critical for the forma-
60. Bataller, R. and Brenner, D.A. Liver fibrosis. J Clin Invest, tion of pseudocanaliculi and modulates E‐cadherin expression in hepatic
2005;115(2):209–18. cells. J Biol Chem, 2007;282(38):28137–48.
61. Wells, R.G. The role of matrix stiffness in regulating cell behavior. 88. Cohen, D., Tian, Y., and Musch, A. Par1b promotes hepatic‐type lumen
Hepatology, 2008;47(4):1394–400. polarity in Madin Darby canine kidney cells via myosin II‐ and E‐cadherin‐
62. Cohen, D. et al. Mammalian PAR‐1 determines epithelial lumen polarity by dependent signaling. Mol Biol Cell, 2007;18(6):2203–15.
organizing the microtubule cytoskeleton. J Cell Biol, 2004;164(5):717–27. 89. Su, X. et al. Single‐cell RNA‐Seq analysis reveals dynamic trajectories dur-
63. Cohen, D. et al. The serine/threonine kinase Par1b regulates epithelial lumen ing mouse liver development. BMC Genomics, 2017;18(1):946.
polarity via IRSp53‐mediated cell‐ECM signaling. J Cell Biol, 90. Wang, Y. et al. Mechanical compaction directly modulates the dynamics of
2011;192(3):525–40. bile canaliculi formation. Integr Biol (Camb), 2013;5(2):390–401.
64. Lazaro‐Dieguez, F. et al. Par1b links lumen polarity with LGN‐NuMA posi- 91. Godoy, P. et al. Recent advances in 2D and 3D in vitro systems using pri-
tioning for distinct epithelial cell division phenotypes. J Cell Biol, mary hepatocytes, alternative hepatocyte sources and non‐parenchymal
2013;203(2):251–64. liver cells and their use in investigating mechanisms of hepatotoxicity, cell
65. Lazaro‐Dieguez, F. and Musch, A. Cell‐cell adhesion accounts for the differ- signaling and ADME. Arch Toxicol, 2013;87(8):1315–530.
ent orientation of columnar and hepatocyte cell division. J Cell Biol, 92. Alessi, D.R., Sakamoto, K., and Bayascas, J.R. LKB1‐dependent signaling
2017;216(11):3847–59. pathways. Annu Rev Biochem, 2006;75:137–63.
66. Bryant, D.M. et al. A molecular network for de novo generation of the apical 93. Baas, A.F. et al. Complete polarization of single intestinal epithelial cells
surface and lumen. Nat Cell Biol, 2010;12(11):1035–45. upon activation of LKB1 by STRAD. Cell, 2004;116(3):457–66.
67. Galvez‐Santisteban, M. et al. Synaptotagmin‐like proteins control the forma- 94. Zhang, L. et al. AMP‐activated protein kinase regulates the assembly of epi-
tion of a single apical membrane domain in epithelial cells. Nat Cell Biol, thelial tight junctions. Proc Natl Acad Sci U S A, 2006;103(46):17272–7.
2012;14(8):838–49. 95. Zheng, B. and Cantley, L.C. Regulation of epithelial tight junction assem-
68. Tuma, P. and Hubbard, A.L. Transcytosis: crossing cellular barriers. Physiol bly and disassembly by AMP‐activated protein kinase. Proc Natl Acad Sci
Rev, 2003;83(3):871–932. U S A, 2007;104(3):819–22.
69. Weisz, O.A. and Rodriguez‐Boulan, E. Apical trafficking in epithelial cells: 96. Aznar, N. et al. AMP‐activated protein kinase fortifies epithelial tight junc-
signals, clusters and motors. J Cell Sci, 2009;122(Pt 23):4253–66. tions during energetic stress via its effector GIV/Girdin. Elife, 2016;5.
70. Boulant, S. et al. Actin dynamics counteract membrane tension during clath- 97. Yano, T. et  al. The association of microtubules with tight junctions is
rin‐mediated endocytosis. Nat Cell Biol, 2011;13(9):1124–31. promoted by cingulin phosphorylation by AMPK. J Cell Biol, 2013;
­
71. Szalinski, C.M. et al. PIP5KIbeta selectively modulates apical endocytosis 203(4):605–14.
in polarized renal epithelial cells. PLoS One, 2013;8(1):e53790. 98. Woods, A. et al. LKB1 is required for hepatic bile acid transport and cana-
72. Sato, T. et al. Regulation of the assembly and adhesion activity of E‐cadherin licular membrane integrity in mice. Biochem J, 2011;434(1):49–60.
by nectin and afadin for the formation of adherens junctions in Madin‐Darby 99. Fu, D. et al. Coordinated elevation of mitochondrial oxidative phosphoryla-
canine kidney cells. J Biol Chem, 2006;281(8):5288–99. tion and autophagy help drive hepatocyte polarization. Proc Natl Acad Sci
73. Takai, Y. et  al. Nectins and nectin‐like molecules: roles in cell adhesion, U S A, 2013;110(18):7288–93.
migration, and polarization. Cancer Sci, 2003;94(8):655–67. 100. Moghe, P.V. et  al. Culture matrix configuration and composition in the
74. Tsuchiya, B. et al. Differential expression of N‐cadherin and E‐cadherin in maintenance of hepatocyte polarity and function. Biomaterials,
normal human tissues. Arch Histol Cytol, 2006;69(2):135–45. 1996;17(3):373–85.
75. Straub, B.K. et al. E‐N‐cadherin heterodimers define novel adherens junc- 101. Mooney, D. et al. Switching from differentiation to growth in hepatocytes:
tions connecting endoderm‐derived cells. J Cell Biol, 2011;195(5):873–87. control by extracellular matrix. J Cell Physiol, 1992;151(3):497–505.
76. Wheelock, M.J. et  al. Cadherin switching. J Cell Sci, 2008;121(Pt 102. Ezzell, R.M. et al. Effect of collagen gel configuration on the cytoskeleton
6):727–35. in cultured rat hepatocytes. Exp Cell Res, 1993;208(2):442–52.
77. Rajasekaran, A.K. et al. Catenins and zonula occludens‐1 form a complex 103. Berthiaume, F. et al. Effect of extracellular matrix topology on cell struc-
during early stages in the assembly of tight junctions. J Cell Biol, ture, function, and physiological responsiveness: hepatocytes cultured in a
1996;132(3):451–63. sandwich configuration. FASEB J, 1996;10(13):1471–84.
48 THE LIVER: REFERENCES

104. Fassett, J., Tobolt, D., and Hansen, L.K. Type I collagen structure regulates 131. Sheff, D.R. et  al. The receptor recycling pathway contains two distinct
cell morphology and EGF signaling in primary rat hepatocytes through populations of early endosomes with different sorting functions. J Cell
cAMP‐dependent protein kinase A. Mol Biol Cell, 2006;17(1):345–56. Biol, 1999;145(1):123–39.
105. Godoy, P. et al. Extracellular matrix modulates sensitivity of hepatocytes to 132. Wang, E. et al. Apical and basolateral endocytic pathways of MDCK cells
fibroblastoid dedifferentiation and transforming growth factor beta‐induced meet in acidic common endosomes distinct from a nearly‐neutral apical
apoptosis. Hepatology, 2009;49(6):2031–43. recycling endosome. Traffic, 2000;1(6):480–93.
106. Desai, S.S. et al. Physiological ranges of matrix rigidity modulate primary 133. Maxfield, F.R. and McGraw, T.E. Endocytic recycling. Nat Rev Mol Cell
mouse hepatocyte function in part through hepatocyte nuclear factor 4 Biol, 2004;5(2):121–32.
alpha. Hepatology, 2016;64(1):261–75. 134. Odorizzi, G. et al. Apical and basolateral endosomes of MDCK cells are
107. Rodriguez‐Boulan, E., Kreitzer, G., and Musch, A. Organization of vesicu- interconnected and contain a polarized sorting mechanism. J Cell Biol,
lar trafficking in epithelia. Nat Rev Mol Cell Biol, 2005;6(3):233–47. 1996;135(1):139–52.
108. Bartles, J.R. et al. Biogenesis of the rat hepatocyte plasma membrane in 135. Apodaca, G., Katz, L.A., and Mostov, K.E. Receptor‐mediated transcytosis
vivo: comparison of the pathways taken by apical and basolateral proteins of IgA in MDCK cells is via apical recycling endosomes. J Cell Biol,
using subcellular fractionation. J Cell Biol, 1987;105(3):1241–51. 1994;125(1):67–86.
109. Kipp, H. and Arias, I.M. Newly synthesized canalicular ABC transporters 136. Barroso, M. and Sztul, E.S. Basolateral to apical transcytosis in polarized
are directly targeted from the Golgi to the hepatocyte apical domain in rat cells is indirect and involves BFA and trimeric G protein sensitive passage
liver. J Biol Chem, 2000;275(21):15917–25. through the apical endosome. J Cell Biol, 1994;124(1–2):83–100.
110. Schell, M.J. et al. 5′nucleotidase is sorted to the apical domain of hepato- 137. Rahner, C., Stieger, B., and Landmann, L. Apical endocytosis in rat hepato-
cytes via an indirect route. J Cell Biol, 1992;119(5):1173–82. cytes In situ involves clathrin, traverses a subapical compartment, and leads
111. Barr, V.A., Scott, L.J., and Hubbard, A.L. Immunoadsorption of hepatic to lysosomes. Gastroenterology, 2000;119(6):1692–707.
vesicles carrying newly synthesized dipeptidyl peptidase IV and polymeric 138. Marbet, P. et  al. Quantitative microscopy reveals 3D organization and
IgA receptor. J Biol Chem, 1995;270(46):27834–44. kinetics of endocytosis in rat hepatocytes. Microsc Res Tech,
112. Saucan, L. and Palade, G.E. Differential colchicine effects on the transport 2006;69(9):693–707.
of membrane and secretory proteins in rat hepatocytes in vivo: bipolar 139. Hoekstra, D., Tyteca, D., and van IJzendoorn, S.C. The subapical compart-
secretion of albumin. Hepatology, 1992;15(4):714–21. ment: a traffic center in membrane polarity development. J Cell Sci,
113. Bastaki, M. et al. Absence of direct delivery for single transmembrane api- 2004;117(Pt 11):2183–92.
cal proteins or their “Secretory” forms in polarized hepatic cells. Mol Biol 140. Casanova, J.E. et al. Association of Rab25 and Rab11a with the apical recy-
Cell, 2002;13(1):225–37. cling system of polarized Madin‐Darby canine kidney cells. Mol Biol Cell,
114. Deborde, S. et al. Clathrin is a key regulator of basolateral polarity. Nature, 1999;10(1):47–61.
2008;452(7188):719–23. 141. Hales, C.M., Vaerman, J.P., and Goldenring, J.R. Rab11 family interacting
115. Gravotta, D. et al. The clathrin adaptor AP‐1A mediates basolateral polar- protein 2 associates with Myosin Vb and regulates plasma membrane recy-
ity. Dev Cell, 2012;22(4):811–23. cling. J Biol Chem, 2002;277(52):50415–21.
116. Folsch, H. Role of the epithelial cell‐specific clathrin adaptor complex 142. Jing, J. and Prekeris, R. Polarized endocytic transport: the roles of Rab11
AP‐1B in cell polarity. Cell Logist, 2015;5(2):e1074331. and Rab11‐FIPs in regulating cell polarity. Histol Histopathol,
117. Schuck, S. and Simons, K. Polarized sorting in epithelial cells: raft cluster- 2009;24(9):1171–80.
ing and the biogenesis of the apical membrane. J Cell Sci, 2004;117(Pt 143. Lapierre, L.A. et al. Phosphorylation of Rab11‐FIP2 regulates polarity in
25):5955–64. MDCK cells. Mol Biol Cell, 2012;23(12):2302–18.
118. Yeaman, C. et al. Protein kinase D regulates basolateral membrane protein 144. Fu, D. et al. Regulation of bile canalicular network formation and mainte-
exit from trans‐Golgi network. Nat Cell Biol, 2004;6(2):106–12. nance by AMP‐activated protein kinase and LKB1. J Cell Sci, 2010;123(Pt
119. Valente, C., Luini, A., and Corda, D. Components of the CtBP1/BARS‐ 19):3294–302.
dependent fission machinery. Histochem Cell Biol, 2013;140(4):407–21. 145. Wakabayashi, Y. et al. Rab11a and myosin Vb are required for bile canali-
120. Pagliuso, A. et  al. Golgi membrane fission requires the CtBP1‐S/BARS‐ cular formation in WIF‐B9 cells. Proc Natl Acad Sci U S A,
induced activation of lysophosphatidic acid acyltransferase delta. Nat 2005;102(42):15087–92.
Commun, 2016;7:12148. 146. Wakabayashi, Y., Kipp, H., and Arias, I.M. Transporters on demand: intra-
121. Kreitzer, G. et al. Kinesin and dynamin are required for post‐Golgi trans- cellular reservoirs and cycling of bile canalicular ABC transporters. J Biol
port of a plasma‐membrane protein. Nat Cell Biol, 2000;2(2):125–7. Chem, 2006;281(38):27669–73.
122.. Salvarezza, S.B. et  al. LIM kinase 1 and cofilin regulate actin filament 147. Gatmaitan, Z.C., Nies, A.T., and Arias, I.M. Regulation and translocation of
population required for dynamin‐dependent apical carrier fission from the ATP‐dependent apical membrane proteins in rat liver. Am J Physiol,
trans‐Golgi network. Mol Biol Cell, 2009;20(1):438–51. 1997;272(5 Pt 1):G1041–9.
123. Breuza, L., Fransen, J., and Le Bivic, A. Transport and function of syntaxin 148. Misra, S. et  al. The role of phosphoinositide 3‐kinase in taurocholate‐
3 in human epithelial intestinal cells. Am J Physiol Cell Physiol, induced trafficking of ATP‐dependent canalicular transporters in rat liver. J
2000;279(4):C1239–48. Biol Chem, 1998;273(41):26638–44.
124. Ikonen, E. et  al. Different requirements for NSF, SNAP, and Rab pro- 149. Misra, S. et  al. Phosphoinositide 3‐kinase lipid products regulate ATP‐
teins  in apical and basolateral transport in MDCK cells. Cell, dependent transport by sister of P‐glycoprotein and multidrug resistance
1995;81(4):571–80. associated protein 2 in bile canalicular membrane vesicles. Proc Natl Acad
125. Low, S.H. et al. Differential localization of syntaxin isoforms in polarized Sci U S A, 1999;96(10):5814–19.
Madin‐Darby canine kidney cells. Mol Biol Cell, 1996;7(12):2007–18. 150. Misra, S., Varticovski, L., and Arias, I.M. Mechanisms by which cAMP
126. Pocard, T. et  al. Distinct v‐SNAREs regulate direct and indirect apical increases bile acid secretion in rat liver and canalicular membrane vesicles.
delivery in polarized epithelial cells. J Cell Sci, 2007;120(Pt 18):3309–20. Am J Physiol Gastrointest Liver Physiol, 2003;285(2):G316–24.
127. Reales, E. et al. Basolateral sorting of syntaxin 4 is dependent on its N‐ter- 151. Cao, H. et  al. Cortactin is a component of clathrin‐coated pits and
minal domain and the AP1B clathrin adaptor, and required for the epithelial participates in receptor‐mediated endocytosis. Mol Cell Biol,
­
cell polarity. PLoS One, 2011;6(6):e21181. 2003;23(6):2162–70.
128. Puertollano, R. et al. The MAL proteolipid is necessary for normal apical 152. Smythe, E. and Ayscough, K.R. Actin regulation in endocytosis. J Cell Sci,
transport and accurate sorting of the influenza virus hemagglutinin in 2006;119(Pt 22):4589–98.
Madin‐Darby canine kidney cells. J Cell Biol, 1999;145(1):141–51. 153. Ortiz, D.F. et al. Identification of HAX‐1 as a protein that binds bile salt
129. Cheong, K.H. et al. VIP17/MAL, a lipid raft‐associated protein, is involved export protein and regulates its abundance in the apical membrane of
in apical transport in MDCK cells. Proc Natl Acad Sci U S A, Madin‐Darby canine kidney cells. J Biol Chem, 2004;279(31):32761–70.
1999;96(11):6241–8. 154. Miyaoka, Y. et al. Hypertrophy and unconventional cell division of hepato-
130. Ramnarayanan, S.P. et al. Exogenous MAL reroutes selected hepatic apical cytes underlie liver regeneration. Curr Biol, 2012;22(13):1166–75.
proteins into the direct pathway in WIF‐B cells. Mol Biol Cell, 155. Ragkousi, K. and Gibson, M.C. Cell division and the maintenance of epi-
2007;18(7):2707–15. thelial order. J Cell Biol, 2014;207(2):181–8.
4:  Hepatocyte Surface Polarity 49

156. Reinsch, S. and Karsenti, E. Orientation of spindle axis and distribution of 160. Bartles, J.R. and Hubbard, A.L. Preservation of hepatocyte plasma mem-
plasma membrane proteins during cell division in polarized MDCKII cells. brane domains during cell division in situ in regenerating rat liver. Dev Biol,
J Cell Biol, 1994;126(6):1509–26. 1986;118(1):286–95.
157. Tuncay, H. and Ebnet, K. Cell adhesion molecule control of planar spindle 161. Slim, C.L. et  al. Par1b induces asymmetric inheritance of plasma mem-
orientation. Cell Mol Life Sci, 2016;73(6):1195–207. brane domains via LGN‐dependent mitotic spindle orientation in proliferat-
158. McNally, F.J. Mechanisms of spindle positioning. J Cell Biol, ing hepatocytes. PLoS Biol, 2013;11(12):e1001739.
2013;200(2):131–40. 162. Meyer, K. et  al. A predictive 3D multi‐scale model of biliary fluid
159. Kotak, S. and Gonczy, P. Mechanisms of spindle positioning: cortical force ­dynamics in the liver lobule. Cell Syst, 2017;4(3):277–90.
generators in the limelight. Curr Opin Cell Biol, 2013;25(6):741–8.
Primary Cilia
5 Carolyn M. Ott
Janelia Research Campus, Ashburn, VA, USA

INTRODUCTION Primary cilia are present on a subset of cells in the liver;


cholangocytes have primary cilia, while hepatocytes do not. The
A multicellular organism has a need to coordinate multiple pro- absence of cilia may be essential to specialization. Outside the
cesses in space and time. During development, multiple cell lin- hematopoietic lineage, all progenitor cells have primary cilia, so
eages migrate and differentiate simultaneously. Individual organ cilia‐less, differentiated cells have lost the ability to respond to
function, though physically isolated, must be coordinated with messages in the same way, and thus may be insulated. Like
the activity of other organs. Similarly, each specialized cell hepatocytes, adipocytes and myocytes lack cilia, but are derived
coordinates with neighboring cells. Coordination is achieved as from ciliated progenitors. The liver is a source of signals that
individual cells respond to messages from other cells and the can be perceived by primary cilia elsewhere in the body. For
environment, and transmit signals about their own state and example, insulin‐like growth factor‐1 (IGF‐1) is produced by
needs. Amid the cacophony of potential messages, individual the liver. The IGF‐1 receptor is expressed in diverse tissues and
cells must discern which messages are relevant. Cilia project can localize to cilia. IGF‐1 signaling through cilia can initiate
away from the cell surface and function as a tunable sensing cilia resorption as cells enter the mitotic cycle [3].
organelle. Because cilia are continuous with both the cellular Cells have many modes of perceiving the extracellular envi-
cytoplasm and plasma membrane, specialized barriers restrict ronment. What advantages do cilia provide for signal perception
entry and exit so that cells can form and adjust the composition and transmission? Location, structure, isolation, and size make
of cilia to sense and respond to signals. the primary cilium a unique environment.
The term “cilia” has been generalized to encompass all
types of ciliated structures, including flagella, and motile and Location
non‐motile cilia. Cilia and flagella, present throughout the
eukaryotic lineage, are distinct in structure and composition Like a probe or an antenna, primary cilia extend away from the
from prokaryotic flagella. In eukaryotes, microtubules pro- cell and are immersed in the extracellular environment. The
vide both structural support and a transportation highway structure is positioned to detect and report the conditions out-
within cilia. Unicellular eukaryotes use cilia to process side the cell. Cells can specify which part of the environment to
responses to stimuli from the environment, mate, and feed. In monitor using the cilium. For example, polarized cells typically
mice and other mammals, genetic ablation of primary cilia position cilia at the apical surface.
terminates development. In adults, ciliary loss leads to dis-
ease [1]. Primary cilia are typically solitary projections from
the cell surface and they lack the structural components that
Structure
facilitate active beating in motile cilia and flagella. However, Two structural features of primary cilia are essential to consider.
primary cilia are not stationary, but can move with the First, because the cilium is a long narrow structure, it has a very
extracellular environment or pivot due to forces from the high surface area‐to‐volume ratio. The plasma membrane pro-
intracellular actin cytoskeleton [2]. vides a platform for detecting the extracellular environment.

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
5:  Primary Cilia 51

Second, the polarized microtubules that provide structural sup- directed movement along axonemal microtubules (reviewed in
port also act as highways for traffic along the length of the cil- [9]). IFT was first discovered as movement in the flagella of the
ium. Transport along these microtubules can be regulated and unicellular alga Chlamydomonas reinhardtii [10]. The IFT
facilitates signaling. Tubulin in the microtubules themselves complexes form trains (approximately 200 nm long in
may also have direct effects on proteins in the membrane [4]. Chlamydomonas flagella) that bind to microtubule motor pro-
teins, cargo, such as tubulin, and cargo adaptors that bind solu-
ble and membrane proteins. IFT complexes are large multiprotein
Isolation complexes about the size of the ribosomal small subunit (16S)
Although small proteins may be able to access primary cilia, that can be biochemically separated into two subcomplexes,
entry of larger proteins is restricted and targeting sequences or IFT‐A and IFT‐B [11]. The IFT‐B subcomplexes associate with
interacting partners are thought to be essential. Regulation of kinesin‐2 to facilitate anterograde transport along the B tubules
both the cytoplasmic and plasma membrane content of the pri- of the axoneme doublets. The retrograde motor, dynein, is car-
mary cilium tunes the cilium to detect specific ligands. In addi- ried to the plus end of ciliary microtubules through association
tion, downstream signaling molecules within the cilium are with the IFT‐A subcomplex. At the tip, kinesin‐2 dissociates
protected from modification by enzymes in the cytoplasm. from the IFT complex, retrograde trains form, and IFT‐A‐asso-
Although ions can pass between the cilium and the cytoplasm, ciated dynein drives retrograde movement along the A tubules
channels localized in the ciliary plasma membrane create a of the axoneme doublets (Figure 5.1) [12]. Two kinesin‐2 family
unique ionic environment inside cilia that may influence protein members, Kif3 and Kif17, process along ciliary microtubules:
activity [5]. Kif3 is required for ciliogenesis while Kif17 appears to partici-
pate in more specialized trafficking. The specific tubulin iso-
types in cilia, as well as the multiple tubulin modifications are
Size thought to affect IFT transport.
The volume of a 5 μm cilium is approximately 0.35 μm3, which Membrane surrounds cilia microtubules as they extend away
is less than 0.05% of the total eukaryotic cell volume [6]. As a from the centriole (also called a basal body). While some cen-
result, within the cilium the concentration of a single molecule trosomes are at the cell surface, many are recessed and have a
is much larger than it would be in the cytoplasm. In addition, cililary pocket: a specialized membrane domain with active
because changes in cilia length change cilia volume, concentra- exocytosis and endocytosis [13]. The base of the cilium, near
tions within the cilium could also be regulated by processes that the centriole, serves as the access point for both cytoplasmic and
affect cilia growth. membrane components. The composition of a cilium defines the
sensitivity and responsiveness of the organelle. While the ciliary
membrane is continuous with the plasma membrane of the cell,
access to both the ciliary membrane and cytoplasm is restricted
CILIUM COMPOSITION by a boundary created by a structure located inside the cilium,
AND ORGANIZATION just above the centrosome, called the transition zone. Mutations
in transition zone proteins or other proteins responsible for cre-
In the cytoplasm, most microtubules are hollow tubes created by ating and regulating cilia composition are responsible for many
assembly of α and β tubulin monomers assembled into 13 cilia‐related diseases, termed ciliopathies. Protein entry through
protofilaments. Cilia microtubules are doublets composed the transition zone is gated and although small molecules can
of a 13‐protofilament tubule (designated the A tubule) and a pass through the transition zone, several active processes main-
10‐protofilament B tubule that closes onto the outside of tain the unique environment of the cilium.
three A tubule protofilaments (Figure 5.1). Ciliary microtubules Most ciliopathies are pleiotropic diseases that affect several
receive several post‐translational modifications: acetylation, overlapping organ systems including liver, kidney, retina, heart,
detyrosination, glutamylation, and glycylation (Figure 5.2a) [7]. and bones [1]. Several ciliopathies, including nephronophthisis
The nine doublet microtubules are direct extensions of centriole (NPHP), Meckel syndrome (MKS), and Joubert syndrome, are
microtubules. As a consequence, primary cilia are always caused by mutation of transition zone proteins [14]. While often
anchored in the cytoplasm. Collectively, cilia microtubes are considered a single defined structure, transition zone organiza-
referred to as an axoneme. Although polarized, axoneme micro- tion and length can vary between cell types [15]. Tissue‐specific
tubules do not treadmill like cytoplasmic microtubules because differences in transition zones might contribute to the differ-
there is no depolymerization at the minus end within the centro- ences in symptoms upon loss or mutation of different transition
some. Axoneme length is mediated by addition or removal of zone components.
tubulin at the plus end at the cilium tip. Unlike flagella and Genetic, biochemical, and proteomic strategies identified
motile cilia that have internal structures to maintain the radial many components of the transition zone [16–18]. By electron
symmetry of the microtubules along the entire length of the microscopy the transition zone is visible as “Y”‐shaped bridges
cilium, primary cilia microtubules can rearrange toward the dis- with two branches attached to the ciliary membrane and a single
tal end. Primary cilia can also narrow as individual doublets ter- anchor at each microtubule doublet (Figure 5.1). Subdiffraction
minate (Figure 5.1) [8]. imaging strategies have been employed to reconcile the parts list
As protein synthesis does not occur in cilia, growth requires with the visible structure, and to begin to attribute functions to
tubulin monomers to be imported into the cilium and travel to individual components. Three key transition zone protein
the tip. The intraflagellar transport (IFT) complex facilitates complexes were named in part after the diseases caused by
52 THE LIVER:  CILIUM COMPOSITION AND ORGANIZATION

Tubulin kin
dimer es
in

(+)

IFT-A
IFT-B

Microtubule dynein
doublet
IFT-B IFT-A

kinesin

Primary cilium

dynein

Y-link
Intraflagellar transport
(IFT)

Transition zone
Ciliary pocket

Basal body
(Mother centriole)

Distal appendages

200 nm

Figure 5.1  Primary cilia architecture and intraflagellar transport. The cross‐sections show the structural changes as the microtubules extend from
the centrosome (described from bottom to top). The drawings are based on EM images in [8]. At the centrosome there are nine microtubule triplets.
Distal appendage proteins appear like a pinwheel emanating from the centriole microtubules. At the transition zone, Y‐links are visible. The doublet
microtubules of the ciliary axoneme are extensions of two of the triplet centriolar microtubules. Toward the distal end, the axoneme doublets can
rearrange and terminate. The doublets are composed of α and β tubulin dimers that assemble a closed cylinder (A tubule) and a second cylinder that
closes on the outside of the first (B tubule). On the right is a representation of intraflagellar transport (IFT). The IFT complex is a large multi‐subunit
complex that can be biochemically separated into the IFT‐A and IFT‐B subcomplexes. The IFT‐A subcomplex interacts with dynein and facilitates
retrograde transport along the A tubule. For anterograde transport toward the plus end of the microtubules at the cilium tip, the IFT‐B subcomplex
associates with kinesin‐2. IFT complexes associate with cargo in both directions. The retrograde motor, dynein, is an anterograde cargo, but the
kinesin motor dissociates at the tip and diffuses back to the base. Because the time for kinesin‐2 to return to the base is length‐dependent, the con-
centration of kinesin‐2 has been proposed as a length regulator. Adapted from Chien 2017 and Hendel 2018 [109, 110].

mutation of the complex. MKS and NPHP complex proteins The structure of the transition zone cannot yet explain its
create the Y‐links while Cep290 localizes adjacent to the Y‐ function as a barrier capable of selective transport. It has
links, proximal to the centrosome [19, 20]. How stable are the been shown, however, that depletion of transition zone pro-
Y‐link structures? Mating experiments in Chlamydomonas teins that cause disease permit typically excluded proteins
revealed NPHP4 is static, but CEP290 can exchange between into the cilium [22]. In addition to restricting entry, the tran-
transition zones of different flagella [21, 22]. Researchers found sition zone can also limit exit and may function as a staging
no fluorescent recovery within 30 min after photobleaching platform because several dynamic cilia proteins have been
green fluorescent protein‐tagged MKS complex proteins indi- found to accumulate above this region [24, 25]. To pass
cating that the proteins do not exchange with the unbleached through the barrier, many proteins utilize escorts such as
pool [23]. These results suggests that the MKS and NPHP com- nuclear import effectors, the BBSome complex, and tubby
plexes form stable structures, while the pool of CEP290 may be family proteins (the latter two are discussed in the section
more dynamic. later on ciliary signaling).
5:  Primary Cilia 53

(a) (c)
CEP164 FBF1 Merge

200 nm
(d)
FBF1 and SCLT1 CEP164 and SCLT1

5 m
200 nm
anti-acetylated tubulin anti-detyrosinated tubulin Hoecst
(e)
(b) DAM
Y-links

Ciliary
pocket
DABs

xy 5 m
CEP164 FBF1 SCLT1 CEP89 CEP83
xz
cilia-enriched GPCR centrosome CP110 T TBK2 IFT88 EHD1 Arl13B

Figure 5.2  Conventional and super‐resolution views of primary cilia. (a) Microtubules within cilia are post‐translationally modified. These cilia
on Madin–Darby canine kidney (MDCK) cells have been stained with antibodies that recognize acetylated and detyrosinated tubulin and visualized
using confocal microscopy. The image is a maximum intensity projection of a z‐stack. The inset images show that both acetylated and detyrosinated
tubulin are found along the entire length, however the distributions of the two are not identical. (b) Cilia‐enriched GPCR localization. Live imaging
of NIH3T3 fibroblast cell shows the concentration of the GPCR adjacent to the centrosome in cells expressing the GPCR, melanin concentrating
hormone receptor 1 (tagged with the fluorescent protein tdTomato), and the centrosome localization sequence of pericentrin (PACT; tagged with
tag‐blue fluorescent protein). The upper image is an xy projection; the narrow, lower panel is an xz projection of the same cell. (c–e) Super‐resolu-
tion imaging brings together protein localization and structure to reveal organization at the base of the cilium. (c) Two‐color axial view dSTORM
images of distal appendage proteins, FBF1 and CEP164. This approach was used to calculate the relative angular positions of different distal
appendage proteins (represented by symbols in the far right image). From [92]. (d) Two‐color lateral view dSTORM images of FBF1, SCLT1, and
CEP164 (three representative images of each). The relative height of FBF and SCLT1 are designated by the arrowheads on the far left. From [92].
(e) Three‐dimensional computational model created by combining information about the relative axial and lateral positions of distal appendage
proteins and several ciliary proteins. DAB, dorsal appendage blade; DAM, dorsal appendage matrix. From Yang 2018 https://www.nature.com/
articles/s41467‐018‐04469‐1#rightslink. Licensed under CCBY 4.0 [92].

The mechanism remains to be elucidated, but several lines of cytoplasm. Like the nucleus, the cilium is also rich in RanGTP
evidence suggest that entry and exit of ciliary matrix proteins is and the nuclear import adaptor importin β binds to ciliary locali-
related to gating at the nuclear pore. Small soluble proteins can zation sequences and can facilitate delivery of soluble proteins
freely enter the cilium, but, like the nuclear pore that excludes through the boundary at the base of the cilium [29]. Because
proteins above ~50 kDa, large proteins require active transport primary cilia lack structures that resemble nuclear pores, it is
into the primary cilium [26]. Several fluorescently labeled not yet clear where NUP proteins assemble or how they create a
nucleoporin (NUP) proteins localize to the base of the primary permeability barrier.
cilium and disrupting their function alters the transport of pro- Although the full definition and diversity of ciliary lipids is
teins into and out of cilia [27, 28]. The NUPs found in cilia are not known, several lines of evidence indicate that, like the pro-
not the membrane or inner, nuclear basket proteins, but rather tein composition, the lipid environment of cilia is unique and
NUPs that contain phenylalanine–glycine repeats that are actively regulated. Early studies revealed that cilia membranes
thought to create the permeability barrier. To pass through the are generally cholesterol rich, however, the transition zone may
nuclear pore, cargo with import or export signals bind to trans- have less cholesterol [30, 31]. Cilia have condensed membranes
port adaptor proteins, such as importin β. Adaptor binding and and ciliogenesis requires phosphatidylinositol 4‐phosphate
release on the opposite side of the nuclear envelope is facilitated adaptor protein‐2 (FAPP2) [32]. One of the causes of the
by a gradient of RanGTP inside the nucleus and RanGDP in the ciliopathy Joubert syndrome is mutation of INPP5E, a ciliary
54 THE LIVER:  CILIA‐MEDIATED SIGNALING

phosphoinositide 5‐phosphatase that localizes along the length have co‐evolved with motility in early organisms as suggested
of the cilium, where it facilitates the conversion of phosphoi- by a study that revealed several cilia signaling pathway compo-
nositide 4,5‐bisphosphate (PI(4,5)P2) to phosphatidylinositol nents in organisms that are thought to share an ancient early
4‐phosphate (PI(4)P) [33–36]. While adjacent membranes are ancestor with modern animals [43].
rich in PI(4,5)P2, INPP5E localization to cilia creates a PI(4)P‐ Olfaction and vision are both sensory functions mediated by
rich membrane. Targeting of INPP5E and other farnesylated primary cilia [44]. Rod outer segments are modified cilia filled
proteins to cilia is thought to be mediated by release of cargo with membrane disks containing rhodopsin, a G protein‐
from the prenyl‐binding protein, phosphodiesterase 6 delta coupled receptor (GPCR). Smell is the consequence of an
subunit, mediated by Arl3 at the cilium [37]. Disrupting the action potential initiated by ligand binding to olfactory GPCRs
transition zone permits PI(4,5)P2 phospholipids to enter ­primary on the cilia of olfactory neurons. Stereocilia, which detect
cilia [38]. sound during hearing, are actin‐based structures and therefore
As mentioned earlier, the volume of the cilium is a fraction of are not true cilia. The hair cells that project stereocilia do have
the volume of the rest of the cell and it is membrane bound a true cilium, called a kinocilium. Although the kinocilium is
except at the base. Although ions can pass into and out of the not required for mechanosensation of sound, mutations that
base of the cilium, transport of ions through ciliary membrane prevent cilia formation on hair cells disrupt proper stereocilia
channels helps maintain the unique ionic environment inside organization possibly through failures to signal polarity cues
cilia. The resting membrane potential of cilia on cultured human during development [45].
retina pigment epithelial cells is ~30 mV more positive than the Sensing and signaling performed by primary cilia are essen-
resting potential of the plasma membrane of the rest of the cell tial for differentiation and tissue morphogenesis during devel-
[39]. The concentration of Ca2+ in cilia is approximately seven opment. Mutations that cause systemic loss of primary cilia
times higher than in the cytoplasm of cultured human pigment result in embryonic lethality accompanied by patterning
epithelial cells [39]. While the cytoplasm and cytoplasmic orga- defects such as left–right asymmetry [46, 47]. Loss of cilia in
nelles serve as a sink for Ca2+, channels in the ciliary membrane adult animals leads to kidney cyst formation and obesity due to
may import enough Ca2+ to maintain a concentration greater abnormal appetite signaling [48]. Engineered loss of cilia in
than 500 nM. defined cell types causes many additional phenotypes (see, for
Ciliary ion channels are essential components of the vision example, [49–51]). Some ciliary proteins may contribute to
and olfaction signal transduction pathways. Several other ciliary signaling in cancer [52] and signaling through primary cilia
ion channels, including polycystin 2 (PKD2) and PKD2‐like 1 also affects response to injury [52, 53]. Primary cilia in diverse
(PKD2‐L1), are members of the transient receptor potential tissues do not participate in a single, stereotyped signaling
(TRP) family of cation channels. Homomeric PKD2 conducts pathway. Rather, the expression and localization of receptors
Na+ and K+ and is impermeable to Ca2+, while PKD2‐L1 is per- and downstream effectors varies by both cell type and life
meable to Na+, K+, and Ca2+ [40, 41]. The ionic environment stage. There is a growing list of genes that contribute to cili-
defined by the channels contributes to cell and organismal opathies; mutation of many of which alter signal perception or
health. Mutations in PKD2 cause autosomal dominant polycys- transmission [1].
tic kidney disease (ADPKD) and deletion of the gene causes A key paradigm for ciliary signaling is through GPCRs.
embryonic lethality. Without PKD2‐L1, mice have reduced Figure 5.3 summarizes some of the potential cilia GPCR signal
responses to hedgehog signaling and mild hedgehog‐related transduction outcomes described below. Similar to rhodopsin
developmental phenotypes [39]. It is not yet clear yet which and olfactory GPCRs, cilia targeting of many GPCRs is so effi-
cilia proteins are influenced by the unique ionic environment. cient that the concentration of the GPCR in the ciliary mem-
Metabolites can pass between the cilium and the cytoplasm; brane far exceeds the receptor concentration anywhere else in
however, evidence suggests that active processes can sustain the cell (Figure  5.2b). Sequence features on the intracellular
different metabolite concentrations in cilia relative to the cyto- face GPCRs contribute to ciliary targeting [53–55]. These fea-
plasm as a whole. Several isoforms of adenylyl cyclase, which tures are recognized by tubby (Tub) family proteins including
generates cyclic adenosine monophosphate (cAMP) through a Tub and tubby‐like protein 3 (TULP3). Tub and TULP3 associ-
cyclizing reaction of ATP, localize to primary cilia [39, 40]. ate with the plasma membrane through binding PI(4,5)P2,
cAMP concentrations in primary cilia are ~5 times higher than where they capture GPCRs with cilia localization sequences.
in the cell body and the levels can be altered through ciliary Tub and TULP3 binding to IFT‐A facilitates cilia delivery.
signaling pathways [42]. Ciliary Ca2+ concentration also regu- Within the cilium there is little PI(4,5)P2, so Tub and TULP3
lates adenylyl cyclase activity, and thus could regulate ciliary dissociate from the membrane and GPCR cargos are released
signal responses. [56]. Loss of Tub, which is predominantly expressed in the
brain, prevents cilia localization of GPCRs critical for appetite
regulation and results in obesity [57].
Delivery of membrane proteins, including GPCRs and ion
CILIA‐MEDIATED SIGNALING channels, to the base of the cilium can be mediated by lateral
passage from the plasma membrane, direct targeting of Golgi‐
Primary cilia are positioned to sample the extracellular environ- derived vesicles or indirect trafficking through endosome‐
ment near the cell. The ability to detect defined molecules and derived vesicles [58]. Bardet–Biedl syndrome (BBS) is a
signal their presence to the rest of the cell is a feature primary ciliopathy caused by mutation in proteins that form a complex
cilia share with motile cilia and flagella. Sensory functions may called the BBSome [59, 60]. GPCRs are absent from cilia of
5:  Primary Cilia 55

(a) Without ligand


?

Rab8 Tub GPCR


?
BBSome
PI(4)P Adenylyl
cyclase
Tub GDP
PKA
Kinase
target

Ion
channel

(b) Upon ligand binding

Ligand

PI(4)P Adenylyl
cyclase
cAMP cAMP
+ GTP
cAMP
P
PKA
P
BBSome

Possible consequences of ligand binding

G Protein GTP
GDP

Adenylyl cyclase ATP or cAMP

GPCR cilium exit with BBSome

Kinases inactive ACTIVE

Channels closed open

Figure 5.3  Generalized perspective of signaling through primary cilia. (a) Many GPCRs localize to primary cilia. Trafficking of GPCRs to
primary cilia is mediated by Rab8 and the BBSome. Tub family proteins escort GPCRs into primary cilia and release them when they reach the
PI(4)P‐rich membrane (orange). In the absence of ligand, GPCRs can associate with G proteins. (b) Ligand binding to GPCRs can have many
potential consequences. Activated receptors can trigger replacement of Gα‐associated GDP with GTP. As a result, the GTP‐bound Gα subunit dissociates
from the G protein β/γ subunits and both propagate signaling cascades. One possibility shown here is that Gα can stimulate adenylyl cyclase, which
produces cAMP. cAMP can activate kinases such as PKA and could affect ion channel activity. Receptor activation also leads to BBSome‐dependent
internalization of GPCRs that may reset signaling pathways. As indicated at the bottom, signaling cascades could also have the opposite effects,
such as decreasing cAMP production, inactivating kinases, or closing channels.

primary cells cultured from animals lacking individual BBS subunits and, in general, different GPCRs are thought to recruit
proteins [61]. The BBSome complex forms a coat structure that G proteins containing a specific subtype. However, the GPCR
can interact directly with both GPCRs and secretory pathway TGR5 associates with a different Gα isoform in ciliated versus
directors such as Rab8 and facilitate delivery of GPCRs to unciliated cells [63]. Ciliary G proteins can activate different,
the cilium. and sometimes opposing, downstream effector pathways. For
Ciliary GPCRs bind diverse ligands including bile acids, example, once liberated, stimulatory Gα subunits (Gαs) increase
nucleotides, neuropeptides, and proteins [62]. Upon ligand adenylyl cyclase activity while inhibitory Gαi subunits decrease
binding, the receptors facilitate replacement of Gα‐associated cAMP levels. cAMP acts as a second messenger and promotes
GDP with GTP. The trimeric G protein then dissociates from the activity of downstream kinases including protein kinase A
receptor and the Gα and Gβγ subunits both initiate downstream (PKA). Ligand binding leads to retrieval of activated GPCRs
signaling events. There are multiple isoforms of the G protein from the cilium [64–67]. Although additional signaling of
56 THE LIVER:  BUILDING, MAINTAINING, AND DISMANTLING CILIA

internalized GPCRs has been demonstrated in other contexts critical part of signal transduction in olfaction and vision [80,
[68], no activity has been reported for GPCRs internalized 81]. Ciliary ion channels may have multiple effects on signaling
from cilia. cascades. Adenylyl cyclase 3 activity is affected by Ca2+ [82]
Internalization of activated receptors could help reset and it has been hypothesized that changes in intracellular Ca2+ in
signaling pathways. The BBSome, which is implicated in cilia might alter downstream signaling components such as ade-
trafficking proteins to cilia, also facilitates exit of activated nylyl cyclase [83]. In animal models of PKD, tissue levels of
GPCRs through the transition zone in complex with the cAMP are elevated [84]. Activation of ion channels could, if the
small G protein Arl6 [65]. Upon receptor activation, Gαi paradigm is true, exert global effects on any ciliary proteins that
signaling, which reduces adenylyl cyclase and thus PKA are sensitive to changes in ion concentrations. Such a scenario
activity, causes BBSome proteins, Arl6, and retrograde IFT might explain why PKD2‐L1‐deficient mice have reduced
proteins to accumulate at the distal end where the proteins responses to Shh [39].
form large IFT trains, which then bind to activated GPCRs, It is not clear whether cilia act as mechanosensitive flow
transit to the base of the cilium and escort the GPCR through sensors. The prevailing paradigm that primary cilia bend-
the transition zone [69]. ing communicates the presence of external fluid flow via
An alternate mechanism to remove GPCRs and reset sign- mechanosensitive activation of calcium ion channels in the
aling pathways is through direct release of the GPCRs in ciliary membrane has been reconsidered. Detailed studies
cilia‐derived extracellular vesicles (EVs) [70]. Rhodopsin using a fluorescent Ca2+ sensor localized to primary cilia
and the membrane disks of photoreceptor cells are proces- revealed no changes in ciliary Ca2+ levels upon deflection
sively removed from the distal end of the outer segment [71]. with flow from a micropipette [5]. If cilia are mechanosensi-
EVs containing primary cilia proteins have also been col- tive, the signal transduction does not appear to be mediated by
lected in urine [72]. Actin, which is generally not polymer- changes in ciliary calcium. In cartilage, chondrocyte cilia par-
ized in cilia, can drive release of cilia membranes in response ticipate in mechanotransduction, not as a mechanosensor, but
to increased PI(4,5)P2 when INPP5E is disabled or removed as a chemical signal transducer. Receptors in chondrocyte
from cilia [73, 74]. cilia membranes bind ATP that is released by the cell body
Many critical signaling pathways center on primary cilia upon compression [85].
[75]. Obesity generated in ciliopathies is caused by disruptions Although signal reception in cilia is compared to radio signal
in satiety signaling through neuropeptide‐binding GPCRs in reception using antennae, cilia are likely capable of being more
hypothalamus cilia. The hedgehog signaling pathway is critical than just a passive sensor. It is possible that, like other
for developmental and can contribute to cancer. During chronic EVs, cilia‐derived vesicles may be biologically active.
liver disease, hedgehog signaling in liver progenitor cells con- Chlamydomonas releases vesicles from flagella that facilitate
tributes to regeneration [76]. Most components and effectors of breakdown of the wall encapsulating the daughter cells after
this complicated signaling cascade function in cilia. In the cil- mating [86]. In addition to releasing particles, primary cilia may
ium, the activity of the hedgehog receptor Patched influences participate in signaling through direct touch. Cholangiocyte
the ciliary localization of the GPCR Smoothened. The activity cilia within the bile duct and other cilia that project into tubules
of Smoothened in the cilium affects processing of Gli transcrip- can adhere to one another [87], which might facilitate direct
tion factors. Another GPCR, Gpr161, also localizes to the cil- intracellular communication. Primary cilia in many other tissues
ium and mediates hedgehog signaling [77]. are not in a lumen, but rather, can be enmeshed in extracellular
Although it is clear that ligand binding to receptors in the matrix, project between layers of cells, or entwined with other
unique ciliary environment initiates signaling cascades, how extracellular processes (examples include cilia on chondrocytes,
the physiological consequences of these cascades propagate mammary myoepithelial cells, and neurons) [88, 89]. In deep
and cause changes in the rest of the cell is generally less clear. tissue, it is possible that primary cilia could signal through
An exception is the hedgehog signaling pathway: the Gli fam- direct contact.
ily transcription factors localize to primary cilia and can be
processed and trafficked out of the cilium in response to path-
way activation [78, 79]. There are many other possible out-
comes in addition to transcription alterations; however, all of BUILDING, MAINTAINING,
these are subject to the same limitation that Delling and col- AND DISMANTLING CILIA
leagues revealed regarding Ca2+: upon exiting, soluble mole-
cules will be significantly diluted [5]. Delivery of cilia As cells cycle and differentiate, primary cilia are deconstructed
effectors into the nucleus could be streamlined by the overlap- and recreated. When maintained, cilia length can vary by an
ping components of the cilia and nuclear import/export path- order of magnitude between cell types (from ~2 to >20 μm) but
ways. As the centrosome serves as both the basal body and the is typically uniform within a population. The presence and
microtubule organizing center, it is possible that molecules length of a cilium define the ability of a cell to detect and
exiting cilia have a transient local concentration high enough respond to extracellular cues, so the following questions are rel-
to transmit signals to proteins recruited to the centrosome. evant to understanding the role of cilia in health and develop-
Association with carrier molecules or membrane vesicles ment: How do cilia form and how is biogenesis regulated? How
could be an additional strategy. is cilium length determined and maintained? When and how are
cAMP and cGMP generated by signaling cascades can stimu- cilia removed from the cell surface? Partial answers to these
late ion channel activity and alter ciliary Ca2+ levels; this is a questions are discussed below.
5:  Primary Cilia 57

Ciliogenesis cilia proteins, including BBS4 and CEP290, associate with the
key scaffolding protein pericentriolar material‐1 (PCM‐1) in
How is the critical and unique ciliary environment created? centriolar satellites. Impaired autophagy can also prevent cilio-
Primary cilia can grow from a centriole deep in the cytoplasm or genesis because oral–facial–digital syndrome 1 proteins are not
docked at the plasma membrane (Figure  5.4a,b; reviewed in degraded but instead accumulate at centriolar satellites [99].
[90]). In both pathways microtubule doublets extend from the The two centrioles inside each cell are different and only one
triplet microtubules of the centriole, membrane is added, then of them undergoes remodeling to become the basal body. During
presumably differentiated as the transition zone forms. The inter- each cell division cycle, the two centrioles duplicate and then
nal pathway requires the recruitment of vesicles that anchor and separate to form the spindle pole. Each daughter inherits a pair
fuse to form a large ciliary vesicle docked to the mother centri- of centrioles, a duplicate (called the daughter) and an older orig-
ole. As microtubules extend from the centriole, membrane is inal (called the mother). The centriole that becomes the basal
added and the ciliary vesicle deforms to ensheath the nascent body is always the older centriole. Cilia formation can be differ-
cilium. The final stage of internal ciliogenesis requires the mem- ent between sister cells because the two mother centrioles they
brane around the tip of the growing cilium to fuse with the inherit are not equivalent. In the preceding cell cycle one of the
plasma membrane. Ciliogenesis at the plasma membrane is mother centrioles had been a daughter. The sister cell that inher-
thought to begin when a centriole docks at the plasma mem- its the older mother centriole forms a cilium earlier, which could
brane. Membrane is added at the plasma membrane as the micro- influence perception of extracellular developmental cues and
tubules extend to ensheath the cilium. The membrane‐associated downstream differentiation [100]. The older mother centriole
GTPases Rab11 and Rab8 target vesicles to nascent cilia to pro- may be faster because it already has distal appendages and can
vide both lipids and key ciliary membrane proteins [91]. retain association of ciliary membranes [101].
Every centriole is not capable of becoming a basal body. Many polarized cells build cilia at the apical plasma mem-
Structures called distal appendages or transition fibers must be brane and several lines of evidence indicate that polarization
added to the distal end of the ninefold symmetric centriole tri- may position and promote ciliogenesis (reviewed in [102]).
plet microtubules to recruit vesicles or dock at the plasma mem- Proteins that localize to the boundary between apical and baso-
brane. Several components assemble to form distal appendages, lateral membranes in polarized epithelial cells also localize to
and recent super‐resolution imaging efforts have revealed the the base of the primary cilium. Although how polarity proteins
organization of many key distal appendage proteins (Figure 5.2) influence cilia formation and maintenance is unclear, their
[92]. CEP164 and SCLT1 are components of the distinctive importance has been demonstrated. Deletion, mutation, or phar-
blade structure that is visible in electron micrographs. In con- macological disruption of polarity proteins prevents ciliogene-
trast, FBF1 localizes to the distal appendage matrix – a space sis. PAR complex proteins  –  PAR3, PAR6  –  and an atypical
that appears empty in electron micrographs. PKC (aPKC) associate with the anterograde kinesin‐2 compo-
Distal appendage proteins contribute to several aspects of nent Kif3a. The lipid ceramide influences ciliogenesis and is
ciliogenesis. Studies of initiation steps in intracellular ciliogen- required for association of aPKC with other PAR proteins [103].
esis [93] revealed that fusion of small vesicles recruited to distal Cdc42, another PAR complex component, interacts with and
appendages requires membrane tubulating proteins: EHD1 and promotes the ciliary localization of exocyst proteins. The exo-
EHD3. Subsequently, CEP164 in the distal appendage blade cyst complex facilitates targeting and tethering of post‐Golgi
participates in both microtubule growth and membrane expan- vesicles and can interact with Rab8 to direct trafficking in polar-
sion (reviewed in [94]). CEP164 recruits tau tubulin kinase 2 ized cells [104].
(TTBK2), a protein required for a key transformation at the cen- Ciliogenesis at the plasma membrane may be facilitated in
trosome – removal of proteins that cap the plus ends of the cen- some way by the midbody remnant (Figure  5.4b). After the
trosome microtubules. TTBK2 facilitates removal of CP110 cytokinetic bridge is severed during cell division, part of
from the distal end of the centriole. Although it must be removed the bundle of microtubules and membrane that made up the
prior to ciliogenesis, mice with no CP110 have fewer cilia. This bridge can remain associated with the plasma membrane of one
suggests that CP110 may be required for early steps in centro- of the daughter cells. Movement of the remnant across the
some remodeling and then be removed prior to microtubule apical surface of the cell to where the centriole is docked pre-
extension [95]. In the internal ciliogenesis pathway CEP164 cedes cilia growth. Although cilia and midbodies share many
interactions with both Rab8a and its GTPase‐exchange factor molecular components, it is not clear if any of these are directly
Rabin 8 are thought to promote membrane docking of early cili- transferred to contribute to cilia growth [105].
ary membranes [96].
Targeting and transport of ciliary components is also essen-
tial for cilia formation. Mutation or deletion of IFT‐B complex
Maintaining or altering cilia length
proteins, which facilitates anterograde transport, disrupts cilia Cilia length is typically constant across a population but can
formation. FBF1 in the distal appendage matrix colocalizes with vary across cell types. This suggests that cellular parameters
IFT88 and facilitates entry of IFT complexes into cilia [92, 97]. such as protein expression levels or turnover rates influence the
Members of the IFT‐B complex bind tubulin and may escort it cilia length setpoint. Several factors that alter the cilia length
through the permeability barrier where it is concentrated in the setpoint have been identified. For example, tubulin concentra-
cilium [28, 98]. Interference with the formation of centriolar tions and modifications can affect cilia length [28, 106, 107]. In
satellites, which are thought to function as a protein complex addition, proteins involved in IFT are known to affect cilia
assembly platform, can also prevent cilia formation. Several length: increased expression of IFT‐B components yields longer
58 THE LIVER:  BUILDING, MAINTAINING, AND DISMANTLING CILIA

(a)

Ciliary
Distal vesicle
appendages

Centriole

(b)

Midbody
remnant

Transition
? zone

(c)

Dissasembly Severing Resorption

Figure 5.4  Ciliogenesis and cilia elimination. (a) Internal ciliogenesis begins with maturation of the mother centriole. Although they are depicted
as discrete stages, initial vesicle recruitment and axoneme extension may occur simultaneously. As the axoneme extends, the transition zone forms
and additional membrane must be delivered. When the membrane encapsulating the cilium contacts the plasma membrane, the bilayers are thought
to fuse so the membrane adjacent to the microtubules becomes continuous with the plasma membrane. The outer membrane of the ciliary vesicle
may become part of the ciliary pocket membrane. (b) Ciliogenesis at the plasma membrane requires formation of distal appendages and movement
of the mother centriole to the cell surface. It is not currently clear how the midbody remnant promotes cilia assembly at the cell surface. Like in the
internal pathway, the transition zone forms and microtubule growth must be coupled to membrane addition to extend the nascent cilium. (c) Three
­possibly strategies that could all contribute to cilium disappearance are depicted. Microtubule disassembly may be stimulated by activation of
the tubulin deacetylase HDAC6. Membrane removal via endocytosis could also contribute. Severing of ciliary membrane is known to occur and has
been associated with cilium removal. A third strategy involves dissolving the barriers that separate the ciliary cytoplasm and membrane from the
rest of the cell. There has been some evidence to support each possibility.

cilia [108]. IFT‐B delivery of tubulin to the tip may increase the four positive regulators of branched F‐actin [112]. Cilia elongation
local concentration of tubulin and drive polymerization. upon treatment with cytochalasin D is accompanied by an influx
Kinesin‐2 diffuses back to the base of the cilium after it carries of membrane‐associated actin‐binding proteins in cilia [113].
IFT trains to the tip. Because diffusion of kinesin‐2 back to the Cdc42 and aPKC, components of the PAR complex, participate
base of the cilium where it can form new anterograde trains is in signaling to influence F‐actin [114]. How disrupting branched
length dependent, kinesin‐2 concentration has been proposed as F‐actin affects cilia length is not yet clear. Perhaps the actin
a key regulator of cilia length [109, 110]. motor myosin Va, which can bind Rab8 and a transition zone
The actin cytoskeleton can also influence cilia length [111]. protein [115], participates in actin regulation of cilia length.
Cilia length increases in cells when F‐actin is destabilized by Several signaling pathways and disease conditions alter cilia
treatment with cytochalasin D. Similar increases in cilia length length. Inhibition of adenylyl cyclase by treatment with lithium
have been reported by depleting Arp3, which nucleates F‐actin causes cilia elongation [116]. In physiological contexts, adeny-
branching and by treatment with a microRNA that downregulates lyl cyclase activity can be stimulated or inhibited by Gα proteins
5:  Primary Cilia 59

upon activation of ciliary GPCRs. It is not yet clear whether REFERENCES


cAMP as a second messenger directly influences a length regu-
lating pathways described above, or if it promotes cilia exten- 1. Reiter, J.F. and Leroux, M.R. Genes and molecular pathways underpinning
sion through other binding partners. Another candidate ciliopathies. Nat Rev Mol Cell Biol, 2017;18(9):533–47.
2. Battle, C., Ott, C.M., Burnette, D.T. et al. Intracellular and extracellular forces
downstream of signaling pathways that may also influence cilia
drive primary cilia movement. Proc Natl Acad Sci U S A, 2015;112(5): 1410–15.
length is the proteasome, which associates with a transition 3. Yeh, C., Li, A., Chuang, J.‐Z. et al. IGF‐1 activates a cilium‐localized nonca-
zone protein and cleaves Gli transcription factors [117]. nonical Gβγ signaling pathway that regulates cell‐cycle progression. Dev Cell,
2013;26(4):358–68.
4. Li, Q., Montalbetti, N., Wu, Y. et al. Polycystin‐2 cation channel function is
Cilia elimination under the control of microtubular structures in primary cilia of renal epithe-
lial cells. J Biol Chem, 2006;281(49):37566–75.
While some organisms retain cilia or flagella, mammalian cells 5. Delling, M., Indzhykulian, A.A., Liu, X. et al. Primary cilia are not calcium‐
disassemble cilia in G1 phase of the cell cycle or at the begin- responsive mechanosensors. Nature, 2016;531(7596):656–60.
ning of mitosis. Cilia removal could be achieved by: (i) sever- 6. Nachury, M.V. How do cilia organize signalling cascades? Philos Trans R
ing; (ii) disassembly; and (iii) resorption upon decommissioning Soc Lond B Biol Sci, 2014;369(1650):20130465.
7. Wloga, D., Joachimiak, E., Louka, P., and Gaertig, J. Posttranslational modi-
the boundaries that segregate the cilium (Figure 5.4c). The three fications of tubulin and cilia. Cold Spring Harb Perspect Biol,
mechanisms could work together and there is evidence to sug- 2017;9(6):a028159.
gest that all three processes can happen [93]. Electron micros- 8. Gluenz, E., Höög, J.L., Smith, A.E. et al. Beyond 9+0: noncanonical axo-
copy of epithelial cells provided early support for the model that neme structures characterize sensory cilia from protists to humans. FASEB J,
2010;24(9):3117–21.
cilia are resorbed into the cytoplasm [118]. Stimulated release
9. Lechtreck, K.F. IFT–cargo interactions and protein transport in cilia. Trends
of vesicles from the tip can precipitate cilia disassembly [74]. Biochem Sci, 2015;40(12):765–78.
Aurora A kinase, which is regulated by many factors and func- 10. Kozminski, K.G., Johnson, K.A., Forscher, P., and Rosenbaum, J.L. A motil-
tions as a part of the cell cycle, phosphorylates both INPP5E ity in the eukaryotic flagellum unrelated to flagellar beating. Proc Natl Acad
and the deacetylase HDAC6. HDAC6 destabilization of acety- Sci U S A, 1993;90(12):5519–23.
11. Cole, D.G., Diener, D.R., Himelblau, A.L. et al. Chlamydomonas kinesin‐II‐
lated microtubules may facilitate axoneme deconstruction. The
dependent intraflagellar transport (IFT): IFT particles contain proteins
activity of microtubule‐depolymerizing kinesins associated required for ciliary assembly in Caenorhabditis elegans sensory neurons. J
with the axoneme may also contribute to reducing axoneme Cell Biol, 1998;141(4):993–1008.
length. It is not clear if there is a role for endocytosis in recover- 12. Stepanek, L. and Pigino, G. Microtubule doublets are double‐track railways
ing membrane from cilia. Polo‐like kinase 1, a kinase regulated for intraflagellar transport trains. Science, 2016;352(6286):721–4.
13. Benmerah, A. The ciliary pocket. Curr Opin Cell Biol, 2012;25(1):78–84.
by the cell cycle, promotes cilia disassembly and phosphoryl- 14. Gonçalves, J. and Pelletier, L. The ciliary transition zone: finding the pieces
ates a component of the transition zone, possibly disrupting the and assembling the gate. Mol Cells, 2017;40(4):243–53.
integrity of the ciliary gate (see [119]). 15. Jana, S.C., Mendonça, S., Machado, P. et al. Differential regulation of transi-
tion zone and centriole proteins contributes to ciliary base diversity. Nat Cell
Biol, 2018;20(8):928–41.
16. Sang, L., Miller, J.J., Corbit, K.C. et  al. Mapping the NPHP‐JBTS‐MKS
protein network reveals ciliopathy disease genes and pathways. Cell,
CONCLUSION 2011;145(4):513–28.
17. Dean, S., Moreira‐Leite, F., Varga, V., and Gull, K. Cilium transition zone
Because of their distinct structure in electron microscopy, primary proteome reveals compartmentalization and differential dynamics of ciliopa-
cilia were identified on many cell types throughout the twentieth thy complexes. Proc Natl Acad Sci U S A, 2016;113(35):E5135–43.
century. A small community of researchers built foundational 18. Diener, D.R., Lupetti, P., and Rosenbaum, J.L. Proteomic analysis of isolated
ciliary transition zones reveals the presence of ESCRT oroteins. Curr Biol,
knowledge about the structure and function of primary cilia. The 2015;25(3):379–84.
discovery that proteins move along cilia, followed by insights into 19. Yang, T.T., Su, J., Wang, W.‐J. et  al. Superresolution pattern recognition
the many links between primary cilia function and health stimu- reveals the architectural map of the ciliary transition zone. Sci Rep,
lated expansion of cilia research. Investigators in disparate fields 2015;5(1):14096.
20. Williams, C.L., Li, C., Kida, K. et al. MKS and NPHP modules cooperate to
have found that cilia biology is central to many aspects of cell
establish basal body/transition zone membrane associations and ciliary gate
signaling and development. These, in turn, have provided new function during ciliogenesis. J Cell Biol, 2011;192(6):1023–41.
insights into fundamental aspects of cilia biology. Technological 21. Craige, B., Tsao, C.‐C., Diener, D.R. et al. CEP290 tethers flagellar transi-
advances, including super‐resolution microscopy, have provided tion zone microtubules to the membrane and regulates flagellar protein con-
new insights into the movement and molecular architecture of tent. J Cell Biol, 2010;190(5):927–40.
22. Awata, J., Takada, S., Standley, C. et al. NPHP4 controls ciliary trafficking
cilia. However, many exciting questions remain. Emerging studies
of membrane proteins and large soluble proteins at the transition zone. J Cell
will likely provide additional insights into how defects in cilia Sci, 2014;127(Pt 21):4714–27.
form and function contribute to disease and will hopefully lead to 23. Lambacher, N.J., Bruel, A.‐L., van Dam, T.J.P. et al. TMEM107 recruits cili-
novel interventions to alleviate the symptoms of ciliopathies. opathy proteins to subdomains of the ciliary transition zone and causes
Joubert syndrome. Nat Cell Biol, 2016;18(1):122–31.
24. Shi, X., Garcia, G., III, Van De Weghe, J.C. et al. Super‐resolution micros-
copy reveals that disruption of ciliary transition‐zone architecture causes
Joubert syndrome. Nat Cell Biol, 2017;19(10):1178–88.
ACKNOWLEDGMENT 25. Yang, T.T., Hampilos, P.J., Nathwani, B. et  al. Superresolution STED
microscopy reveals differential localization in primary cilia. Cytoskeleton
(Hoboken), 2012;70(1):54–65.
The author thanks Christine Kettenhofen, Corey Valinsky, and 26. Takao, D. and Verhey, K.J. Gated entry into the ciliary compartment. Cell
Shu‐Hsien Sheu for helpful suggestions on the text. Mol Life Sci, 2016;73(1):119–27.
60 THE LIVER: REFERENCES

27. Kee, H.L., Dishinger, J.F., Blasius, T.L. et al. A size‐exclusion permeability 53. Berbari, N.F., Johnson, A.D., Lewis, J.S. et  al. Identification of ciliary
barrier and nucleoporins characterize a ciliary pore complex that regulates localization sequences within the third intracellular loop of G protein‐cou-
transport into cilia. Nat Cell Biol, 2012;14(4):431–7. pled receptors. Mol Biol Cell, 2008;19(4):1540–7.
28. Endicott, S.J. and Brueckner, M. NUP98 Sets the size‐exclusion diffusion 54. Loktev, A.V. and Jackson, P.K. Neuropeptide Y family receptors traffic via
limit through the ciliary base. Curr Biol, 2018;28(10):1643–50.e3. the Bardet‐Biedl syndrome pathway to signal in neuronal primary cilia. Cell
29. Dishinger, J.F., Kee, H.L., Jenkins, P.M. et al. Ciliary entry of the kinesin‐2 Rep, 2013;5(5):1316–29.
motor KIF17 is regulated by importin‐β2 and RanGTP. Nat Cell Biol, 55. Corbit, K.C., Aanstad, P., Singla, V. et al. Vertebrate Smoothened functions
2010;12(7):703–10. at the primary cilium. Nature, 2005;437(7061):1018–21.
30. Souto‐Padron, T. and de Souza, W. Freeze‐fracture localization of filipin‐ 56. Badgandi, H.B., Hwang, S.‐H., Shimada, I.S. et al. Tubby family proteins are
cholesterol complexes in the plasma membrane of Trypanosoma cruzi. J adapters for ciliary trafficking of integral membrane proteins. J Cell Biol,
Parasitol, 1983;69(1):129. 2017;216(3):743–60.
31. Cuevas, P. and Gutierrez Diaz, J.A. Absence of filipin‐sterol complexes from 57. Sun, X., Haley, J., Bulgakov, O.V. et al. Tubby is required for trafficking G
the ciliary necklace of ependymal cells. Anat Embryol, 1985;172(1). protein‐coupled receptors to neuronal cilia. Cilia, 2012;1(1):21.
32. Vieira, O.V., Gaus, K., Verkade, P. et al. FAPP2, cilium formation, and com- 58. Lu, L. and Madugula, V. Mechanisms of ciliary targeting: entering importins
partmentalization of the apical membrane in polarized Madin‐Darby canine and Rabs. Cell Mol Life Sci, 2017;75(4):597–606.
kidney (MDCK) cells. Proc Natl Acad Sci U S A, 2006;103(49):18556–61. 59. Nachury, M.V., Loktev, A.V., Zhang, Q. et al. A core complex of BBS pro-
33. Garcia‐Gonzalo, F.R., Phua, S.C., Roberson, E.C. et al. Phosphoinositides teins cooperates with the GTPase Rab8 to promote ciliary membrane biogen-
regulate ciliary protein trafficking to modulate hedgehog signaling. Dev Cell, esis. Cell, 2007;129(6):1201–13.
2015;34(4):400–9. 60. Khan, S.A., Muhammad, N., Khan, M.A. et al. Genetics of human Bardet–
34. Chávez, M., Ena, S., Van Sande, J. et  al. Modulation of ciliary phosphoi- Biedl syndrome, an updates. Clin Genet, 2016;90(1):3–15.
nositide content regulates trafficking and sonic hedgehog signaling output. 61. Berbari, N.F., Lewis, J.S., Bishop, G.A. et al. Bardet‐Biedl syndrome pro-
Dev Cell, 2015;34(3):338–50. teins are required for the localization of G protein‐coupled receptors to pri-
35. Bielas, S.L., Silhavy, J.L., Brancati, F. et al. Mutations in INPP5E, encoding mary cilia. Proc Natl Acad Sci U S A, 2008;105(11):4242–6.
inositol polyphosphate‐5‐phosphatase E, link phosphatidyl inositol signaling 62. Mykytyn, K. and Askwith, C. G‐Protein‐coupled receptor signaling in cilia.
to the ciliopathies. Nat Genet, 2009;41(9):1032–6. Cold Spring Harb Perspect Biol, 2017;9(9):a028183.
36. Jacoby, M., Cox, J.J., Gayral, S. et al. INPP5E mutations cause primary cil- 63. Masyuk, A.I., Huang, B.Q., Radtke, B.N. et al. Ciliary subcellular localiza-
ium signaling defects, ciliary instability and ciliopathies in human and tion of TGR5 determines the cholangiocyte functional response to bile acid
mouse. Nat Genet, 2009;41(9):1027–31. signaling. Am. J Physiol Gastrointest Liver Physiol,
37. Fansa, E.K., Kösling, S.K., Zent, E. et  al. PDE6δ‐mediated sorting of 2013;304(11):G1013–24.
INPP5E into the cilium is determined by cargo‐carrier affinity. Nat Commun, 64. Green, J.A., Schmid, C.L., Bley, E. et al. Recruitment of β‐arrestin into neu-
2016;7:11366. ronal cilia modulates somatostatin receptor subtype 3 ciliary localization.
38. Jensen, V.L., Li, C., Bowie, R.V. et al. Formation of the transition zone by Mol Cell Biol, 2016;36(1):223–35.
Mks5/Rpgrip1L establishes a ciliary zone of exclusion (CIZE) that compart- 65. Ye, F., Nager, A.R., and Nachury, M.V. BBSome trains remove activated
mentalises ciliary signalling proteins and controls PIP2 ciliary abundance. GPCRs from cilia by enabling passage through the transition zone. J Cell
EMBO J, 2015;34(20):2537–56. Biol. 2018;217(5):1847–68.
39. Delling, M., DeCaen, P.G., Doerner, J.F. et al. Primary cilia are specialized 66. Pal, K., Hwang, S.‐H., Somatilaka, B. et al. Smoothened determines β‐arres-
calcium signalling organelles. Nature, 2013;504(7479):311–14. tin–mediated removal of the G protein–coupled receptor Gpr161 from the
40. Shen, P.S., Yang, X., DeCaen, P.G. et al. The structure of the polycystic kid- primary cilium. J Cell Biol, 2016;212(7):861–75.
ney disease channel PKD2 in lipid nanodiscs. Cell, 2016;167(3):763–73. 67. Mashukova, A., Spehr, M., Hatt, H., and Neuhaus, E.M. β‐Arrestin2‐medi-
e11. ated internalization of mammalian odorant receptors. J Neurosci,
41. Liu, X., Vien, T., Duan, J. et al. Polycystin‐2 is an essential ion channel subu- 2006;26(39):9902–12.
nit in the primary cilium of the renal collecting duct epithelium. Elife, 68. Luttrell, L.M., Ferguson, S.S., Daaka, Y. et al. Beta‐arrestin‐dependent for-
2018;7:E2363. mation of beta2 adrenergic receptor‐Src protein kinase complexes. Science,
42. Moore, B.S., Stepanchick, A.N., Tewson, P.H. et al. Cilia have high cAMP 1999;283(5402):655–61.
levels that are inhibited by Sonic Hedgehog‐regulated calcium dynamics. 69. Ye, F., Nager, A.R., and Nachury, M.V. BBSome trains remove activated
Proc Natl Acad Sci U S A, 2016;113(46):13069–74. GPCRs from cilia by enabling passage through the transition zone. J Cell
43. Sigg, M.A., Menchen, T., Lee, C. et al. Evolutionary proteomics uncovers Biol, 2018;217(5):1847–68.
ancient associations of cilia with signaling pathways. Dev Cell, 70. Wood, C.R. and Rosenbaum, J.L. Ciliary ectosomes: transmissions from the
2017;43(6):744–62.e11. cell’s antenna. Trends Cell Biol, 2015;25 (5):276–85.
44. Singla, V. and Reiter, J.F. The primary cilium as the cell’s antenna: signaling 71. LaVail, M.M. Rod outer segment disk shedding in rat retina: relationship to
at a sensory organelle. Science, 2006;313(5787):629–33. cyclic lighting. Science, 1976;194(4269):1071–4.
45. Sipe, C.W. and Lu, X. Kif3a regulates planar polarization of auditory hair 72. Hogan, M.C., Manganelli, L., Woollard, J.R. et al. Characterization of PKD
cells through both ciliary and non‐ciliary mechanisms. Development, protein‐positive exosome‐like vesicles. J Am Soc Nephrol,
2011;138(16):3441–9. 2009;20(2):278–88.
46. Huangfu, D., Liu, A., Rakeman, A.S. et al. Hedgehog signalling in the mouse 73. Nager, A.R., Goldstein, J.S., Herranz‐Pérez, V. et al. An actin network dis-
requires intraflagellar transport proteins. Nature, 2003;426(6962):83–7. patches ciliary GPCRs into extracellular vesicles to modulate signaling. Cell,
47. Nonaka, S., Tanaka, Y., Okada, Y. et al. Randomization of left‐right asym- 2017;168(1–2):252–63.e14.
metry due to loss of nodal cilia generating leftward flow of extraembryonic 74. Phua, S.C., Chiba, S., Suzuki, M. et al. Dynamic remodeling of membrane
fluid in mice lacking KIF3B motor protein. Cell, 1998;95(6):829–37. composition drives cell cycle through primary cilia excision. Cell,
48. Davenport, J.R., Watts, A.J., Roper, V.C. et al. Disruption of intraflagellar 2017;168(1–2):264–79.e15.
transport in adult mice leads to obesity and slow‐onset cystic kidney disease. 75. Hilgendorf, K.I., Johnson, C.T., and Jackson, P.K. The primary cilium as a
Curr Biol, 2007;17 (18):1586–94. cellular receiver: organizing ciliary GPCR signaling. Curr Opin Cell Biol,
49. Haycraft, C.J., Zhang, Q., Song, B. et al. Intraflagellar transport is essential 2016;39:84–92.
for endochondral bone formation. Development, 2007;134(2):307–16. 76. Grzelak, C.A., Martelotto, L.G., Sigglekow, N.D. et al. The intrahepatic sig-
50. Dinsmore, C. and Reiter, J.F. Endothelial primary cilia inhibit atherosclero- nalling niche of hedgehog is defined by primary cilia positive cells during
sis. EMBO Rep, 2016;17(2):156–66. chronic liver injury. J Hepatol, 2014;60(1):143–51.
51. Foerster, P., Daclin, M., Asm, S. et al. mTORC1 signaling and primary cilia 77. Mukhopadhyay, S. and Rohatgi, R. G‐protein‐coupled receptors, Hedgehog
are required for brain ventricle morphogenesis. Development, signaling and primary cilia. Semin Cell Dev Biol, 2014;33:63–72.
2017;144(2):201–10. 78. Haycraft, C.J., Banizs, B., Aydin‐Son, Y. et al. Gli2 and Gli3 localize to cilia
52. Liu, H., Kiseleva, A.A., and Golemis, E.A. Ciliary signalling in cancer. Nat and require the intraflagellar transport protein polaris for processing and
Rev Cancer, 2018;18(8):511–24. function. PLoS Genet, 2005;1(4):e53.
5:  Primary Cilia 61

79. Liu, J., Zeng, H., and Liu, A. The loss of Hh responsiveness by a non‐ciliary 99. Park, S.M., Lim, J.S., Ramakrishina, S. et al. Brain somatic mutations in
Gli2 variant. Development, 2015;142(9):1651–60. MTOR disrupt neuronal ciliogenesis, leading to focal cortical dyslamina-
80. Nakamura, T. and Gold, G.H. A cyclic nucleotide‐gated conductance in tion. Neuron, 2018;99(1):83–97.e7.
olfactory receptor cilia. Nature, 1987;325(6103):442–4. 100. Anderson, C.T. and Stearns, T. Centriole age underlies asynchronous primary
81. Fesenko, E.E., Kolesnikov, S.S., and Lyubarsky, A.L. Induction by cyclic cilium growth in mammalian cells. Curr Biol, 2009;19(17):1498–502.
GMP of cationic conductance in plasma membrane of retinal rod outer 101. Paridaen, J.T.M.L., Wilsch‐Bräuninger, M., and Huttner, W.B. Asymmetric
­segment. Nature, 1985;313(6000):310–13. inheritance of centrosome‐associated primary cilium membrane directs
82. Choi, E.J., Xia, Z., and Storm, D.R. Stimulation of the type III olfactory ciliogenesis after cell division. Cell, 2013;155(2):333–44.
adenylyl cyclase by calcium and calmodulin. Biochemistry, 1992;31(28): 102. Bernabé‐Rubio, M. and Alonso, M.A. Routes and machinery of primary
6492–8. cilium biogenesis. Cell Mol Life Sci, 2017;74 (22):4077–95.
83. Pablo, J.L., DeCaen, P.G., and Clapham, D.E. Progress in ciliary ion channel 103. He, Q., Wang, G., Dasgupta, S. et al. Characterization of an apical cera-
physiology. J Gen Physiol, 2017;149(1):37–47. mide‐enriched compartment regulating ciliogenesis. Mol Biol Cell, 2012;23(16):
84. Torres, V.E. and Harris, P.C. Strategies targeting cAMP signaling in the 3156–66.
treatment of polycystic kidney disease. J Am Soc Nephrol, 2014; 104. Wu, B. and Guo, W. The exocyst at a glance. J Cell Sci, 2015;
25(1):18–32. 128(16):2957–64.
85. Wann, A.K.T., Zuo, N., Haycraft, C.J. et al. Primary cilia mediate mecha- 105. Bernabé‐Rubio, M., Andrés, G., Casares‐Arias, J. et al. Novel role for the
notransduction through control of ATP‐induced Ca2+ signaling in compressed midbody in primary ciliogenesis by polarized epithelial cells. J Cell Biol,
chondrocytes. FASEB J, 2012;26(4):1663–71. 2016;214(3):259–73.
86. Wood, C.R., Huang, K., Diener, D.R., and Rosenbaum, J.L. The cilium 106. Sharma, N., Kosan, Z.A., Stallworth, J.E., Berbari, N.F., and Yoder, B.K.
secretes bioactive ectosomes. Curr Biol, 2013;23 (10):906–11. Soluble levels of cytosolic tubulin regulate ciliary length control. Mol Biol
87. Ott, C., Elia, N., Jeong, S.Y. et al. Primary cilia utilize glycoprotein‐depend- Cell, 2011;22(6):806–16.
ent adhesion mechanisms to stabilize long‐lasting cilia‐cilia contacts. Cilia, 107. Gadadhar, S., Dadi, H., Bodakuntla, S. et al. Tubulin glycylation controls
2012;1(1):3. primary cilia length. J Cell Biol, 2017;216(9):2701–13.
88. McDermott, K.M., Liu, B.Y., Tlsty, T.D., and Pazour, G.J. Primary cilia 108. Kim, S. and Dynlacht, B.D. Assembling a primary cilium. Curr Opin Cell
regulate branching morphogenesis during mammary gland development. Biol, 2013;25(4):506–11.
Curr Biol, 2010;20(8):731–7. 109. Chien, A., Shih, S.M., Bower, R. et al. Dynamics of the IFT machinery at
89. Jensen, C.G., Poole, C.A., McGlashan, S.R. et  al. Ultrastructural, tomo- the ciliary tip. Elife, 2017;6:979.
graphic and confocal imaging of the chondrocyte primary cilium in situ. Cell 110. Hendel, N.L., Thomson, M., and Marshall, W.F. Diffusion as a ruler: mod-
Biol Int, 2004;28(2):101–10. eling kinesin diffusion as a length sensor for intraflagellar transport.
90. Sorokin, S.P. Reconstructions of centriole formation and ciliogenesis in Biophys J, 2018;114(3):663–74.
mammalian lungs. J Cell Sci, 1968;3(2):207–30. 111. Mirvis, M., Stearns, T., and Nelson, W.J. Cilium structure, assembly, and disas-
91. Lu, L. and Madugula, V. Mechanisms of ciliary targeting: entering importins sembly regulated by the cytoskeleton. Biochem J, 2018;475(14):2329–53.
and Rabs. Cell Mol Life Sci, 2017;75(4):597–606. 112. Yan, X. and Zhu, X. Branched F‐actin as a negative regulator of cilia formation.
92. Yang, T.T., Chong, W.M., Wang, W.‐J. et al. Super‐resolution architecture of Exp Cell Res, 2013;319(2):147–51.
mammalian centriole distal appendages reveals distinct blade and matrix 113. Kohli, P., Höhne, M., Jüngst, C. et  al. The ciliary membrane‐associated
functional components. Nat Commun, 2018;9(1):2023. proteome reveals actin‐binding proteins as key components of cilia. EMBO
93. Sánchez, I. and Dynlacht, B.D. Cilium assembly and disassembly. Nat Cell Rep, 2017;18(9):1521–35.
Biol, 2016;18(7):711–17. 114. Drummond, M.L., Li, M., Tarapore, E. et al. Actin polymerization controls
94. Garcia‐Gonzalo, F.R. and Reiter, J.F. Open sesame: how transition fibers and cilia‐mediated signaling. J Cell Biol, 2018;217(9):3255–66.
the transition zone control ciliary composition. Cold Spring Harb Perspect 115. Assis, L.H.P., Silva‐Junior, R.M.P., Dolce, L.G. et al. The molecular motor
Biol, 2017;9(2):a028134. Myosin Va interacts with the cilia‐centrosomal protein RPGRIP1L. Sci
95. Yadav, S.P., Sharma, N.K., Liu, C. et al. Centrosomal protein CP110 controls Rep, 2017;7(1):43692.
maturation of mother centriole during cilia biogenesis. Development, 116. Ou, Y., Ruan, Y., Cheng, M. et al. Adenylate cyclase regulates elongation of
2016;143(9):1491–501. mammalian primary cilia. Exp Cell Res, 2009;315(16):2802–17.
96. Schmidt, K.N., Kuhns, S., Neuner, A. et al. Cep164 mediates vesicular dock- 117. Gerhardt, C., Leu, T., Lier, J.M., and Rüther, U. The cilia‐regulated protea-
ing to the mother centriole during early steps of ciliogenesis. J Cell Biol., some and its role in the development of ciliopathies and cancer. Cilia,
2012;199(7):1083–101. 2016;5(1):14.
97. Wei, Q., Xu, Q., Zhang, Y. et al. Transition fibre protein FBF1 is required for 118. Rieder, C.L., Jensen, C.G., and Jensen, L.C. The resorption of primary cilia
the ciliary entry of assembled intraflagellar transport complexes. Nat during mitosis in a vertebrate (PtK1) cell line. J Ultrastruct Res,
Commun, 2013;4(1):2750. 1979;68(2):173–85.
98. Kubo, T., Brown, Jason M., Bellve, Karl et  al. Together, the IFT81 and 119. Seeger‐Nukpezah, T., Liebau, M.C., Höpker, K. et  al. The centrosomal
IFT74 N‐termini form the main module for intraflagellar transport of tubulin. kinase plk1 localizes to the transition zone of primary cilia and induces
J Cell Sci, 2016;129(10):2106–19. phosphorylation of nephrocystin‐1. PLoS One, 2012;7(6):e38838.
Endocytosis in Liver Function
6 and Pathology
Micah B. Schott, Barbara Schroeder, and Mark A. McNiven
Department of Biochemistry and Molecular Biology, Division of Gastroenterology and Hepatology, Mayo Clinic,
Rochester, MN, USA

INTRODUCTION extracellular fluid, whereas selective pathways, such as recep-


tor‐mediated endocytosis, import specific soluble ligands and
A prominent function of the hepatocyte is the regulated uptake transmembrane receptors. Regardless of the mechanism of
of extracellular material for subsequent processing and/or trans- internalization, endocytic vesicles deliver cargo to the early
­
port into bile. This process, known as endocytosis, depends on endosome (EE) for sorting to different destinations, including
elaborate vesicle trafficking machinery that is linked to specific recycling back to the cell surface, secretion to the extracellular
lipid–membrane subdomains and the cytoskeletal matrix. This milieu, or degradation by late endosomes and lysosomes.
provides a mechanism to sequester and internalize transmem- During receptor‐mediated endocytosis, extracellular ligands,
brane receptor/ligand complexes, such as epidermal growth fac- such as LDL, transferrin, epidermal growth factor (EGF), hor-
tor, hepatocyte growth factor, and iron‐bound transferrin, and to mones, and many others, bind with high affinity to specific
help maintain normal lipid serum levels through the endocytosis receptors at the plasma membrane to propagate intracellular
of low‐density lipoproteins (LDLs). Of equal importance is the signaling cascades that lead to gene transcription and other
fact that this highly evolved machinery can be “hijacked” by processes. In addition to signal transduction, receptor–ligand
many pathogens including bacteria, viruses, and parasites to binding also initiates endocytic uptake, an event that “desensi-
infect the liver, leading to inflammation and hepatitis. This tizes” signaling by reducing receptor availability at the cell
review will outline the molecular and cell biological mecha- surface. More recently, insights into “housekeeping” trophic
nisms that underlie endocytosis in the liver, including the diverse receptors, such as the transferrin receptor (TfR) and the LDL
roles of endocytic vesicles in the cytoplasm and how these path- receptor (LDLR), which have been assumed not to play a
ways are altered in liver disease. prominent role in cell signaling, have been observed to activate
specific signaling cascades.
The formation of endocytic vesicles requires the coordination
of a wide variety of adaptor proteins that link cargo to the
ENDOCYTIC VESICLE FORMATION ­endocytic machinery and cytoskeleton. As a brief summary, the
AT THE PLASMA MEMBRANE various modes of endocytosis and their molecular machinery
are described in the following section.
Endocytic vesicles are formed at the plasma membrane as inward
budding events that invaginate toward the cytoplasm. These
membrane structures form vesicles that are packed with various
types of “cargo” such as integral membrane proteins, receptor– CLATHRIN‐DEPENDENT ENDOCYTOSIS
ligand complexes, lipids, fluid, and nutrients. The cargo content
of endocytic vesicles differs considerably between the various Clathrin‐dependent endocytosis is a central and intensely studied
modes of uptake. For example, nonselective endocytic pathways, process by which most surface receptors are internalized. The initial
such as macropinocytosis, mediate the import of nutrient‐rich steps of this event require receptor–ligand interactions at the plasma

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
6:  Endocytosis in Liver Function and Pathology 63

membrane, which trigger the recruitment of adaptor proteins that and temporal coordination of over 50 different adaptors [1].
orchestrate the assembly of a clathrin coat (Figure 6.1). A model of these events is depicted in Figure 6.1.
Clathrin‐coated vesicles are relatively uniform in size (100–
150 nm diameter) and are present in all cell types. They were first Coated pit/vesicle formation
observed by Roth and Porter, who noted that the endocytosis of
and associated factors
yolk protein into the oocyte of the mosquito was associated with
a marked increase in invaginations of the oocyte cell membrane, The plasma membrane is markedly enriched in the phospholipid
which was coated on the cytoplasmic face with what they termed phosphatidylinositol 4,5‐bisphosphate (PI(4,5)P2) compared to
a bristle coat [3]. Soon thereafter, clathrin was identified as the other cellular compartments, which aids in recruitment of adap-
major protein component of the coat [4]. tors during the initial stages of vesicle formation. The mecha-
Clathrin‐dependent endocytosis is a highly coordinated pro- nisms that initiate receptor sequestration and membrane bending
cess that remains a topic of great research interest. First, plasma at these sites are not clearly understood. At minimum, these
membrane subdomains undergo phospholipid remodeling at early steps seem to require the PI(4,5)P2‐binding protein Fes/
sites of vesicle formation, a step that is critical for the recruit- CIP4 homology (FCH) domain only 1 and 2 (FCHO1 and
ment of adaptors that link surface receptors with clathrin and FCHO2), as well as the adaptors epidermal growth factor recep-
other endocytic proteins. Next, inward membrane curvature is tor (EGFR) pathway substrate 15 (Eps15) and intersectins 1 and
achieved by the actin cytoskeleton and curvature‐sensing BAR 2 [1]. FCHO proteins also contain curvature‐sensing BAR
domain proteins. Clathrin is recruited from the cytosol to stabi- domains that likely aid in their localization at early budding
lize the forming vesicle and aid in its displacement from the vesicles. Eps15, along with other cargo‐specific adaptors, helps
plasma membrane. Forming vesicles are separated from the cell to recruit a critical endocytosis protein known as adaptin 2
surface by dynamin oligomers that induce vesicle scission by (AP2), the main link between surface receptors and clathrin.
generating constrictive force around the thin vesicle neck. Once The role of receptor–ligand interactions in signaling clathrin‐
internalized, the clathrin coat is disassembled to allow for vesi- coated pit formation is unclear, but some studies suggest that
cle fusion with downstream target endosomes. All of these steps receptors may directly recruit a specific subset of endocytic
occur rapidly on the order of 1–5 minutes and require spatial adaptors. For example, transferrin receptor contains a tyrosine

(a) Adaptor, cargo


recruitment Membrane bending Scission Uncoating

(b) (c) BAR-domain proteins


(endophilin, amphiphisin)
Synd
apin Cortactin
Actin
Dynamin
Arp2/3

AP2
Clathrin

GPCR RTK
AP2

Arrestin Clathrin

Figure 6.1  Model of clathrin‐mediated endocytosis. (a) Endocytosis occurs at plasma membrane sites of receptor–ligand cargo where clathrin and
clathrin adaptors are recruited. Invaginating membranes are separated from the plasma membrane by scission machinery in cooperation with the
cytoskeleton. Vesicle uncoating allows for participation in downstream endocytic trafficking routes. Cartoon adapted from [1] with permission of
Springer Nature. (b) Scanning electron micrograph of clathrin‐coated vesicles show an intracellular view of clathrin‐coated pits. Reprinted from [2]
with permission of Rockefeller University Press. (c) Components of clathrin‐coated vesicles, including fission machinery and associated actin
cytoskeleton. These components come together to complete the process of cargo sequestration, vesicle formation, and membrane scission.
64 THE LIVER: HEPATOCELLUAR TRAFFICKING OF ENDOCYTIC VESICLES

recognition motif that activates AP2 to recruit clathrin and that  nonselectively imports extracellular fluid and nutrients,
stimulate endocytosis [5]. Another example is β‐adrenergic
­ ­phagocytosis that imports extracellular pathogens, and caveolin‐­
receptors that signal the recruitment of the clathrin adaptor β‐ mediated endocytosis. Although relatively little is known
arrestin [6]. regarding these mechanisms in comparison to clathrin‐mediated
Clathrin is recruited from the cytosol to bind AP2 or other endocytosis, perhaps the most broadly appreciated is the forma-
cargo‐specific adaptors that are directly linked to surface recep- tion of caveolae, which are small flask‐shaped vesicles at the
tors destined for internalization. Consisting of both heavy and plasma membrane that are enriched in sphingolipids and choles-
light chains, clathrin forms heterodimers that further assemble terol. In contrast to clathrin‐mediated endocytosis, caveolae are
into three‐legged triskelions. Clathrin assembly can occur relatively stable structures at the plasma membrane and are
around nascent vesicle pits, but has also been shown to assemble internalized at a very slow rate [26, 27]. This has led to some
at the plasma membrane prior to vesicle formation as flat, planar controversy regarding the role of caveolae in hepatocellular
lattices [7, 8]. Whereas clathrin heavy chains are capable of endocytosis, especially given the moderate expression of caveo-
binding to adaptins and other accessory proteins, the light chains lin, the main structural component of caveolae, in the liver.
prevent premature lattice assembly in the cytosol and are regu- Nonetheless, hepatocyte caveolae are internalized by dynamin
lated by calcium binding and phosphorylation [8–13]. scission machinery [28, 29], and caveolin has been shown to
Clathrin does not provide sufficient force for inward play several important hepatic functions, including lipid metab-
­membrane curvature. Thus, inward curvature on the plasma olism and liver regeneration [30].
membrane is thought to require actin polymerization that syner-
gizes with the clathrin coat and vesicle scission machinery. In
yeast, actin binds clathrin adaptors such as Sla2 and Ent1 to
“pull” coated pits inward, while actin polymerization and myo- HEPATOCELLUAR TRAFFICKING
sin at the plasma membrane aids in constriction of the endocytic OF ENDOCYTIC VESICLES
vesicle neck [1]. BAR domain proteins, such as endophilin and
amphiphysin, also sense membrane curvature and facilitate Once internalized by either clathrin‐dependent or clathrin‐inde-
endocytic neck constriction prior to vesicle scission [14, 15]. pendent endocytosis, endocytic vesicles fuse with EEs that
Other adaptors are also known to link the actin cytoskeleton to serve as a central hub for sorting and trafficking of endocytic
the endocytic vesicles, including the actin‐binding proteins pro- cargo. From here, cargo can be directed through multiple sort-
filin, synapsin, syndapin, and cortactin, some of which interact ing pathways, such as recycling back to the plasma membrane,
with the molecular machinery that drives vesicle scission. retrograde transport to the Golgi, degradation by lysosomes and
Scission of clathrin‐coated vesicles marks their separation from late endosomes, or secretion from endolysosomal vesicles that
the plasma membrane. This is accomplished by a family of large fuse with the plasma membrane (Figure 6.2). Recycling requires
GTPase mechanoenzymes known as dynamins. Dynamin binds to the sorting and concentration of cargo along membrane tubules
GTP near its N‐terminus, while membrane attachment is mediated that extend from endosomal vesicles. Cargo that is not recycled
by a plekstrin homology (PH) domain. Dynamin proteins also bind can be directly internalized within small intraluminal vesicles
a variety of effectors via a C‐terminal proline‐rich domain (PRD) (ILVs), forming distinct endosomal structures known as multi-
[16–18], including BAR domain‐containing proteins and cytoskel- vesicular bodies (MVBs). These cargo‐rich ILVs are then
etal adaptors. Dynamin has been referred to as a “pinchase” degraded by the late endosomal pathway, marked by the acidifi-
because of its ability to generate discrete vesicles from invaginated cation of MVBs by fusion with lysosomes. Hepatocytes may
coated pits. Dynamin proteins can self‐assemble into oligomeric also secrete the contents of MVBs, late endosomes, and lys-
rings along lipid vesicles [19] and membrane tubules [20]. In vitro osomes into the bile as these vesicles fuse with the apical plasma
studies of various dynamin mutations have revealed that GTP membrane. These diverse events are dictated by the spatial and
hydrolysis generates the constrictive force for vesicle scission [21]. temporal coordination of various protein complexes, post‐trans-
Following vesicle scission, the clathrin coat is disassembled, lational modifications, phospholipid dynamics, and cytoskeletal
and nascent vesicles fuse with other peripheral endosomes. regulators. In addition, it has become increasingly appreciated
Clathrin disassembly requires the ATPase Hsc70 and the remod- that endosomal vesicles play diverse roles in cellular homeosta-
eling of membrane PI(4,5)P2 phospholipids to phosphatidylino- sis beyond that of cargo trafficking [31].
sitol 4‐phosphate (PI(4)P) by synaptojanin. Hsc70 works
together with its co‐chaperone GAK/auxilin that is recruited in
Cargo sorting at early endosomes
part by PI(4)P phospholipids [22, 23]. After GAK/auxilin
recruitment, the clathrin coat is released from the vesicle. It is Newly formed endocytic vesicles converge at the early endo-
worth noting that auxilin/GAK activity is not restricted to clath- some (EE), also called the “sorting endosome,” that decides
rin uncoating but is also required for the dynamin‐dependent the  fate of endocytosed cargo. The importance of this sorting
constriction of coated pits [24]. process cannot be overstated as it is essential to hepatocyte
function and insures the proper degradation, recycling, or
­transcytosis of trophic and growth factor receptors, bile acid
Clathrin‐independent endocytosis transporters and other integral membrane proteins [32]. EEs
Clathrin‐independent endocytosis is a broad category of contain multiple subdomains that are thought to mediate differ-
­internalization pathways that can be dynamin dependent and ent ­
sorting pathways [33]. For example, tubular membrane
independent [25]. This includes fluid‐phase endocytosis extensions arising from specific endosomal subdomains are
6:  Endocytosis in Liver Function and Pathology 65

Pl(4,5)P2 Basolateral membrane

Rab4
TJ
Rab11 Rab5
WASH
retromer
BC
ESCRT
EHD1
EHD3 Early/sorting
endosome Rab27b
EHD1 MVB
Rab11 EHD3
Endocytic
recycling Pl(3)P Rab27a
Pl(3,4)P2 Rab5
compartment
WASH
retromer Lys

Nucleus Golgi
Rab7
Late
Lys
endosome

Pl(3,5)P2

Figure 6.2  Hepatocyte endocytic trafficking routes and the distribution of Rab GTPases and phosphatidylinositides at the compartments within
the cell. Rab GTPases mediate trafficking from coated vesicles to early/sorting endosomes (Rab5) or late endosomes (Rab7) as well as recycling
back to the plasma membrane directly (Rab4) or indirectly via the endocytic recycling compartment (Rab11). Note that the plasma membrane and
the endosomal pathways are enriched in a subset of phosphatidylinositides that contribute to Rab targeting and additional specificity of the different
compartments. TJ, tight junction; BC, bile canaliculus; MVB, multivesicular body.

destined for recycling, whereas large vesicular subdomains are subdomains, followed by actin polymerization that generates
thought to “mature” down the late endosomal pathway. Even force for tubule initiation and extension. Membrane tubules
within a single EE, there is thought to be variation in pH and further extend along microtubule tracks, and scission machin-
phospholipid composition at different regions that aid in sorting ery separates the tubule from the EE, releasing tubular vesicle
between recycling, secretory, and degradative routes. carriers that transport the sorted cargo to the plasma
The trafficking routes of endosomal vesicles is largely facili- ­membrane, endocytic recycling compartment (ERC), or Golgi
tated by a family of small GTPases known as Ras‐associated apparatus. Recent work has considerably advanced our under-
binding (Rab) proteins. Approximately 60 different Rab pro- standing of the molecular mechanisms underlying each of
teins play important roles in membrane and vesicle trafficking these steps.
pathways in humans, most of which are expressed in the hepato- In order for endocytic sorting to occur, endosomes must be
cyte [34, 35]. Rab proteins guide endocytic vesicle trafficking able to recognize and sequester specific cargo proteins. This
by binding to effector proteins that modulate endocytic vesicle “cargo recognition” process appears to rely on a diverse class of
functions. For EE function, Rab5 is predominant along with proteins known as the sorting nexins (Snx) [39]. Snx proteins
binding effectors such as phosphoinositide 3‐kinase (PI3K)/ contain a conserved PX domain that binds PI(3)P phospholipids
VPS34, EEA1, rabenosyn‐5, and others [36]. PI3K/VPS34 enriched on EEs, as well as cargo‐specific binding domains that
changes the phospholipid composition of EEs to phosphati- vary between Snx family members. Snx proteins also contain
dylinositol 3‐phosphate (PI(3)P), giving these vesicles a distinc- additional motifs such as BAR domains that sense membrane
tive membrane signature that aids in the recruitment of other curvature, FERM domains that aid in cargo recognition, and
Rab5 effectors. EEA1 binds both Rab5 and PI(3)P and mediates others [40]. Sorting nexins work in concert with the retromer, a
the fusion of endocytic vesicles by coordinating SNARE pro- trimeric protein complex composed of Vps26, Vps35, and
teins such as syntaxin6 and syntaxin13. Rabenosyn‐5 is thought Vps29 that is critical to endocytic sorting. More recently, a new
to regulate “fast recycling” of cargo such as the transferrin retromer‐like complex known as retriever (consisting of DSCR3,
receptor, and delivery of cargo to the endocytic recycling center C160rf62, and VPS29) was also identified to work in concert
for “slow recycling” [37]. with Snx17 [41]. Retriever/Snx17 promotes the recycling of a
Cargo sorting is driven by dynamic membrane tubules that subset of cargo molecules that is distinct from retromer‐depend-
extend from EE subdomains [38]. Although the mechanisms ent sorting. Thus, the ability of EEs to accurately decipher dif-
that form these tubules are still unclear, several lines of evi- ferent types of cargo, and to organize cargo at endosomal
dence suggest that cargo is first concentrated at specific EE subdomains, is absolutely critical to hepatocellular vesicle
66 THE LIVER: HEPATOCELLUAR TRAFFICKING OF ENDOCYTIC VESICLES

transport, and involves an impressive synchronization between regulated by the NRTK c‐Abl kinase. Whereas TfR is normally
cargo adaptors, protein complexes, and the cytoskeleton. recycled back to the plasma membrane, inhibition of c‐Abl
In addition to sorting cargo at endosomal subdomains, kinase redirects TfR to late endosomes for degradation [49].
Snx–retromer/retriever complexes coordinate actin dynamics Serine–threonine kinases, such as cyclic AMP (cAMP)‐depend-
at these sites by recruiting the WASH complex [40, 42], ent protein kinase A (PKA), are well‐described to regulate
which consists of five individual proteins (WASH1, Fam21, GPCR sorting. For example, activation of the β‐adrenergic
Strumpellin, SWIP, and CCDC53). WASH stimulates Arp2/3 receptor leads to multiple PKA phosphorylation events at its C‐
for actin nucleation at EE sites where membrane tubulation terminal cytoplasmic tail, which recruits β‐arrestins that induce
occurs. It is thought that actin generates the force required GPCR internalization and desensitization [6]. In addition, PKA
for tubule formation, extension, and/or possibly scission. phosphorylation of the β‐adrenergic adaptor gravin is essential
Fam21 of the WASH complex seems to act as the main teth- for receptor recycling and resensitization back to the plasma
ering link to other pathways. For example, Fam21 is known membrane [50–53]. Thus, it is clear that cargo‐specific phos-
to bind Vps35 of the retromer complex [43]. Fam21 also phorylation by NRTKs and serine–threonine kinases drive many
links the WASH complex to retriever by binding yet another important aspects of endocytic regulation, including cargo inter-
recently described protein complex known as CCC nalization and sorting between degradative and recycling
(COMMD1, CCDC22, CCDC93) [41]. pathways.
Early endosomal tubules elongate and move along microtu-
bule tracks. This requires tethering to molecular motor proteins
that travel toward (minus‐end dyneins) or away from (plus‐end Recycling endosomes and the endocytic
kinesins) the microtubule organizing center (MTOC). The
recycling compartment
small GTPase Rab7 facilitates the tethering of EE tubules to
molecular motor proteins in cooperation with Snx‐BAR pro- Whereas early endosomal cargo can be recycled directly back to
teins. The tethering of EE tubules to dyneins versus kinesins the plasma membrane (fast recycling), a “slow recycling” path-
plays an important role in dictating cargo destination. For way also exists whereby cargo is shunted through tubular recy-
example, dynein‐tethered tubules are effectively transported cling endosomes (RE) that are located at the cell periphery or
back to the perinuclear endocytic recycling compartment and/ clustered within the perinuclear endocytic recycling compart-
or Golgi apparatus, whereas kinesin‐tethered tubules are ment (ERC). REs are enriched in Rab11, distinguishing these
­transported to the plasma membrane. The scission machinery vesicles from Rab5‐positive EEs.
that separates nascent tubules from EEs is currently unclear, The delivery of cargo from Rab5‐positive EEs to Rab11‐
but likely involves actin, the ATPase EHD1, and/or dynamin positive REs seems to rely on a family of Eps15 homology
GTPases. domain (EHD) proteins that play diverse roles at various stages
of endosomal sorting and recycling [54, 55]. EHD1 is thought
to cause scission of tubular REs arising from the ERC, and
Regulation of cargo sorting by post‐ perhaps EE tubules as well [56]. Indeed, EHD1‐depleted cells
show aberrant recycling of the transferrin receptor [57], as do
translational modifications EHD3 and EHD4, which also aid in the transition of the TfR
As described above, cargo sorting between recycling, degrada- from EE to RE [58, 59]. In addition to their well‐described
tion, and secretion pathways relies on various endosomal adap- roles in the RE, EHD proteins also affect sorting at the EEs by
tors that bind specific cargo. These cargo–adaptor interactions, regulating Rab5 and its effectors rabenosyn‐5 and rabakyrin‐5
as well as cargo fate, are guided by post‐translational modifica- [58–60].
tions, such as ubiquitination and phosphorylation, that direct Tubular REs arising from the perinuclear ERC deliver cargo
downstream sorting pathways. back to the plasma membrane. The generation of these struc-
Ubiqutin (Ub) is an 8 kDa, highly conserved protein that can tures is still under investigation, but likely involves a variety of
be covalently linked to lysine residues on target proteins. This mechanisms. Interestingly, the WASH complex, which is impor-
can occur as a polyubiquitin chain or as a single monoubiquitin tant in the actin‐based budding of EE tubules, is also present on
attached to one or multiple lysine residues. Whereas ubiquitina- REs and suggests its role in the generating tubular carriers from
tion of soluble proteins leads to their degradation by the protea- the ERC [61]. However, EHD proteins also promote both the
some [44, 45], ubiquitination of endocytic cargo is a signal for scission of ERC‐derived tubular carriers (EHD1) and the stabi-
MVB internalization and degradation by the late endosomal lization of these structures (EHD3). EHD1 seems to operate
pathway (described in a later section). A prototypical example within a protein complex containing the endosomal trafficking
of this process is the EGFR that is ubiquitinated following adaptors MICAL‐L1 and syndapin2, both of which can generate
ligand stimulation, which prevents EGFR recycling and pro- tubules from phosphatidylserine (PS)‐rich liposomes in vitro
motes its sorting to degradative pathways [46, 47]. [62]. The EHD1–MICAL‐L1–syndapin2 complex may also
Phosphorylation by nonreceptor tyrosine kinases (NRTK) work on EE‐derived tubular carriers, perhaps even through
and by serine–threonine kinases also play important roles in binding to the Vps26 subunit of the retromer complex [60, 63,
endocytic cargo trafficking. For example, the NRTK Src was 64]. Thus, the generation of tubules is critical to cargo recycling
shown to phosphorylate the EGFR adaptor CIN85 to regulate directly from EE vesicles and through the ERC, but the synergy
EGFR ubiquitination and degradation by the late endosomal between the different molecular machineries on early endosomes
pathway [48]. In addition, the transferrin receptor (TfR) is and REs is still unclear.
6:  Endocytosis in Liver Function and Pathology 67

Multivesicular bodies and the late many questions remain regarding the mechanisms that package
endosomal pathway soluble cytosolic materials into ILVs, the role of EVs in cell
communication, and their potential synthetic use in drug deliv-
Hepatocellular cargo that is not recycled will progress down ery systems.
the  late endosomal pathway for degradation or secretion into
the  apical bile canaliculus. Due to the acidic pH of late
endosomes and their role in degradation of proteins and lipids, Lysosomes/late endosomes
late endosomes are more akin to lysosomes than EEs. Late As described above, lysosomes and late endosomes share some
endosomes can be distinguished experimentally from EEs by overlapping functions in hepatocytes. In conjunction with their
their acidic luminal pH, enrichment in PI(3,5)P2 phospholipids, role in degrading proteins and lipids, lysosomes also serve
and association with Rab7 and other protein markers. Although as  nutrient sensors and platforms for signal transduction.
the transition from Rab5‐positive EEs to Rab7‐positive late Lysosomal nutrient sensing is largely mediated by the presence
endosomes is not fully understood, the prevailing model is that or absence of amino acids that regulate the mammalian target of
EEs themselves “mature” to become gradually more acidic by rapamycin (mTOR), a nutrient‐sensing kinase that is critical in
fusion with smaller lysosomes that donate acid hydrolases metabolic liver diseases such as non‐alcoholic fatty liver disease
responsible for the breakdown of proteins, lipids, and nucleic (NAFLD), NASH, and hepatocellular carcinoma [72]. mTOR
acids at low pH [65]. At the same time, endosomal cargo that is operates as part of two distinct complexes: mTOR complex 1
destined for degradation will be invaginated within small, <100 (mTORC1) and mTOR complex 2 (mTORC2). While both of
nm ILVs. Early and late endosomes containing ILVs are these complexes regulate specific facets of cell growth and pro-
­identified as multivesicular bodies (MVBs) due to their unique liferation, mTORC1 is functionally linked to nutrient sensing at
morphology. Cargo‐rich ILVs are degraded by lysosomal acidi- the lysosome/late endosome. Amino acids stimulate the lysoso-
fication via the late endosomal pathway, as is the case with the mal recruitment of mTORC1 to regulate enzymes involved in
EGFR, but MVBs may also fuse with the plasma membrane to protein synthesis (ribosomal p70 S6 kinase and 4‐EBP1), lipid
secrete ILVs as extracellular vesicles/exosomes [66]. synthesis (SREBP1) and others. Active mTORC1 also sup-
Ubiquitination of endosomal cargo serves as the main signal for presses “starvation” pathways such as autophagy (discussed in
sequestration into MVB ILVs [67]. Cargo that is destined for MVB more detail later) by the inhibitory phosphorylation of Ulk1 and
internalization relies on endosomal sorting complexes required transcription factor EB (TFEB). In general, active mTORC1
for transport (ESCRT‐0, ‐I, ‐II, and ‐III) that recognize ubiquitin‐ signifies a “fed” state of nutrient abundance and thus activates
tagged membrane cargo for sorting into endosomal subdomains. pathways related to cell growth and proliferation, while simulta-
Ubiquitin recognition is achieved by ESCRT‐0, ESCRT‐I, and neously suppressing pathways that are important during “starved”
ESCRT‐II components that contain ubiquitin‐binding motifs and states where nutrient supply is limited.
cluster cargo in cooperation with endosomal clathrin lattices and Lysosomes and late endosomes play important roles in cell
the actin cytoskeleton. ESCRT‐III contains no ubiquitin recogni- signaling by storing intracellular calcium (Ca2+) that serves as a
tion, but oligomerizes as a concentric spiral around clustered cargo second messenger for signal transduction. Since Ca2+ concentra-
and is presumed to generate force, much like the action of a spring, tions in the cytoplasm are very low, release of lysosomal cal-
to push cargo inward into the lumen of the MVB [67, 68]. ESCRT‐ cium by lysosomal calcium channels can trigger nearby
III filaments are subsequently disassembled by the action of the substrates. For example, lysosomal calcium release can activate
ATPase Vps4, and this step is believed to mediate the final scission TFEB, a master transcriptional regulator of lysosome biogene-
of invaginated ILVs into the MVB lumen. It should be noted that sis, lipid catabolism, autophagy, and mitochondrial β‐oxidation
while ubiquitin modifications are important for cargo internaliza- [73]. Under basal conditions, TFEB is phosphorylated by
tion into MVB, recent studies indicate that ESCRTs also partici- mTORC1 and is sequestered in the cytosol. However, lysosomal
pate in ubiquitin‐independent cargo sorting. This may involve calcium release activates the phosphatase calcineurin, leading to
direct binding of cargo to ESCRTs, ESCRT adaptors, and/or endo- TFEB dephosphorylation and translocation into the nucleus. A
somal phospholipids [69]. similar pathway was later described for TFE3, which is closely
Cargo‐rich ILVs are degraded by the late endosomal pathway, related to TFEB, in the regulation of lysosome biogenesis and
but may be spared from degradation if the MVB fuses with the energy metabolism [74]. Thus, the signaling events propagated
plasma membrane. This unconventional secretory method by lysosomes/late endosomes are closely related to their
releases ILVs as extracellular vesicles (EVs), or “exosomes,” ­important roles as nutrient sensors in energy metabolism and
into the extracellular milieu. Indeed, EVs have been detected in autophagy.
blood, bile, and urine, indicating the potentially widespread
nature of this pathway to transmit cellular information through-
out the body. Various EV analytical studies have demonstrated Hepatocyte polarity in endocytic trafficking
that these vesicles contain much more than integral membrane Hepatocytes possess a unique epithelial polarity that contributes
cargo, but contain specific classes of lipids and soluble cytosolic to specialized endocytic vesicle trafficking routes [75]. In tradi-
materials such as microRNAs and various protein enzymes. In tional epithelium, cells possess an apical plasma membrane that
liver pathology, EVs are thought to play a role in tissue signal- faces the luminal space, a lateral domain that contacts adjacent
ing cell stress during non‐alcoholic steatohepatitis (NASH) and epithelial cells, and a basal domain anchored to the basal ­lamina.
other hepatic insults [70, 71]. Although it is clear that EVs play In hepatocytes, the apical (canalicular) domain faces the bile
an important role in tissue homeostasis and disease progression, canalicular lumen, and this domain is separated from the basal
68 THE LIVER:  AUTOPHAGY AND THE ENDOCYTIC MACHINERY

and lateral domains by tight junctions. The basolateral (sinusoi- the cytoplasm prior to fusion with lysosomes/late endosomes.
dal) domain interfaces with adjacent hepatocytes and the perisi- This fusion results in an acidic “autolysosome” in which the
nusoidal space of Disse that contains blood plasma, fenestrated autophagic cargo is degraded.
endothelial cells, and hepatic stellate cells. Endosomal vesicles contribute to several stages of autophagy
A subset of endocytic cargo undergoes a unique hepatocyte‐ [77]. Most critical is the contribution of acidic vesicles derived
specific mode of transcytosis (Figure 6.2), whereby endocytic from the endocytic pathway (i.e. lysosomes, late endosomes)
vesicles formed at the basolateral domain eventually deposit that fuse with autophagosomes to form autolysosomes that
their contents at the apical bile canaliculus via specialized, sub- degrade autophagic cargo. In hepatocytes, the late endosomal
apical endosomes [32]. Lysosomes also secrete cytosolic mate- Rab7, as well as the membrane scission protein dynamin2, are
rials into the bile through an upstream pathway called autophagy essential autophagy proteins that mediate a process known as
(discussed in the next section) [76]. Together with late autophagic lysosomal reformation (ALR), whereby nascent lys-
endosomes, these vesicles deliver bile acids, EVs, phospholip- osomes are generated from recycling tubules that extend from
ids, cholesterol, polymeric immunoglobulin A (pIgA – part of autolysosomes and late endosomes [84]. Rab7 has also been
mucosal system), and other materials into the bile [31]. implicated in the fusion of autophagosomes with lysosomes/late
endosomes [85]. In addition to late endosomes, other endoso-
mal vesicles are also known to play important roles in autophagic
progression. For example, EEs serve as one of the membrane
AUTOPHAGY AND THE ENDOCYTIC sources for phagophore formation [86]. Rab11‐positive REs are
MACHINERY known to harbor various components of the autophagic machin-
ery, and Rab11 itself has been implicated in the fusion of
Beyond cargo internalization and sorting, endocytic vesicles autophagosomes with MVBs, forming a structure known as the
play additional roles in cell physiology [31]. A prominent amphisome [87, 88].
example of this is the intersection of late endosomes with the Endosomal vesicles also facilitate other forms of autophagy
process of autophagy, a “self‐eating” pathway that degrades that are independent of autophagosome formation [89]. One
cytosolic materials (whole organelles, lipids, proteins, and prominent example of this is chaperone‐mediated autophagy,
other materials) to meet energy demands during periods of low whereby lysosomes directly target cytosolic proteins that con-
nutrient availability [77, 78]. Autophagy also plays a critical tain a pentapeptide (KFERQ) motif that binds to the chaperone
role in liver function by clearing damaged proteins and orga- Hsc70. This interaction facilitates the translocation of substrate
nelles that induce cellular stress. Dysregulation of autophagy proteins into the lysosomal lumen via lysosome‐associated
contributes to a wide range of liver diseases, including steato- membrane protein type 2A [90]. Dysregulation of this process
sis, steatohepatitis, cirrhosis, cancer, drug‐induced liver injury, in mouse livers leads to gross metabolic dysregulation,
and others [reviewed in 79]. ­especially the accumulation of hepatocellular lipids and the
Autophagy mediates the transport of cytosolic materials to the depletion of glycogen stores [91].
lysosome for degradation. While different subtypes of autophagy
exist, the canonical form, macroautophagy, is often referred to as Mitochondrial autophagy protects against
simply “autophagy.” This process can be activated in the liver by
prolonged fasting as well as glucagon stimulation, and is sup-
liver injury
pressed by amino acids and/or insulin. These pathways are A prominent function of the liver is clearance of toxins from the
thought to converge at the regulation of mTOR activity. mTOR blood by metabolizing blood alcohol, pharmaceutical drugs,
attenuates autophagy by inhibiting the Atg conjugation system and other nutrients. Accordingly, liver injury is a prominent
that drives the formation of double‐membrane phagophores concern in drug development and accounts for >50% of acute
[80]. Phagophores are tubular membranes derived from the liver failure [79]. The importance of autophagy in this process
endoplasmic reticulum (ER) and a variety of other cellular has been observed particularly in the context of acetaminophen
sources such as Golgi, plasma membrane, and endosomal (APAP) overdose, which induces severe liver injury [92].
vesicles [81–83]. These membranous structures (as well as
­ Autophagy is activated by excessive APAP to clear damaged
autophagosomes and autolysosomes) are decorated by a key mitochondria that produce reactive oxygen species (ROS), lead-
autophagy protein known as microtubule‐associated protein 1 ing to hepatocellular death. The degradation of mitochondria by
light chain 3, or LC3, which is “activated” by the Atg conjuga- autophagy is a protective mechanism, as pharmacological stim-
tion system during autophagy initiation and links cytosolic ulation of autophagy protects against APAP‐induced liver injury
materials to autophagic membranes. LC3 accomplishes this task in mice [93].
by interacting with a family of autophagy receptors (p62/ The selective autophagy of damaged mitochondria, termed
SQSTM1, NBR1, OPTN, and others) that recognize ubiquity- “mitophagy,” seems to utilize both autophagic and endosomal
lated cargo in the cytoplasm, but LC3 may also directly bind to machineries. The process is driven by mitochondrial fission that
autophagic cargo proteins containing LC3‐interacting regions produces smaller sized mitochondria that are more susceptible
(LIR). The binding of LC3 to autophagy receptors aids in trap- to autophagic engulfment [94]. In canonical mitophagy,
ping cytosolic materials to the phagophore, as this membrane damaged mitochondria accumulate PTEN‐induced putative
­
extends to envelope the autophagic cargo. When fully engulfed, kinase 1 (PINK1) that recruits E3 ubiquitin ligase parkin to the
the phagophore closes to become a double‐membrane auto­ ­mitochondrial outer membrane [95]. This generates ubiquitin
phagosome. Autophagosomes sequester autophagic cargo from signatures on the mitochondrial surface that are recognized by
6:  Endocytosis in Liver Function and Pathology 69

autophagy receptors that recruit LC3‐positive phagophores for autophagic mechanism for LD breakdown [115–117]. Although
engulfment within the autophagosome [95, 96]. Recent reports the mechanisms that traffic hepatocellular LDs from the cyto-
indicate that this engulfment may also be mediated directly plasm to the lumen of lysosomes and late endosomes are not
by  endosomal vesicles even in the absence of traditional completely understood, it is clear that numerous Rab proteins
autophagosome‐based mitophagy [97, 98]. In this model,
­ and other endocytic machinery play important roles in this pro-
termed “endosomal mitophagy,” parkin recruits damaged mito- cess. For example, both Rab7 and Rab10 are activated on the
chondria to Rab5‐positive EEs that utilize the ESCRT machin- LD surface during periods of low nutrient availability. Rab10
ery to bind ubiquitylated mitochondria. ESCRT recognition of was shown to form a trimeric complex with its effector Eps15
ubiquitylated mitochondria leads to an MVB‐like mitochondrial homology binding protein 1 (EHBP1) and EHD2 on the LD sur-
engulfment and degradation by the late endosomal pathway. face, and this complex is important for phagophore recruitment
In addition to mitophagy, endosomal vesicles have recently and extension around the LD [118]. Rab7 drives LD interaction
been reported to play a novel role in mitochondrial fission and with multivesicular bodies and late endosomes marked by the
fusion. These findings suggest that lysosomes make an intimate membrane tetraspanin CD63 [119]. The direct interaction of
connection with mitochondria, especially at sites of mitochon- LDs with MVBs/endosomes raises the possibility that alterna-
drial fission [99]. The tethering of mitochondria to lysosomes tive forms of “microlipophagy” may exist in parallel with tra-
requires active, GTP‐bound Rab7 on the lysosome, whereas ditional autophagosome‐based “macrolipophagy” (Figure  6.3)
untethering requires a protein known as TBC1D15 that inacti- Interestingly, Rab7 was also shown to be inhibited by chronic
vates Rab7 by facilitating GTP hydrolysis and dissociation. ethanol exposure, which alters lysosomal morphology and
Interestingly, TBC1D15 is recruited to the mitochondria to reg- motility to perturb the lipophagic catabolism of LDs during
ulate Rab7 activity, suggesting a bidirectional mechanism alcoholic fatty liver progression [120]. Other endocytic proteins
whereby mitochondria regulate lysosomal positioning via Rab7 also appear to be involved in hepatocellular lipophagy, includ-
inactivation, and lysosomes act on mitochondria to mediate ing clathrin, dynamin2, and Vps4 [84, 119].
fission. In addition to lipophagy, several studies now show a bidirec-
tional synergy between autophagy and cytosolic lipases. First,
various autophagy pathways are known to promote ATGL
Autophagy and lipid droplet homeostasis
­activity in hepatocytes to protect against fatty liver progression.
The liver is a central hub for the storage and regulation of neutral For  example, chaperone‐mediated autophagy was recently
lipids such as triglycerides and cholesterol esters. These lipids shown to degrade LD proteins that inhibit ATGL‐mediated
are assembled from free fatty acids (FFA) that are synthesized de lipolysis [121]. Macroautophagy also seems to facilitate
novo or imported from extracellular sources (i.e. dietary fat, ­hepatocyte lipolysis, as both HSL and ATGL contain several
lipids released from adipose tissue) and stored within specialized LC3‐interacting regions that link these cytosolic lipases to
organelles known as lipid droplets (LDs) [32]. Dysregulation of autophagosomes that function to deliver HSL and ATGL to the
these important hepatocellular functions leads to profound lipo- surface of LDs [122]. Conversely, ATGL‐mediated lipolysis has
dystrophies that impact both liver function and whole‐body lipid been reciprocally shown to induce autophagy at the transcrip-
homeostasis. Most common are fatty liver diseases that affect tional level. In this model, ATGL activity releases FFAs
20–30% of the world population and coincide with other meta- that  serve as signaling ligands for PGC‐1α and PPAR‐α. The
bolic diseases (obesity, diabetes, metabolic syndrome) and/or ­mechanism for this relies on the activation of NAD‐dependent
chronic alcohol consumption [100, 101]. Central to hepatic fat deacetylase sirtuin 1, which facilitates the interaction of PPAR‐α
storage and energy utilization, LDs store intracellular neutral with deacetylated PGC‐1α to promote autophagic gene tran-
lipids surrounded by a phospholipid monolayer that associates scription and other metabolic programs [123, 124].
with proteins involved in lipid storage, breakdown, and fusion
[102–104]. In addition, several proteomic studies have shown
that LDs associate with numerous Rab GTPases, as well as
endosomal vesicle machinery found on endosomes, multivesicu-
VIRAL INFECTION
lar bodies, and lysosomes [105–109]. AND HEPATOCELLULAR
LDs are catabolized by two different types of lipases – ­neutral ENDOCYTOSIS
lipases that reside in the cytoplasm and acid lipases that reside
in lysosomes/late endosomes. Neutral lipases, such as adipose Human pathogens such as hepatitis B virus (HBV) and hepatitis
triglyceride lipase (ATGL) and hormone‐sensitive lipase (HSL), C virus (HCV) are a leading cause for liver diseases such as
are recruited from the cytosol to the LD following β‐adrenergic steatohepatitis, cirrhosis, and hepatocellular carcinoma. These
receptor activation of the cAMP/PKA pathway. It was once viruses hijack normal hepatocellular processes, particularly host
thought that these neutral lipases played a more prominent role endocytic pathways that are utilized for infection, propagation,
in adipose tissue and skeletal muscle, but numerous studies have and  secretion to neighboring cells and tissues [reviewed in
now demonstrated their importance in liver function and meta- 125].  Five hepatitis viruses have been described that hijack
bolic disease pathology [110–114]. Acid lipases such as lysoso- different  endocytic pathways for replication and secretion:
­
mal acid lipase (LAL) are active within the low pH environment HBV  (Hepadnaviridae family), HDV (Deltaviridae), HCV
of lysosomes/late endosomes, and thus cannot be recruited to (Flaviviridae family), HAV  (Hepatovirus Picornaviridae),
the LD surface in the same manner as the ­cytosolic lipases. LAL and  HEV (Hepervirus Heperviridae). Although the genetic
gains access to LDs via lipophagy (Figure  6.3), a selective and  molecular components of these hepatic viruses vary
70 THE LIVER: VIRAL INFECTION AND HEPATOCELLULAR ENDOCYTOSIS

(a) Plasma membrane


Clathrin-mediated
endocytosis Microlipophagy
LD
?

MVB/Late
EE endosome

Macrolipophagy

ALR
LD Lysosome
Dyn2
Phagophore Autophagosome Autolysosome

(b) (c)

Figure 6.3  Lipophagy in hepatocytes. (a) A working model of two distinct lipophagy pathways. Microlipophagy is proposed to occur by the direct
uptake of lipid droplets by endosomal vesicles such as multivesicular bodies (MVBs) and late endosomes for degradation by lysosomal enzymes.
Macrolipophagy utilizes traditional autophagy machinery whereby lipid droplets (LDs) are targeted by phagophores for complete engulfment within
autophagosomes which fuse with lysosomes to become degradative autolysosomes. The terminal stages of macrolipophagy incorporate autolyso-
some reformation (ALR), whereby membrane tubules extending from the autolysosome undergo scission by dynamin 2 (Dyn2). This generates
nascent lysosomes that contribute to autophagic and late endosomal degradation. (b) Transmission electron micrograph shows an LD (white arrow)
encased within an autophagic membrane in Huh7 human hepatoma cells. Reproduced from [118] with permission of American Association for the
Advancement of Science. (c) Electron micrograph of a siRNA Dyn2‐depleted Hep3B human hepatoma cell shows an autolysosome tubule (white
arrows) that is elongated in the absence of tubule scission machinery. Reproduced from [84] with permission of Rockefeller University Press.

considerably, each virus consists of a nucleotide genome of facilitate membrane attachment. HAV or HEV attachment to
ssRNA (D,C,A,E) or dsDNA (B) that is encapsulated by a pro- the  ­hepatocyte is not well understood and seems to require
teinaceous core and surrounded by an outer envelope consisting HSPGs, asialoglycoprotein receptor (ASPGR), and others
of proteins and host‐derived lipids. A model for how each of [reviewed in 125].
these viruses are assembled as they advance through hepatocyte Hepatic viruses utilize clathrin‐mediated endocytosis for
endocytic pathways is depicted in Figure 6.4. hepatocellular uptake, but the contributions of other endocytic
routes such as caveolin‐mediated endocytosis, pinocytosis, and
phagocytosis cannot be ruled out [125]. Upon internalization,
Viral attachment and endocytosis these viruses traverse through hepatocyte endocytic pathways
Hepatitis viruses attach to the cell surface and can move prior to genome release into the nucleus via the cytoplasm. For
­laterally along the plasma membrane prior to endocytosis. For HBV, genome release into the nucleus is pH dependent and
HBV, this interaction is facilitated by hepatocyte heparin sul- requires early‐to‐late endosomal transitions mediated by Rab5
fate proteoglycans (HSPGs) and by sodium taurocholate and Rab7 [126]. The mechanisms supporting these important
cotransporting polypeptide (NTCP), both of which bind to steps, including how these viruses evade late endosomal degra-
HBV envelope ­proteins. HCV attachment is also facilitated by dation, remain unclear. In contrast to HBV and other hepatic
HSPGs, but has also been reported to interact with apolipopro- viruses, the HCV genome is translated in the cytoplasm follow-
teins, junctional proteins, surface receptors, and others that help ing release from EEs in a pH‐dependent manner [127].
6:  Endocytosis in Liver Function and Pathology 71

Virus BCDE ABE


C
C?
HSPGs Clathrin-mediated
endocytosis Autophagosome

Endosome

B ACDE
RNA MVB
Nucleus C
Golgi B
Viral
genome

DNA

B C? ESCRT:
RNA Lysosome
ABE
ER

Figure 6.4  Model depicting how the different hepatitis viruses (ABCDE) utilize common and distinct endocytic pathways in the hepatocyte
­during internalization, infection, maturation, and release. Modified from [125] with permission of John Wiley & Sons.

Viral assembly, replication, and Rab27b. These later Rabs are particularly important in
and hepatocellular secretion MVB‐based secretion, as Rab27a is thought to mediate MVB
docking/fusion with the plasma membrane, whereas Rab27b
HBV‐infected cells secrete both infectious virions and non‐ participates in the transfer of MVBs from microtubules to the
infectious subviral particles (SVPs) made up of envelope pro- actin‐rich regions at the cell periphery [134].
teins that likely serve as “decoys” against the host immune In conclusion, hepatitis viruses utilize hepatocyte membrane
system. SVPs are assembled in the ER and packaged for secre- trafficking machinery for infection and propagation. Despite the
tion within the ER–Golgi intermediate compartment (ERGIC) substantial progress that has been made toward understanding
[128]. For HCV, assembly occurs within ER‐derived double‐ how the host hepatocellular endocytic pathways are modified by
membrane vesicles (~150 nm in diameter) that are constructed these viruses, future work will need to further define the syn-
in response to viral infection. Interestingly, HCV also stimulates ergy between the endosomal compartments, nucleus, ER, Golgi,
the formation of LDs that are contained within these vesicles and the autophagic machinery in the viral life cycle.
[129]. LDs are thought to facilitate HCV replication, as core and
nonstructural proteins are found on the LD surface. LDs may
also contain viral double‐stranded RNA, suggesting LDs are
also a site of genomic replication. Other work has shown that FUTURE DIRECTIONS
LD‐associated proteins, such as Rab18 [130] and Plin3/Tip47
[131], also help facilitate HCV replication. New imaging and biochemical techniques have provided
Hepatitis viruses utilize different endo‐membrane vesicle detailed insights into the mechanisms of vesicle formation and
pathways for secretion. For example, HCV secretion requires post‐endocytic trafficking. Nevertheless, the list of protein and
trans‐Golgi machinery and Rab11, suggesting interplay between lipid components of the vesicle‐forming machinery continues
the Golgi and RE compartments [132]. Other viruses, such as to expand while the functions of many endocytic components
HBV, utilize the MVB for uptake and secretion through a mech- in specific transport processes are yet to be established. It is
anism that requires the ESCRT machinery. It was recently found important to understand how the hepatocyte manages to
that HBV induces the dramatic tubulation of MVBs and ­coordinate and control a massive protein and lipid network that
autophagosomes by modulating Rab7 activity, which stimulates supports vesicular trafficking events. In this context, ­challenges
lysosomal fusion and pH‐dependent secretion [133]. for the future will be to understand how different endocytic
Recent studies reveal that hepatic viruses utilize late ­endocytic cargos are sequestered and sorted, how membrane subdomains
compartments for maturation and release. Some of these viruses are maintained even within individual vesicles, and which
(HAV, HEV) utilize MVB‐derived membranes to envelop nas- phospho‐substrates are targeted by kinase signaling cascades
cent virions as their genomes do not encode envelope proteins. to control endocytic trafficking events. Equally exciting will
In addition to serving as a membrane source, MVBs also gener- be  the elucidation of alternative roles for endocytic vesicles
ate ILVs that support an exosome‐like shedding of hepatic in  processes such as the autophagy of mitochondria and
viruses. Trafficking of viruses to MVBs, as well as MVB fusion LDs.  Insights into these processes will provide a foundation
with the plasma membrane, seems to require the participation of for  understanding the function of hepatocytes in health and
several Rab proteins including Rab2b, Rab5a, Rab9b, Rab27a, disease.
72 THE LIVER:  REFERENCES

REFERENCES 27. Pelkmans, L., Burli, T., Zerial, M., and Helenius, A. Caveolin‐stabilized
membrane domains as multifunctional transport and sorting devices in endo-
cytic membrane traffic. Cell, 2004;118(6):767–80.
1. Kaksonen, M. and Roux, A. Mechanisms of clathrin‐mediated endocytosis.
28. Yao, Q., Chen, J., Cao, H. et al. Caveolin‐1 interacts directly with dynamin‐2.
Nature reviews Mol Cell Biol, 2018;19(5):313–26.
J Mol Biol, 2005;348(2):491–501.
2. Heuser, J. Effects of cytoplasmic acidification on clathrin lattice morphol-
29. Henley, J.R., Krueger, E.W., Oswald, B.J., and McNiven, M.A. Dynamin‐
ogy. J Cell Biol, 1989;108(2):401–11.
mediated internalization of caveolae. J Cell Biol, 1998;141(1):85–99.
3. Roth, T.F. and Porter, K.R. Yolk protein uptake in the oocyte of the mosquito
30. Fernandez‐Rojo, M.A. and Ramm, G.A. Caveolin‐1 function in liver physi-
Aedes aegypti. L. J Cell Biol, 1964;20:313–32.
ology and disease. Trends Mol Med, 2016;22(10):889–904.
4. Pearse, B.M. Clathrin: a unique protein associated with intracellular transfer
31. Schroeder, B. and McNiven, M.A. Importance of endocytic pathways in liver
of membrane by coated vesicles. Proc Natl Acad Sci U S A, 1976;73(4):
function and disease. Compr Physiol, 2014;4(4):1403–17.
1255–9.
32. Schulze, R.J. et al. The cell biology of the hepatocyte. J Cell Biol,
5. Kadlecova, Z., Spielman, S.J., Loerke, D. et  al. Regulation of clathrin‐­
2019;218(7):2096–112.
mediated endocytosis by hierarchical allosteric activation of AP2. J Cell
33. Short, B. Sorting out endosome form and function. J Cell Biol,
Biol, 2017;216(1):167–79.
2015;210(6):870.
6. Goodman, O.B., Jr., Krupnick, J.G., Santini, F. et al. Beta‐arrestin acts as a
34. Zhen, Y. and Stenmark, H. Cellular functions of Rab GTPases at a glance.
clathrin adaptor in endocytosis of the beta2‐adrenergic receptor. Nature,
J Cell Sci, 2015;128(17):3171–6.
1996;383(6599):447–50.
35. Stenmark, H. Rab GTPases as coordinators of vesicle traffic. Nat Rev Mol
7. Larkin, J.M., Donzell, W.C., and Anderson, R.G. Potassium‐dependent
Cell Biol, 2009;10(8):513–25.
assembly of coated pits: new coated pits form as planar clathrin lattices.
36. Jovic, M., Sharma, M., Rahajeng, J., and Caplan, S. The early endosome: a
J Cell Biol, 1986;103(6 Pt 2):2619–27.
busy sorting station for proteins at the crossroads. Histol Histopathol.
8. Heuser, J. Three‐dimensional visualization of coated vesicle formation in
2010;25(1):99–112.
fibroblasts. J Cell Biol, 1980;84(3):560–83.
37. Navaroli, D.M., Bellve, K.D., Standley, C. et  al. Rabenosyn‐5 defines the
9. Ungewickell, E. and Ungewickell, H. Bovine brain clathrin light chains
fate of the transferrin receptor following clathrin‐mediated endocytosis. Proc
impede heavy chain assembly in vitro. J Biol Chem, 1991;266(19):
Natl Acad Sci U S A, 2012;109(8):E471–80.
12710–14.
38. van Weering, J.R. and Cullen, P.J. Membrane‐associated cargo recycling by
10. Vigers, G.P., Crowther, R.A., and Pearse, B.M. Three‐dimensional structure
tubule‐based endosomal sorting. Semin Cell Dev Biol, 2014;31:40–7.
of clathrin cages in ice. EMBO J, 1986;5(3):529–34.
39. Cullen, P.J. Endosomal sorting and signalling: an emerging role for sorting
11. Nathke, I., Hill, B.L., Parham, P., and Brodsky, F.M. The calcium‐binding
nexins. Nat Rev Mol Cell Biol, 2008;9(7):574–82.
site of clathrin light chains. J Biol Chem, 1990;265(30):18621–7.
40. Wang, J., Fedoseienko, A., Chen, B. et al. Endosomal receptor trafficking:
12. Hill, B.L., Drickamer, K., Brodsky, F.M., and Parham, P. Identification of
retromer and beyond. Traffic, 2018;19(8):578–90.
the  phosphorylation sites of clathrin light chain LCb. J Biol Chem,
41. McNally, K.E. et al. Retriever is a multiprotein complex for retromer-inde-
1988;263(12):5499–501.
pendent endosomal cargo recycling. Nat Cell Biol, 2017;19(10):1214–25.
13. Schmid, S.L., Braell, W.A., Schlossman, D.M., and Rothman, J.E. A role for
42. Duleh, S.N. and Welch, M.D. WASH and the Arp2/3 complex regulate
clathrin light chains in the recognition of clathrin cages by ‘uncoating
endosome shape and trafficking. Cytoskeleton (Hoboken), 2010;67(3):
ATPase.’ Nature, 1984;311(5983):228–31.
193–206.
14. Meinecke, M., Boucrot, E., Camdere, G. et  al. Cooperative recruitment of
43. Harbour ME., Breusegem, S.Y., and Seaman, M.N. Recruitment of the endo-
dynamin and BIN/amphiphysin/Rvs (BAR) domain‐containing proteins leads
somal WASH complex is mediated by the extended “tail” of Fam21 binding
to GTP‐dependent membrane scission. J Biol Chem, 2013;288(9):6651–61.
to the retromer protein Vps35. Biochem J, 2012;442(1):209–20.
15. Dawson, J.C., Legg, J.A., and Machesky, L.M. Bar domain proteins: a role
44. Elsasser, S. and Finley, D. Delivery of ubiquitinated substrates to protein‐
in tubulation, scission and actin assembly in clathrin‐mediated endocytosis.
unfolding machines. Nat Cell Biol, 2005;7(8):742–9.
Trends Cell Biol, 2006;16(10):493–8.
45. Miller, J. and Gordon, C. The regulation of proteasome degradation by multi‐
16. Kim, Y. and Chang, S. Ever‐expanding network of dynamin‐interacting pro-
ubiquitin chain binding proteins. FEBS Lett, 2005;579(15):3224–30.
teins. Mol Neurobiol, 2006;34(2):129–36.
46. Huang, F., Goh, L.K., and Sorkin, A. EGF receptor ubiquitination is not
17. McNiven, M.A., Cao, H., Pitts, K.R., and Yoon, Y. The dynamin family
­necessary for its internalization. Proc Natl Acad Sci U S A, 2007;104(43):
of  mechanoenzymes: pinching in new places. Trends Biochem Sci,
16904–9.
2000;25(3):115–20.
47. Ravid, T., Heidinger, J.M., Gee, P., Khan, E.M., and Goldkorn, T. c‐Cbl‐
18. Praefcke, G.J. and McMahon, H.T. The dynamin superfamily: universal
mediated ubiquitinylation is required for epidermal growth factor receptor
membrane tubulation and fission molecules? Nat Rev Mol Cell Biol,
exit from the early endosomes. J Biol Chem, 2004;279(35):37153–62.
2004;5(2):133–47.
48. Schroeder, B., Srivatsan, S., Shaw, A., Billadeau, D., and McNiven, M.A.
19. Tuma, P.L. and Collins, C.A. Dynamin forms polymeric complexes in the
CIN85 phosphorylation is essential for EGFR ubiquitination and sorting into
presence of lipid vesicles. Characterization of chemically cross‐linked
multivesicular bodies. Mol Biol Cell, 2012;23(18):3602–11.
dynamin molecules. J Biol Chem, 1995;270(44):26707–14.
49. Cao, H., Schroeder, B., Chen, J., Schott, M.B., and McNiven, M.A. The
20. Takei, K., McPherson, P.S., Schmid, S.L., and De Camilli, P. Tubular mem-
endocytic fate of the transferrin receptor is regulated by c‐Abl kinase. J Biol
brane invaginations coated by dynamin rings are induced by GTP‐gamma S
Chem, 2016;291(32):16424–37.
in nerve terminals. Nature, 1995;374(6518):186–90.
50. Shih, M., Lin, F., Scott, J.D., Wang, H.Y., and Malbon, C.C. Dynamic com-
21. Sweitzer, S.M. and Hinshaw, J.E. Dynamin undergoes a GTP‐dependent
plexes of β2‐adrenergic receptors with protein kinases and phosphatases and
conformational change causing vesiculation. Cell, 1998;93(6):1021–9.
the role of gravin. J Biol Chem, 1999;274(3):1588–95.
22. Massol, R.H., Boll, W., Griffin, A.M., and Kirchhausen, T. A burst of auxilin
51. Lin, F., Wang, H., and Malbon, C.C. Gravin‐mediated formation of signaling
recruitment determines the onset of clathrin‐coated vesicle uncoating. Proc
complexes in beta 2‐adrenergic receptor desensitization and resensitization.
Natl Acad Sci U S A, 2006;103(27):10265–70.
J Biol Chem, 2000;275(25):19025–34.
23. Lee, D.W., Wu, X., Eisenberg, E., and Greene, L.E. Recruitment dynamics of
52. Fan, G.F., Shumay, E., Wang, H.Y., and Malbon, C.C. The scaffold protein
GAK and auxilin to clathrin‐coated pits during endocytosis. J Cell Sci,
gravin (cAMP‐dependent protein kinase‐anchoring protein 250) binds the
2006;119(Pt 17):3502–12.
β2‐adrenergic receptor via the receptor cytoplasmic Arg‐329 to Leu‐413
24. Sever, S., Skoch, J., Newmyer, S. et al. Physical and functional connection
domain and provides a mobile scaffold during desensitization. J Biol Chem,
between auxilin and dynamin during endocytosis. EMBO J, 2006;25(18):
2001;276(26):24005–14.
4163–74.
53. Tao, J., Wang, H.Y., and Malbon, C.C. Protein kinase A regulates AKAP250
25. Mayor, S., Parton, R.G., and Donaldson, J.G. Clathrin‐independent path-
(gravin) scaffold binding to the beta2‐adrenergic receptor. EMBO J,
ways of endocytosis. Cold Spring Harb Perspect Biol, 2014;6(6):a016758.
2003;22(24):6419–29.
26. Tagawa, A., Mezzacasa, A., Hayer, A. et  al. Assembly and trafficking of
54. Mintz, L., Galperin, E., Pasmanik‐Chor, M. et al. EHD1 – an EH‐domain‐
caveolar domains in the cell: caveolae as stable, cargo‐triggered, vesicular
containing protein with a specific expression pattern. Genomics, 1999;59(1):
transporters. J Cell Biol, 2005;170(5):769–79.
66–76.
6:  Endocytosis in Liver Function and Pathology 73

55. Pohl, U., Smith, J.S., Tachibana, I. et al. EHD2, EHD3, and EHD4 encode 84. Schulze, R.J., Weller, S.G., Schroeder, B. et al. Lipid droplet breakdown
novel members of a highly conserved family of EH domain‐containing requires dynamin 2 for vesiculation of autolysosomal tubules in hepato-
­proteins. Genomics, 2000;63(2):255–62. cytes. J Cell Biol, 2013;203(2):315–26.
56. Xie, S., Bahl, K., Reinecke, J.B. et al. The endocytic recycling compartment 85. Hyttinen, J.M., Niittykoski, M., Salminen, A., and Kaarniranta, K.
maintains cargo segregation acquired upon exit from the sorting endosome. Maturation of autophagosomes and endosomes: a key role for Rab7.
Mol Biol Cell, 2016;27(1):108–26. Biochim Biophys Acta, 2013;1833(3):503–10.
57. Naslavsky, N., Boehm, M., Backlund, P.S., Jr., and Caplan, S. Rabenosyn‐5 86. Ravikumar, B., Moreau, K., Jahreiss, L., Puri, C., and Rubinsztein, D.C.
and EHD1 interact and sequentially regulate protein recycling to the plasma Plasma membrane contributes to the formation of pre‐autophagosomal
membrane. Mol Biol Cell, 2004;15(5):2410–22. structures. Nat Cell Biol, 2010;12(8):747–57.
58. Naslavsky, N., Rahajeng, J., Sharma, M., Jovic, M., and Caplan, S. 87. Szatmari, Z., Kis, V., Lippai, M., et al. Rab11 facilitates cross‐talk between
Interactions between EHD proteins and Rab11‐FIP2: a role for EHD3 in autophagy and endosomal pathway through regulation of Hook localiza-
early endosomal transport. Mol Biol Cell, 2006;17(1):163–77. tion. Mol Biol Cell, 2014;25(4):522–31.
59. Sharma, M., Naslavsky, N., and Caplan, S. A role for EHD4 in the regulation 88. Fader, C.M. and Colombo, M.I. Autophagy and multivesicular bodies: two
of early endosomal transport. Traffic, 2008;9(6):995–1018. closely related partners. Cell Death Differ, 2009;16(1):70–8.
60. Zhang, J., Reiling, C., Reinecke, J.B. et  al. Rabankyrin‐5 interacts with 89. Madrigal‐Matute, J. and Cuervo, A.M. Regulation of liver metabolism by
EHD1 and Vps26 to regulate endocytic trafficking and retromer function. autophagy. Gastroenterology, 2016;150(2):328–39.
Traffic, 2012;13(5):745–57. 90. Tasset, I. and Cuervo, A.M. Role of chaperone‐mediated autophagy in
61. Duleh, S.N. and Welch, M.D. WASH and the Arp2/3 complex regulate metabolism. FEBS J, 2016;283(13):2403–13.
­endosome shape and trafficking. Cytoskeleton, 2010;67(3):193–206. 91. Schneider, J.L., Suh, Y., and Cuervo, A.M. Deficient chaperone‐mediated
62. Giridharan, S.S., Cai, B., Vitale, N., Naslavsky, N., and Caplan, S. Cooperation autophagy in liver leads to metabolic dysregulation. Cell Metab,
of MICAL‐L1, syndapin2, and phosphatidic acid in tubular recycling endo- 2014;20(3):417–32.
some biogenesis. Mol Biol Cell, 2013;24(11):1776–90, S1–15. 92. Chao, X., Wang, H., Jaeschke, H., and Ding, W.X. Role and mechanisms
63. Zhang, J., Naslavsky, N., and Caplan, S. EHDs meet the retromer: complex of  autophagy in acetaminophen‐induced liver injury. Liver Int,
regulation of retrograde transport. Cell Logist, 2012;2(3):161–5. 2018;38(8):1363–74.
64. Gokool, S., Tattersall, D., and Seaman, M.N. EHD1 interacts with retromer 93. Ni, H.M., Bockus, A., Boggess, N., Jaeschke, H., and Ding, W.X. Activation
to stabilize SNX1 tubules and facilitate endosome‐to‐Golgi retrieval. Traffic, of autophagy protects against acetaminophen‐induced hepatotoxicity.
2007;8(12):1873–86. Hepatology, 2012;55(1):222–32.
65. Scott, C.C., Vacca, F., and Gruenberg, J. Endosome maturation, transport and 94. Gomes, L.C., Di Benedetto, G., and Scorrano, L. During autophagy mito-
functions. Semin Cell Dev Biol, 2014;31:2–10. chondria elongate, are spared from degradation and sustain cell viability.
66. Hessvik, N.P. and Llorente, A. Current knowledge on exosome biogenesis Nat Cell Biol, 2011;13(5):589–98.
and release. Cell Mol Life Sci, 2018;75(2):193–208. 95. Lazarou, M., Sliter, D.A., Kane, L.A. et al. The ubiquitin kinase PINK1
67. Schmidt, O. and Teis, D. The ESCRT machinery. Curr Biol, 2012;22(4): recruits autophagy receptors to induce mitophagy. Nature, 2015;
R116–20. 524(7565):309–14.
68. Carlson, L.A., Shen, Q.T., Pavlin, M.R., and Hurley, J.H. ESCRT filaments 96. Harper, J.W., Ordureau, A., and Heo, J.M. Building and decoding ubiquitin
as spiral springs. Dev Cell, 2015;35(4):397–8. chains for mitophagy. Nat Rev Mol Cell Biol, 2018;19(2):93–108.
69. Frankel, E.B. and Audhya, A. ESCRT‐dependent cargo sorting at multive- 97. Hammerling, B.C., Najor, R.H., Cortez, M.Q. et  al. A Rab5 endosomal
sicular endosomes. Semin Cell Dev Biol, 2018;74:4–10. pathway mediates Parkin‐dependent mitochondrial clearance. Nat
70. Shen, J., Huang, C.K., Yu, H. et al. The role of exosomes in hepatitis, liver Commun, 2017;8:14050.
cirrhosis and hepatocellular carcinoma. J Cell Mol Med, 2017;21(5): 98. Yamano, K., Wang, C., Sarraf, S.A. et al. Endosomal Rab cycles regulate
986–92. Parkin‐mediated mitophagy. eLife, 2018;7.
71. Maji, S., Matsuda, A., Yan, I.K., Parasramka, M., and Patel, T. Extracellular 99. Wong, Y.C., Ysselstein, D., and Krainc, D. Mitochondria‐lysosome con-
vesicles in liver diseases. Am J Physiol Gastrointestinal Liver Physiol, tacts regulate mitochondrial fission via RAB7 GTP hydrolysis. Nature,
2017;312(3):G194–G200. 2018;554(7692):382–6.
72. Cornu M., de Caudron de Coquereaumont, G., and Hall, M.N. mTOR signal- 100. Brunt, E.M., Wong, V.W., Nobili, V. et al. Nonalcoholic fatty liver disease.
ing in liver disease, in Signaling Pathways in Liver Diseases 3rd edn (eds. J.‐F. Nat Rev Dis Primers, 2015;1:15080.
Dufour and P.‐A. Clavien), John Wiley & Sons, Chichester, 2015, pp. 314–25. 101. Spengler, E.K. and Loomba, R. Recommendations for diagnosis, referral
73. Settembre, C., De Cegli, R., Mansueto, G. et al. TFEB controls cellular lipid for liver biopsy, and treatment of nonalcoholic fatty liver disease and non-
metabolism through a starvation‐induced autoregulatory loop. Nat Cell Biol, alcoholic steatohepatitis. Mayo Clin Proc, 2015;90(9):1233–46.
2013;15(6):647–58. 102. Thiam, A.R., Farese, R.V., Jr., and Walther, T.C. The biophysics and cell
74. Napolitano, G. and Ballabio, A. TFEB at a glance. J Cell Sci, biology of lipid droplets. Nat Rev Mol Cell Biol, 2013;14(12):775–86.
2016;129(13):2475–81. 103. Walther, T.C. and Farese, R.V., Jr. Lipid droplets and cellular lipid metabo-
75. Gissen, P. and Arias, I.M. Structural and functional hepatocyte polarity and lism. Annu Rev Biochem, 2012;81:687–714.
liver disease. J Hepatol, 2015;63(4):1023–37. 104. Guo, Y., Cordes, K.R., Farese, R.V., Jr., and Walther, T.C. Lipid droplets at
76. Sewell, R.B. et al. Pharmacologic perturbation of rat liver lysosomes: effects a glance. J Cell Sci, 2009;122(Pt 6):749–52.
on release of lysosomal enzymes and of lipids into bile. Gastroenterology, 105. Schmidt, C., Ploier, B., Koch, B., and Daum, G. Analysis of yeast lipid
1988;95(4):1088–98. droplet proteome and lipidome. Meth Cell Biol, 2013;116:15–37.
77. Tooze, S.A., Abada, A., and Elazar, Z. Endocytosis and autophagy: exploita- 106. Zhang, H., Wang, Y., Li, J. et al. Proteome of skeletal muscle lipid droplet
tion or cooperation? Cold Spring Harbor Perspect Biol, 2014;6(5):a018358. reveals association with mitochondria and apolipoprotein a‐I. J Proteome
78. Mizushima, N. Autophagy: process and function. Genes Dev, 2007;21(22): Res, 2011;10(10):4757–68.
2861–73. 107. Sato, S., Fukasawa, M., Yamakawa, Y. et  al. Proteomic profiling of lipid
79. Czaja, M.J., Ding, W.X., Donohue, T.M., Jr. et al. Functions of autophagy in droplet proteins in hepatoma cell lines expressing hepatitis C virus core
normal and diseased liver. Autophagy, 2013;9(8):1131–58. protein. J Biochem, 2006;139(5):921–30.
80. Alers, S., Loffler, A.S., Wesselborg, S., and Stork, B. Role of AMPK‐mTOR‐ 108. Cermelli, S., Guo, Y., Gross, S.P., and Welte, M.A. The lipid‐droplet pro-
Ulk1/2 in the regulation of autophagy: cross talk, shortcuts, and feedbacks. teome reveals that droplets are a protein‐storage depot. Curr Biol,
Mol Cell Biol, 2012;32(1):2–11. 2006;16(18):1783–95.
81. Mercer, T.J., Gubas, A., and Tooze, S.A. A molecular perspective of mam- 109. Brasaemle, D.L., Dolios, G., Shapiro, L., and Wang, R. Proteomic analysis
malian autophagosome biogenesis. J Biol Chem, 2018;293(15):5386–95. of proteins associated with lipid droplets of basal and lipolytically stimu-
82. Lamb, C.A., Yoshimori, T., and Tooze, S.A. The autophagosome: origins lated 3T3‐L1 adipocytes. J Biol Chem, 2004;279(45):46835–42.
unknown, biogenesis complex. Nat Rev Mol Cell Biol, 2013;14(12):759–74. 110. Schott, M.B., Rasineni, K., Weller, S.G. et  al. β‐Adrenergic induction of
83. Reggiori, F. and Klionsky, D.J. Autophagosomes: biogenesis from scratch? lipolysis in hepatocytes is inhibited by ethanol exposure. J Biol Chem,
Curr Opin Cell Biol, 2005;17(4):415–22. 2017;292(28):11815–28.
74 THE LIVER:  REFERENCES

111. Ong, K.T., Mashek, M.T., Bu, S.Y., and Mashek, D.G. Hepatic ATGL 122. Martinez‐Lopez, N., Garcia‐Macia, M., Sahu, S. et  al. Autophagy in the
knockdown uncouples glucose intolerance from liver TAG accumulation. CNS and periphery coordinate lipophagy and lipolysis in the brown a­ dipose
FASEB J, 2013;27(1):313–21. tissue and liver. Cell Metab, 2016;23(1):113–27.
112. Ong, K.T., Mashek, M.T., Bu, S.Y., Greenberg, A.S., and Mashek, D.G. 123. Sathyanarayan, A., Mashek, M.T., and Mashek, D.G. ATGL promotes
Adipose triglyceride lipase is a major hepatic lipase that regulates triacylglyc- autophagy/lipophagy via SIRT1 to control hepatic lipid droplet catabolism.
erol turnover and fatty acid signaling and partitioning. Hepatology, Cell Rep, 2017;19(1):1–9.
2011;53(1):116–26. 124. Khan, S.A., Sathyanarayan, A., Mashek, M.T. et  al. ATGL‐catalyzed
113. Wu, J.W., Wang, S.P., Alvarez, F. et al. Deficiency of liver adipose triglyc- ­lipolysis regulates SIRT1 to control PGC‐1alpha/PPAR‐alpha signaling.
eride lipase in mice causes progressive hepatic steatosis. Hepatology, Diabetes, 2015;64(2):418–26.
2011;54(1):122–32. 125. Inoue, J., Ninomiya, M., Shimosegawa, T., and McNiven, M.A. Cellular
114. Reid, B.N., Ables, G.P., Otlivanchik, O.A. et al. Hepatic overexpression of membrane trafficking machineries used by the hepatitis viruses.
hormone‐sensitive lipase and adipose triglyceride lipase promotes fatty Hepatology, 2018;68(2):751–76.
acid oxidation, stimulates direct release of free fatty acids, and ameliorates 126. Macovei, A. et al. Regulation of hepatitis B virus infection by Rab5, Rab7,
steatosis. J Biol Chem, 2008;283(19):13087–99. and the endolysosomal compartment. J Virol, 2013;87(11):6415–27.
115. Schulze, R.J., Drizyte, K., Casey, C.A., and McNiven, M.A. Hepatic 127. Coller, K.E. et al. RNA interference and single particle tracking analysis of
lipophagy: new insights into autophagic catabolism of lipid droplets in the hepatitis C virus endocytosis. PLoS Pathog, 2009;5(12):e1000702.
liver. Hepatol Commun, 2017;1(5):359–69. 128. Selzer, L. and Zlotnick, A. Assembly and release of hepatitis B virus. Cold
116. Singh, R. and Cuervo, A.M. Lipophagy: connecting autophagy and lipid Spring Harb Perspect Med, 2015 Nov 9;5(12):pii: a021394.
metabolism. Int J Cell Biol, 2012;2012:282041. 129. Miyanari, Y. et al. The lipid droplet is an important organelle for hepatitis C
117. Singh, R., Kaushik, S., Wang, Y. et al. Autophagy regulates lipid metabo- virus production. Nat Cell Biol, 2007;9:1089–97
lism. Nature, 2009;458(7242):1131–5. 130. Salloum, S. et al. Rab18 binds to hepatitis C virus NS5A and promotes
118. Li, Z., Schulze, R.J., Weller, S.G. et  al. A novel Rab10‐EHBP1‐EHD2 interaction between sites of viral replication and lipid droplets. PLoS
­complex essential for the autophagic engulfment of lipid droplets. Sci Adv, Pathog, 2013;9(8):e1003513.
2016;2(12):e1601470. 131. Ploen, D. et al. TIP47 plays a crucial role in the life cycle of hepatitis C
119. Schroeder, B., Schulze, R.J., Weller, S.G. et al. The small GTPase Rab7 as a cen- virus. J Hepatol, 2013;58(6):1081–8.
tral regulator of hepatocellular lipophagy. Hepatology, 2015;61(6):1896–907. 132. Coller, K.E. et al. Molecular determinants and dynamics of hepatitis C
120. Schulze, R.J., Rasineni, K., Weller, S.G. et  al. Ethanol exposure inhibits virus secretion. PLoS Pathog, 2012;8(1):e1002466.
hepatocyte lipophagy by inactivating the small guanosine triphosphatase 133. Inoue, J., Krueger, E.W., Chen, J. et al. HBV secretion is regulated through
Rab7. Hepatol Commun, 2017;1:140–52. the activation of endocytic and autophagic compartments mediated by
121. Kaushik, S. and Cuervo, A.M. Degradation of lipid droplet‐associated pro- Rab7 stimulation. J Cell Sci, 2015;128(9):1696–706.
teins by chaperone‐mediated autophagy facilitates lipolysis. Nat Cell Biol, 134. Ostrowski, M. et al. Rab27a and Rab27b control different steps of the exo-
2015;17(6):759–70. some secretion pathway. Nat Cell Biol, 2010;12:19–30.
The Hepatocellular
7 Secretory Pathway
Catherine L. Jackson1 and Mark A. McNiven2
1
Institut Jacques Monod, UMR7592 CNRS Université Paris‐Diderot, Sorbonne Paris Cité, Paris, France
2
Department of Biochemistry and Molecular Biology, and The Center for Digestive Diseases, Mayo Clinic &
Foundation, Rochester, MN, USA

INTRODUCTION sorting station of the pathway [2]. Generally distributed


­throughout the cell and extending to the periphery, the ER is
Hepatocytes are known to secrete scores of proteins and lipid characterized by a network of interconnected membrane struc­
particles into the sinusoidal space, and also play a key role in tures that include cisternae and tubular networks [3, 4]. The
lipid homeostasis in organisms. An emerging theme over the past Golgi apparatus, in contrast, is localized in the perinuclear
few years has been the intimate connection between membrane region of the cell and forms a continuous ribbon in mammalian
trafficking and lipid metabolism, with important implications for cells [5]. This ribbon is composed of organized stacks of flat­
hepatocyte function. The secretory pathway assures processing tened saccules, interconnected by more complex tubulo‐vesicu­
and delivery of proteins to their correct destination, relying on lar regions [5]. The first cis‐Golgi element and the trans‐Golgi
highly organized vesicular trafficking machinery that includes network (TGN) are tubular membrane meshworks at the entry
numerous enzymes, cytoskeletal proteins, molecular motors, and and exit sides of the Golgi apparatus, respectively.
coat proteins. This machinery is highly conserved in evolution, Early studies of the secretory pathway made use of rat liver
and is found in virtually all eukaryotic cells. Specialized cell (Figure 7.1), as well as exocrine pancreas cells, due to their high
types such as hepatocytes can have more specific secretion levels of secretion [6]. These studies determined that during the
­pathways. Very low density lipoprotein (VLDL) particles are synthesis and transport of secreted proteins, the ER and Golgi
produced uniquely by liver cells, and use the ubiquitous vesicu­ function sequentially, and prior to packaging of secretory cargo
lar trafficking machinery in addition to components specific to into secretory granules or vesicles.
this large cargo. Through conventional biochemical and molecu­ The mechanism by which proteins are transported from a
lar methods, the use of genetic model organisms, and recent donor to an acceptor compartment in eukaryotic cells is via the
advancements in live cell imaging, much has been learned about budding and fusion of vesicular carriers [7] (Figure  7.2). The
these trafficking pathways in hepatocytes and other epithelial first step in formation of a vesicle is activation of a small G pro­
cells. This chapter will focus on the molecular mechanisms by tein of the ADP ribosylation factor (Arf) family by a nucleotide
which nascent proteins and larger cargos are sequestered, pack­ exchange factor, which leads to recruitment of proteins called
aged into vesicle carriers, and targeted to specific hepatocellular effectors to the membrane [10–12]. Cargo adaptors, which are
destinations during the secretory process. coupled to a vesicle coat, are a major class of effectors of the
Sar1 and Arf1 small G proteins that function at the ER and Golgi,
respectively. Both Arf and Sar proteins have an N‐terminal
amphipathic helix that inserts into the membrane, contributing to
THE SECRETORY PATHWAY deformation into a highly curved bud [13]. A variety of different
adaptors bind to the cargo, either transmembrane or luminal pro­
The secretory pathway begins with synthesis of proteins at the teins, to concentrate them into the forming vesicle [14]. Other
endoplasmic reticulum (ER) [1]. Next, proteins are transported effectors of Sar1 and Arf1 include lipid‐modifying enzymes,
to the Golgi apparatus (Figure 7.1), the central processing and which change the composition of the membrane at the budding

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
76 THE LIVER:  EVOLUTIONARY ORIGINS OF VESICLE COATS

membrane proteins and lysosomal enzymes pass through the


secretory apparatus to reach their final destinations, which
reflects the important sorting role of the Golgi apparatus.
A particularly active sorting region of the Golgi is the TGN,
where several distinct types of vesicles containing specific cargo
proteins and harboring particular coat proteins arise [22–24].
Several of these coats contain cargo adaptor proteins (APs)
recruited to membranes by the Arf1 or Arf3 small G proteins.
The APs are four‐subunit complexes containing two large and
two small subunits. AP‐1, and probably AP‐3 as well, recruit
clathrin upon membrane binding to form a clathrin–adaptor
coat. However, this is not the case for AP‐4, which forms a non‐
clathrin coat [22, 23]. Another important set of adaptors at the
TGN are the Golgi‐localized γ‐ear‐containing, ADP ribosyla­
tion factor‐binding (GGA) proteins (GGA1, GGA2, and
GGA3), which bind clathrin and transport distinct cargo from
the TGN to endosomes. Secretory vesicles (or granules) carry­
ing constitutively secreted proteins from the TGN to the cell
surface are believed to be uncoated vesicles, as no coats have yet
been identified in mammalian cells [23].
Figure 7.1  Elaborate vesicular processes in the hepatocyte. Thin‐­
section electron micrograph of a rat liver hepatocyte from the original
collections of, Dr. Keith Porter, about 1962. Deep in the cytoplasm,
parallel cisternae of the endoplasmic reticulum (ER) appear to give rise
to the flattened, stacked saccules of the Golgi apparatus (G). Scores of
EVOLUTIONARY ORIGINS
small vesicular profiles can be seen in the Golgi region, as well as larger OF VESICLE COATS
lipoprotein‐filled secretory vesicles budding from the trans side of the
Golgi. Specific coats, accessory proteins, and address tags characterize The COPII, COPI, and AP–clathrin coats are all thought to
many of these vesicles. G, Golgi; ER, endoplasmic reticulum; Lys, share a common origin, arising from a primordial coat com­
lysosome.
plex. All three vesicle coats have subunits containing a β‐pro­
peller domain followed by an α‐solenoid, also known as a
site, also contributing to membrane deformation [15–17]. helix‐turn‐helix domain [8] (Figure  7.2). Strikingly, only
Polymerization of the outer shell of the coat also aids in shaping eukaryotes, and a few rare prokaryotes with rudimentary vesic­
the membrane into a highly curved bud. ular trafficking pathways, have proteins containing a β‐propel­
Once the bud has formed and undergone scission from the ler motif fused to an α‐solenoid fold [8]. Hence this class of
donor compartment membrane, uncoating of the vesicle ensues. protein is strictly correlated in evolution with the necessity to
The first step is hydrolysis of GTP on the Arf family small G bend membranes, such as in vesicle formation. This β‐
protein, which can commence even before the vesicle has under­ propeller–α‐solenoid structure is also found in nuclear pore
gone fission from the donor compartment [18]. The interaction components that bind to the highly curved membrane of the
between vesicle cargo and the coat maintains the coat on the pore [25]. These observations led to the protocoatomer hypoth­
vesicle membrane until the appropriate time for uncoating, esis, which postulates that COPII, COPI, and clathrin–AP
which must occur prior to fusion of the vesicle with its target coats, as well as some nucleoporins, all shared a primordial
membrane. ancestor [25]. Additional support for this hypothesis comes
Transport in the anterograde direction from the ER to the from clear sequence homology between the β‐ and γ‐COP sub­
Golgi apparatus is mediated by coat protomer II (COPII)‐coated units of COPI and the large subunits of the AP‐1 and AP‐2
vesicles (Figure  7.2). On the ER membrane, subunits of the adaptors [8, 26, 27]. However, despite this homology, the
COPII coat are recruited upon activation of Sar1 by its exchange structures of the COPI and AP–clathrin coats are significantly
factor Sec12, to form buds and ultimately generate vesicles. different (Figure 7.2). The β‐propeller–α‐solenoid subunits of
COPII vesicle formation occurs at specialized regions of the ER both coats polymerize due to interactions between α‐solenoids,
that are called ER exit sites (ERES). In mammalian cells, COPII but in the case of COPI, these subunits make direct contact
vesicles undergo homotypic fusion, and the resulting compart­ with the membrane, and are positioned beside the other sub­
ment recruits exchange factor Golgi brefeldin A resistant factor 1 complex of COPI. In contrast, in the AP–clathrin coat, clathrin
(GBF1), its substrate Arf1, and the coat protomer I (COPI) coat, forms a cage around the vesicle that does not directly touch the
which form COPI vesicles (Figure  7.2). The sorting compart­ membrane (Figure 7.2a). COPII resembles AP–clathrin in that
ment formed upon GBF1–Arf1–COPI recruitment, called the the β‐propeller–α‐solenoid subunit (Sec31), along with Sec13,
ER/Golgi intermediate compartment (ERGIC), is made up of a forms an outer cage that does not directly contact the mem­
complex network of tubules [19]. COPI‐coated vesicles are also brane (Figure 7.2a). However, the geometry of the COPII and
thought to mediate the retrograde transport of resident ER pro­ AP–clathrin coats is quite different based on in vitro structural
teins and Golgi enzymes from the Golgi apparatus back to the data (Figure  7.2a), and visualization of vesicles within cells
ERGIC and to the ER [20, 21]. In addition to secretory proteins, (Figure 7.2b).
7:  The Hepatocellular Secretory Pathway 77

(a) β-propeller
α-solenoid
small GTPase
Sec23
Sec24
AP/COP large subunit (β)
AP/COP large subunit (EGADZ) Clathrin
AP/COP medium subunit
AP/COP small subunit
Transmembrane cargo

Sec13/Sec31 α-COP β’-COP

COPII COPI AP-1 + clathrin

(b)
COPII COPI Clathrin
A B C

D E F

Figure 7.2  Comparison of the COPII, COPI, and AP–clathrin coats. (a) All three coats contain β‐propeller–α‐solenoid components (red spheres
and blue rods), but they are arranged in different ways in each coat. In COPII and AP–clathrin coats, these subunits form an outer polyhedral layer.
In COPI, on the other hand, the β‐propeller–α‐solenoid‐containing subunits directly contact the membrane as well as the other COPI subunits. The
other subunits of COPI and the adaptor subunits share both sequence and structural homology. All three coats are recruited to membranes by a
member of the Arf small G protein (GTPase) family: Sar1 in the case of COPII, and Arf1 for both COPI and AP–clathrin coats. Reproduced from
[8] with permission of Elsevier. (b) Visualization in their native cellular environment of COPII (A,D), COPI (B,E) and AP‐1–clathrin (C,F)‐coated
vesicles by cryoelectron tomography. Imaging was performed on Chlamydomonas reinhardtii cells. Panels A–C show cross‐sections through each
vesicle; panels D–F are grazing slices through the coats at the tops of the same vesicles. Note that the geometry of the COPII and clathrin coats is
visible in the lower panels, revealing the triangular Sec13/31 COPII lattice (D) and clathrin triskelions (F). Scale bars, 50 nm. Reproduced from
Bykov 2017 [9], https://cdn.elifesciences.org/articles/32493/elife‐32493‐v2.pdf. Licensed under CCBY 4.0.
78 THE LIVER:  COPI‐COATED VESICLES

In addition to β‐propeller–α‐solenoid subunits, coats also and two members of the BIG/Sec7 subfamily, brefeldin A inhib­
share similar features such as recruitment to membranes by a ited GEF 1 (BIG1) and BIG2. These two subfamilies have a
member of the Arf family of small G proteins, as described similar domain structure, probably because they have descended
above. Other ancestral motifs found in all coat complexes are from a common ancestor, but have different steady‐state locali­
longin domains (present in the medium and small COPI/AP zations and functions [10, 30, 31]. In animal cells, including
subunits, Figure  7.2a) and coiled coils [8]. This ensemble of hepatocytes, and yeast, the major steady‐state localization
ancient conserved features supports the idea of an ancient pri­ of  GBF/Gea GEFs is at the early Golgi, whereas BIG/Sec7
mordial coat, which split into the three major coats of the secre­ GEFs are present predominantly at the late Golgi, ­including
tory pathway: COPII, COPI, and AP–clathrin. the TGN [10, 31].
COPI vesicles mediate retrograde transport of components
that need to be recycled from the Golgi and ERGIC back to
earlier compartments and to the ER. Arf proteins (primarily
COPI‐COATED VESICLES Arf1 and Arf3) at the TGN recruit the heterotetrameric clathrin
adaptor proteins, AP‐1, AP‐3, and AP‐4, and the three mono­
COPI‐coated vesicles are required for intra‐Golgi trafficking meric Golgi‐localized γ‐ear‐containing, ADP ribosylation fac­
and for recycling from the Golgi apparatus back to the ER. tor‐binding proteins (GGAs 1–3) [24, 38]. To explain how one
The COPI coat consists of seven subunits (α, β, β′, γ, δ, ε, ζ or two Arf proteins can recruit so many different coats to differ­
COP) that form a 680 kDa complex and associate with the small ent membrane sites, several groups have found evidence that the
GTP‐binding protein Arf1. COPI sorts cargo proteins into form­ GEFs activating Arfs at different sites determine the coat that
ing vesicles as it induces membrane curvature [28] (Figure 7.2). will be recruited. GBF1 at ERGIC and the cis‐Golgi is responsi­
COPI‐coated vesicles are best known for their function in the ble for COPI recruitment, and the BIG1 and BIG2 proteins at
retrieval of ER‐resident proteins from the Golgi. Selected COPI the trans‐Golgi are responsible for recruiting the AP–clathrin
subunits are known to associate with dilysine amino acid motifs coats [10]. Demonstration of a direct interaction between GBF1
in the cytoplasmic domains of ER‐resident enzymes [20]. and COPI coat subunits provides a molecular explanation for
Furthermore, several COPI temperature‐sensitive mutants in yeast this specificity [39]. Interaction of GBF1 with COPI prior to
have revealed defects in transport between the ER and the Golgi at Arf1 activation ensures that COPI is stabilized on the membrane
the restrictive temperature [21]. Further, COPI‐coated vesicles where GBF1 activates Arf1 [39].
have been shown to mediate retrograde transport of resident ER Removal of the coat from a vesicle (uncoating) is required for
proteins by using the KDEL (single‐letter code for amino acids) fusion of the vesicle to the acceptor membrane. The first step in
receptor [29], confirming a retrieval role for COPI coat proteins. vesicle uncoating occurs upon hydrolysis of GTP on the Arf
Regulators of Arf–GTP binding and GTP hydrolysis control the family G protein that maintains the coat on the membrane [40,
spatio‐temporal localization of Arf protein activation and down­ 41]. Even after hydrolysis of GTP on Sar1 or Arf1, the coat (or
stream signaling. GDP release from Arf proteins is catalyzed by at least a portion of the coat) remains on the vesicle due to inter­
Arf guanine nucleotide exchange factors (GEFs), allowing the actions with other vesicle proteins such as cargo molecules [40,
more abundant GTP to bind. This nucleotide exchange activity is 42]. In the case of COPI vesicles, GTP hydrolysis on Arf1 is
contained within the Sec7 domain, an evolutionarily conserved mediated by Arf GAP proteins [28, 43, 44]. Arf GAP1 is
sequence first identified in yeast Sec7p [30–32]. A Sec7 domain is recruited only to highly curved membranes through its ALPS
present in all Arf GEFs identified to date. The function of Sec7 motif, which ensures that the coat remains on the budding vesi­
domains was first identified in the yeast Gea1p protein [33]. The cle until it has achieved a mature, highly curved morphology
human ortholog of yeast Gea1p, GBF1, was identified by P. [45, 46]. It was originally thought that the coat was lost quite
Melançon and colleagues [34]. GBF1 is a Golgi‐localized protein quickly after budding. However, interaction of the coat with a
whose overexpression was found to confer resistance to the fungal tethering complex on the acceptor membrane has recently been
toxin brefeldin A, which completely blocks secretion, and causes demonstrated for multiple types of vesicles [47, 48]. For COPI
disassembly of the Golgi apparatus and eventually its fusion with vesicles transported to the ER, the Ds11 complex present on the
the ER [10]. The Arf GTPase‐activating proteins (GAPs) catalyze ER membrane acts as a vesicle docking site through direct inter­
the hydrolysis of GTP on Arf family proteins, a function carried action with the COPI coat [49–51]. Moreover, this coat–tether
out by a conserved GAP domain, which contains a zinc finger [10]. interaction stimulates uncoating of the vesicle [51]. In addition
The primary sequence homology of the catalytic domains of the to fusing with the ER, COPI vesicles also mediate intra‐Golgi
Arf GEFs and GAPs accelerated their identification and allowed retrograde trafficking. Interactions between COPI subunits and
studies of their evolution. Phylogenetic analyses of the Arf GAPs the Golgi COG tethering complex [52, 53], and with the cis‐
has revealed that they have likely co‐evolved with their Arf sub­ Golgi tether p115 [54], are also likely to be involved in targeting
strates [35], and a similar conclusion very likely holds for the Arf vesicles to these Golgi acceptor membranes.
GEFs [36]. From these results we can conclude that Arf proteins GBF1 has been shown to have a crucial function in the
function in a tightly coordinated manner with their regulators. ­replication of numerous viruses, including hepatitis C [55] and
There are seven subfamilies of Arf GEFs in eukaryotic cells hepatitis E [56]. Hepatitis A, B, C, D, and E viruses cause the
[30]. Members of the GBF/Gea and BIG/Sec7 subfamilies of majority of liver disease across the globe, and represent a world­
Sec7 domain GEFs use class I Arf proteins as substrates [10], wide health problem. For the well‐studied hepatitis C virus, lipid
and in addition, GBF1 possibly functions on class II Arf pro­ metabolism is an important component in infection, and lipid
teins [34, 37]. In humans, there is one GBF/Gea member, GBF1, metabolism pathways are subverted for propagation of this virus
7:  The Hepatocellular Secretory Pathway 79

[57]. The catalytic domain of GBF1 is required for HCV to form Sec13/31, and Sec23/24 were found to be necessary for the
functional viral replication complexes [58]. Arf4 and Arf5 are generation of COPII‐coated vesicles from washed micro­
also required for HCV replication, through their roles in lipid somes [62]. Sec12, a GTP‐GEF for Sar1 that is localized to
homeostasis in hepatocytes [58]. Therefore, understanding the the ER initiates COPII vesicle biogenesis. After the activa­
mechanisms by which hepatitis viruses infect hepatocytes could tion of Sar1, the adaptor coat components Sec23/24 are
increase our knowledge of how the secretory pathway and lipid recruited to bind cargo. Subsequently, the structural β‐
homeostasis are coordinated in cells, in addition to providing propeller–α‐solenoid Sec13/31 coats are recruited to form
avenues for treatment of these important human pathogens. the coat and to contribute to membrane deformation. The
selection of cargo including transmembrane proteins and v‐
SNAREs (soluble NSF attachment receptors) by Sec23/24 is
mediated by three different sites for recognizing ER‐to‐Golgi
COPII‐COATED VESICLES v‐SNAREs on Sec24. The interaction is regulated by the
assembly state of the SNAREs, which suggests that COPII
COPII‐coated vesicles function in ER‐to‐Golgi proteins can be involved in vesicle fusion specificity in the
anterograde transport ER‐to‐Golgi step [12].
In the final stages of COPII vesicle biogenesis, the GTP
Protein transport from the ER to the Golgi complex is mediated used during vesicle formation is hydrolyzed by Sec23, a Sar1‐
by vesicular carriers that originate from morphologically specific GAP. Upon GTP binding by Sar1, membrane inser­
defined, specialized ribosome‐free regions of the ER called ER tion of the N‐terminal α helix of Sar1 deforms synthetic
exit sites (or ER export sites) (ERES) [59]. ER buds and an liposomes into narrow tubules. Mutation of the helix led to a
abundance of small vesicular profiles at these sites, which are defect in membrane curvature and the formation of vesicles
distinct from smooth ER could be clearly visualized in the clas­ from native ER, although the recruitment of coat proteins
sic electron micrographs of pancreatic acinar cells by Jamieson appeared to be normal [13]. Moreover, inhibition of GTP
and Palade [60]. Studies using live cell imaging demonstrated hydrolysis by Sar1 resulted in COPII‐coated vesicles that
that these transitional areas are formed transiently and randomly failed to detach from the ER, suggesting that regulation of the
from the ER [61]. N‐terminus of Sar1 by GTP binding and hydrolysis controls
The cargo carriers that mediate ER‐to‐Golgi protein transport COPII vesicle fission [65].
have been subjects of intense study. As for many of the protein Although the minimal requirements for in vitro vesicle bio­
trafficking steps along the secretory pathway, molecular compo­ genesis are the five molecules listed above, other components
nents were first identified in studies using the yeast Saccharomyces essential for COPII vesicle formation in different cell types have
cerevisiae as a model system. Yeast has provided a powerful been identified. Sec16 was first identified in yeast, and found to
combination of genetics, molecular biology, and biochemistry to be essential and required for COPII vesicle formation in vivo
elucidate mechanisms both in vitro and in vivo. Isolation of con­ [12]. Sec16 is present throughout eukaryotes, and is thought to
ditional sec (secretion) mutants allowed the identification and play a role in scaffolding other COPII components and organ­
characterization of numerous components of trafficking path­ izing ERES through interaction with other ERES components
ways, including all essential subunits of the COPII coat [32]. By [14]. Sec16 is also required for formation of stress granules con­
using yeast membranes for ER vesicle budding in vitro, cell‐free taining COPII components that are formed when cells are
assays for vesicle budding and fusion could be reconstituted. starved and consequently shut down the secretory pathway [66].
These systems enabled the isolation of ER‐derived transport ves­ Hence Sec16 has broad functions in regulation of COPII com­
icles that were 60–65 nm in diameter and exhibited a distinct ponents in various cellular physiological states, including cell
electron‐dense coat on their surface, termed COPII [62]. COPII growth and starvation.
coat components were identified from the purified vesicles and The TRK‐fused gene protein (TFK) and the cargo receptor
revealed to be distinct from those of COPI coats [62]. transport and Golgi organization 1 (TANGO1)/cutaneous T
Early studies in various cell types including human hepatoma cell lymphoma‐associated antigen (cTAGE5) proteins are
HepG2 cells [63] reported that secretory cargo protein was present more recently identified ERES components that function in
at bud sites upon ER exit and within small 40–80 nm carrier vesicles ERES organization [67, 68]. TFK binds to the Sec23/24 inner
and tubules referred to as vesicular–tubular clusters (VTCs). In pan­ coat and contributes to uncoating of COPII vesicles [67]. As
creatic cells, similar structures were shown to be COPII‐­positive. described for COPI vesicles, the uncoating of COPII vesicles
The ER cisternae near VTCs often exhibited budding profiles that also occurs late in a transport step, after targeting of the vesi­
were decorated with an electron‐dense, honeycomb‐patterned coat cle to the a­ cceptor compartment [47, 48]. Interestingly, TFK
(Figure 7.2b), which was identified as COPII by immunogold labe­ has an N‐­ terminally unstructured region in addition to a
ling antibodies against COPII coat components [64]. COPII‐binding site, which confers properties of liquid droplet
phase separation in vitro. In the absence of TFK, COPII vesi­
COPII‐coated vesicle protein components cles disperse, suggesting that TFK might function to maintain
vesicles close to their target membranes though forming a liq­
and vesicle formation
uid droplet ­structure between ERES and ERGIC target mem­
The COPII coat components, which are distinct from the branes [67]. In addition to TRK, the COPII vesicle coat
COPI coatomer, were postulated to induce vesicle budding of component Sec23 also interacts with the tethering complex
ER‐derived transport vesicles because only the GTPase Sar1, transport protein particle (TRAPP) at the acceptor
80 THE LIVER:  TRANS‐GOLGI NETWORK‐DERIVED VESICLES

compartment membrane prior to complete release of the TRANS‐GOLGI NETWORK‐DERIVED


COPII coat and fusion [69]. VESICLES
TANGO1 was identified in a screen for genes affecting secre­
tion in flies, and was later demonstrated to play an essential role
in collagen secretion in keratinocytes and fibroblasts [68]. Like
Protein sorting and trafficking events
TFK, TANGO1 interacts with the inner components of the at the trans‐Golgi network
COPII coat, Sec23 and Sec24, but not the outer cage compo­ The TGN is a tubulovesicular network that acts primarily in
nents, and localizes to ERES [68]. TANGO1 is also expressed in sorting proteins to their final destinations. Within this membra­
liver cells [70], where it functions in export of specific cargos, nous reticulum, cargo proteins are segregated efficiently into
as described below. distinct vesicles that will be targeted to various compartments,
In mammalian cells, GST–hSec23‐binding columns allowed including the endosome/lysosome system, the apical or basolat­
the identification of a 125 kDa protein, p125, that is expressed eral domains in polarized cells, and the secretory granule pool.
in liver, and shares sequence homology with p­ hosphatidic acid In regulated secretory cells, this granule pool awaits the appro­
(PA)‐preferring phospholipase A1, which produces lysophos­ priate extracellular stimulus to exocytose its contents.
pholipids [71]. Recently, the function of lysophospholipids in The TGN is known to interact intimately and dynamically
lowering the rigidity of the ER membrane during COPII with the endosomal apparatus, especially during events related
­vesicle budding was demonstrated both in vitro and in cells to receptor recycling, endosomal maturation, and the delivery of
[72]. These results indicate the importance of membrane lysosomal enzymes [23].
­com­position and membrane biophysical properties in vesicle A large number of different types of vesicles are formed at
formation. the trans‐Golgi and TGN, reflecting the intense sorting activity
that occurs at this exit face of the Golgi. The best characterized
are clathrin‐coated vesicles, which are linked to the membrane
LARGE COPII CARRIERS TRANSPORT through various APs, as described briefly above. Other non‐
clathrin‐coated vesicles are characterized by adaptor proteins
VLDL PARTICLES FROM THE ER
homologous to AP‐1, but which carry different cargo.
TO THE GOLGI
Recently, it has become clear that large cargos, too large to be Formation of clathrin‐coated and clathrin‐
incorporated into classic 60 nm COPII vesicles, nevertheless
independent vesicles at the TGN
use COPII for their transport out of the ER to the Golgi in the
secretory pathway [73, 74]. In hepatocytes, VLDL particles are Clathrin was the first coat to be visualized in the Golgi region of
formed from the ER membrane on the luminal side, and from mammalian cells. An electron‐dense, spiked coat on TGN mem­
there are transported to the Golgi apparatus [70, 75]. However, branes in liver hepatocytes could be seen clearly in the original
each VLDL particle is up to 90 nm in diameter, and especially micrographs of Novikoff and Yam [80] (Figure 7.3a). The first‐
if multiple particles enter the same carrier, this cargo is too identified and best‐characterized cargo adaptors are heterotrim­
large to be incorporated into classic COPII vesicles. However, eric adaptor protein complexes AP‐1 to AP‐5, three of which
there is evidence for COPII involvement in ER–Golgi transport mediate sorting at the TGN (AP‐1, AP‐3, and AP‐4). All three of
of VLDL [76, 77]. In addition, the TANGO proteins are impor­ these coats are recruited to membranes by Arf1 and/or Arf3.
tant components in this process, acting as scaffolds for the AP‐1 in addition requires the phosphoinositide phosphatidylin­
assembly of these larger COPII vesicular structures. TANGO1 ositol 4‐phosphate (PI(4)P) (Figure  7.4). TGN‐derived CCVs
was first identified as a transmembrane cargo receptor for are now known to have a coat composed of clathrin triskelia and
luminal procollagen, another cargo too large for classic COPII the AP complexes AP‐1 (made up of γ‐, β1‐, μ1‐, and σ1‐adaptin
vesicles. As a cargo receptor, TANGO1 was found to link lumi­ subunits) [22] and AP‐3 (δ‐, β3‐, μ3‐, and σ3‐adaptin subunits)
nal collagen to the COPII coat on the other side of the ER [81]. AP‐4 does not bind to clathrin. The subunit μ1 of AP‐1 is
membrane [68]. responsible for binding to YXXΦ motifs within cargo proteins,
TANGO1 and an interacting partner called TALI are both whereas β1 interacts with dileucine‐based signals in cargo. The
required for optimal secretion of VLDL particles in the hepato­ γ‐adaptin subunit recruits accessory proteins to the site of vesi­
cyte cell line HepG2 [70]. TANGO1 and TALI not only interact, cle formation [23].
but the complex interacts with apolipoprotein B (ApoB100), There are two forms of the μ1 subunit of AP‐1, μ1A and μ1B.
required for production of VLDL particles via its interaction μ1A is ubiquitous, whereas μ1B is found only in polarized epi­
with triglycerides [70]. All three proteins colocalize in HepG2 thelial cells [22, 23]. The two adaptor complexes AP‐1A and
cells, and interaction of TANGO/TALI with apoB present on AP‐1B, containing μ1A and μ1B, respectively, are both present
VLDL particles results in their recruitment to ER exit sites, in polarized cells, and function in basolateral protein sorting.
where they are packaged into COPII carriers [70]. Interestingly, AP‐1B is responsible for sorting of a specific subset of cargo
ERGIC53, a marker of ERGIC membranes, colocalizes with proteins at the TGN in epithelial cells to the basolateral plasma
apoB‐containing structures at ER exit sites [70]. This and other membrane, including the LDL receptor and interleukin 6 recep­
data support the conclusion that ERGIC is involved in forma­ tor β chain [82].
tion of VLDL particles, as it is for collagen‐containing COPII A second type of coat complex recruited to TGN membranes
carriers [78]. by Arf1 and PI(4)P is the family of monomeric adaptors, the
7:  The Hepatocellular Secretory Pathway 81

(a) (b)
VLDL transport vesicles
ER Golgi
VLDL apparatus
VLDL

Sar1b
Sec23/24
VLDL Sec13/31 VLDL
TANGO
VLDL

ApoB100 VLDL

Sar1a
Sec23/24
Sec13/31

Classic COPII vesicles

Figure 7.3  (a) VLDL particles are present in dilated elements (arrows) within the Golgi apparatus, and in secretory vesicles forming within the
trans‐Golgi and TGN. Rough ER is evident by the presence of ribosomes. G, Golgi; arrows, Golgi elements containing VLDL particles; tGE,
trans‐Golgi elements containing VLDL particles; C, clathrin‐coated vesicle; R, ribosomes; P, peroxisome. Magnification: 44,000×. Reproduced
with permission from [80]. ©1978 P.M. Novikoff and A. Yam. Originally published in Journal of Cell Biology. https://doi.org/10.1083/jcb.76.1.1
(b) Schematic diagram depicting VLDL trafficking from the ER to the Golgi in hepatocytes. VLDL particles form from the ER membrane on the
luminal side, and require apolipoprotein B 100 (apoB100). The particles bud into the ER lumen, then are packaged into VLDL transport vesicles
using the COPII machinery (Sec23/24, Sec13/31) as well as TANGO. The diameter of VLDL transport vesicles is approximately 110 nm [79], suf­
ficient to enclose VLDL‐sized particles, and larger than classic COPII vesicles (60–70 nm diameter on average). The latter contain nascent secretory
and transmembrane cargo proteins synthesized on ER‐localized ribosomes, and translocated into the ER lumen. Both VLDL transport vesicles and
classic COPII vesicles require COPII proteins for their budding from the ER membrane, but use different Sar1 small G proteins: Sar1b for VLDL
transport vesicles and Sar1a for classic COPII vesicles. Upon fission of the vesicle from the ER, both types of vesicles are targeted to the acceptor
compartment, uncoat, then fuse with the acceptor compartment membrane.

Basolateral membrane
6
11

Rab8
ab
ab

R
R

Golgi
4)P
PI( P2
)
,5
(4
PI b2
Ra BC
Rab9
LE/MVB 1
Rab
membrane
Apical

ER

Figure 7.4  Specific small Rab GTPases and phosphoinositides are localized to distinct membrane compartments along the hepatocyte secretory
pathway. Rab1 and Rab2 are believed to mediate the traffic of nascent proteins from the ER to the cis‐Golgi. Rab6 is a kinesin‐associated Rab
that mediates transport within the Golgi stacks, Rab11 regulates exit from the TGN, and Rab8 functions in the targeting of secretory vesicles from
the Golgi to the plasma membrane. Rab9 has a role in late endosome‐to‐TGN transport. PI(4)P is significantly enriched in the Golgi and is
believed to mediate the recruitment of multiple proteins that support vesicle formation at the TGN. PI(4,5)P2 at the Golgi functions with Arf1 and
phospholipase D (PLD) and is also involved in vesicle formation. LE/MVB, late endosome/multivesicular body; ER, endoplasmic reticulum; BC,
bile canalicular membrane.
82 THE LIVER:  COORDINATION BETWEEN THE SECRETORY PATHWAY AND LIPID METABOLISM

Golgi‐localized, γ‐ear‐containing, Arf‐binding proteins (GGAs). which indicates an essential role of the cortactin–dynamin com­
There are three known GGA proteins in mammalian cells: plex in TGN function [90].
GGA1, GGA2, and GGA3 [23]. Each GGA protein contains The dynamins are believed to form large helical polymers from
three domains: VHS (Vps27, Hrs, Stam), GAT (GGA and which many interactive proline‐rich tail domains extend. These
TOM), and γ‐adaptin ear (GAE). Through the VHS domain, the domains bind to a variety of SH3‐domain‐containing proteins,
GGAs recognize DXXLL motifs in specific cargo such as the many of which appear to be actin‐binding proteins such as Abp1
mannose 6‐phosphate receptor (MPR) tail and thus provide a and syndapin. These findings suggest that the dynamin family
sorting function between the TGN and endosomes [83]. The acts as a “polymeric contractile scaffold” at the interface between
GAT domain interacts with Arf; mutations in this domain abro­ biological membranes and filamentous actin (Figure 7.4).
gated recruitment of GGAs to the TGN as well as binding to Arf A number of non‐dynamin‐mediated mechanisms for scis­
[84]. The GAE domain is homologous to the GAE domain of sion of vesicles, including those formed at the TGN, have
the AP‐1 complex‐recruiting accessory proteins [85]. recently been described [91]. In the case of Rab6 vesicles
CCVs generated from the TGN are involved in the transport formed at the TGN, cytoskeletal motors play a crucial role.
of newly synthesized lysosomal enzymes. Newly synthesized, Active Rab6‐GTP recruits myosin II to membranes, where its
soluble lysosomal enzymes acquire mannose 6‐phosphate resi­ motor activity on filamentous actin creates the force necessary
dues at the cis‐Golgi, and these residues are recognized by the to mediate vesicle fission [92].
MPRs at the TGN [23]. The GGAs are essential for packaging
MPRs into CCVs and for transporting them from the TGN to
endosomes [86]. Another adaptor complex at the TGN, AP‐3, is
COORDINATION BETWEEN
expressed ubiquitously, is localized to buds/vesicles associated
with the TGN, and can interact with clathrin and with sorting THE SECRETORY PATHWAY
signals in the cytoplasmic tails of lysosomal membrane proteins AND LIPID METABOLISM
[23]. Correct targeting of lysosomal‐associated membrane pro­
tein‐1 (LAMP‐1) and lysosomal integral membrane protein‐2 The secretory pathway is intimately linked to lipid metabolism
(LIMP‐2) are mediated by the AP‐3 adaptor complex [81]. In pathways [93]. The ER is one of the major sites within the cell for
addition to the heterotetrameric and GGA adaptor complexes, lipid synthesis, with multiple pathways operating to shuttle lipids
additional adaptors have been identified, including epsin‐related to different destinations. Both the phospholipids that make up cel­
proteins [23]. lular membranes, and neutral storage lipids (triglycerides and cho­
In addition to Arf family small G proteins, another family of lesterol esters) are synthesized by enzymes in the ER membrane
small G proteins, called Rabs, are involved in each trafficking [94, 95]. In addition to their synthesis, neutral lipids are also
step of the secretory pathway [87]. Rab1, along with Rab2, are thought to be packaged into lipid droplets in the ER membrane
known to function in trafficking from the ER and ERGIC com­ [95]. A regulatory switch between channeling of fatty acids into
partments [59, 87]. Other Rab GTPases, including Rab6, either phospholipids for secretory membrane production or into
Rab19, Rab30, Rab33, Rab36, Rab40, Rab41, and Rab43 func­ neutral lipids for storage in lipid droplets (LDs) was first identified
tion at the Golgi, and Rab8 (the ortholog of Sec4 in yeast) is in yeast [96]. Subsequent work in mammalian cells has uncovered
crucial in the exocytic pathway for late Golgi to plasma mem­ similar regulatory mechanisms. A critical control point in this
brane transport. Rab3, Rab26, Ra27, Rab37, and Rab2 have switch is the metabolism of phosphatidic acid (PA), which serves
also been implicated in TGN to plasma membrane trafficking as a precursor of both phospholipids and storage triglycerides [97].
[87]. Rab8 and Rab11, along with Arf4 and Arf regulators, Indeed, PA can either be dephosphorylated by PA phosphatases to
function in trafficking from the TGN to cilia in photoreceptor produce diglyceride (DAG), which is then fatty acid acylated to
cells [88] (Figure 7.4). form triglycerides, or converted by CDP‐DAG synthase to CDP‐
DAG, a precursor of all the major cellular phospholipids. Hence
the action of PA phosphatases is responsible for channeling PA
Scission of TGN‐derived vesicles towards storage lipid synthesis. There is one PA phosphatase in
Since the localization of the large GTPase dynamin to the Golgi yeast, Pah1, and three in mammalian cells, lipin1, lipin2, and
apparatus, in addition to its known cell surface distribution, was lipin3. A lipin1‐knockout mouse has a lipodsystrophy phenotype,
reported, it has been suggested that vesicle fission at the Golgi and Pah1 is required for LD formation in yeast, showing the
apparatus is mediated by this molecular pinchase. Dynamin’s importance of these enzymes for lipid storage [97]. Pah1 and lipins
association with Golgi membranes was demonstrated bio­ are localized both to the nucleus and cytoplasm, and are involved
chemically with specific antibodies that detected dynamin in in transcriptional regulation of lipid metabolism genes [97].
immunoblots of purified rat liver Golgi fractions and by the In addition to ubiquitous regulatory pathways linking
immuno‐isolation of Golgi membranes on magnetic beads the  secretory pathway and lipid metabolism, hepatocytes also
coated with those dynamin antibodies [89]. Cortactin, an actin‐ use  the secretory pathway for transport of VLDL particles
binding protein, in a complex with dynamin, functions in post‐ (Figure 7.3a), a central function in organismal lipid homeostasis
Golgi transport in liver cells. By using in vitro or intact cell [75, 79, 98]. The liver plays a key role in the body both in stor­
experiments, it was shown that activation of Arf1 recruits actin, age of lipids and in their distribution to other tissues. In the liver,
cortactin, and dynamin to Golgi membranes. Disruption of cort­ fatty acids from the diet, from triglyceride breakdown, and from
actin–dynamin interactions reduced dynamin recruitment to the biosynthesis are incorporated into LDs for storage, and into
Golgi and blocked the transit of nascent proteins from the TGN, VLDL particles for distribution. Both types of particles have a
7:  The Hepatocellular Secretory Pathway 83

hydrophobic lipid core made up of triglycerides surrounded by REFERENCES


a phospholipid monolayer, likely arising from deposition of
­triglycerides between the leaflets of the bilayer of the ER mem­ 1. Rapoport, T.A., Li, L., and Park, E. Structural and mechanistic insights into
brane [75, 79]. LDs surrounded by cytosolic leaflet phospholip­ protein translocation. Annu Rev Cell Dev Biol, 2017;33:369–90.
2. Glick, B.S. and Nakano, A. Membrane traffic within the Golgi apparatus.
ids subsequently bud towards the cytoplasm, whereas VLDL
Annu Rev Cell Dev Biol, 2009;25:113–32.
particles surrounded by phospholipids of the luminal leaflet bud 3. Nixon‐Abell, J., Obara, C.J., Weigel, A.V. et  al. Increased spatiotemporal
into the ER lumen. VLDL particles then enter the secretory resolution reveals highly dynamic dense tubular matrices in the peripheral
pathway, from where they are released from hepatocytes into ER. Science, 2016;354(6311).
the bloodstream (Figure 7.3b). 4. Shibata, Y., Hu, J., Kozlov, M.M., and Rapoport, T.A. Mechanisms shaping the
membranes of cellular organelles. Annu Rev Cell Dev Biol, 2009;25:329–54.
In hepatocytes, overexpression of lipin1 has been shown to 5. Rambourg, A. and Clermont Y. Three‐dimensional electron microscopy:
repress VLDL secretion [98, 99]. The mechanism has not been structure of the Golgi apparatus. Eur J Cell Biol, 1990;51(2):189–200.
fully elucidated, but an attractive hypothesis is that high levels 6. Farquhar, M.G. and Palade, G.E. The Golgi apparatus (complex)‐(1954–1981)‐
of lipin1 channel PA to triglyceride synthesis rather than from artifact to center stage. J Cell Biol, 1981;91(3 Pt 2):77s–103s.
­phospholipid synthesis, the latter being required for secretory 7. Bonifacino, J.S. and Glick, B.S. The mechanisms of vesicle budding and
fusion. Cell, 2004;116(2):153–66.
pathway activity. This could have important implications for the 8. Dacks, J.B. and Robinson, M.S. Outerwear through the ages: evolutionary
pathogenesis of alcoholic fatty liver disease (FLD), as it has cell biology of vesicle coats. Curr Opin Cell Biol, 2017;47:108–16.
been shown that ethanol increases lipin1 levels [99]. Moreover, 9. Bykov, Y.S., Schaffer, M., Dodonova, S.O. et al. The structure of the COPI
ethanol exposure inhibits VLDL secretion in mouse liver, and coat determined within the cell. eLife, 2017;6.
10. Donaldson, J.G. and Jackson, C.L. ARF family G proteins and their regula­
the increase in lipin1 expression induced by ethanol is mediated
tors: roles in membrane transport, development and disease. Nat Rev Mol
by AMPK‐SREBP1 signaling [75, 99]. In response to changing Cell Biol, 2011;12(6):362–75.
nutritional signals, mTORC1 regulates lipin activity by control­ 11. Gillingham, A.K. and Munro S. The small G proteins of the Arf family and
ling its nuclear localization in hepatocytes [100]. their regulators. Annu Rev Cell Dev Biol, 2007;23:579–611.
12. Lee, M.C., Miller, E.A., Goldberg, J., Orci, L., and Schekman R. Bi‐­
directional protein transport between the ER and Golgi. Annu Rev Cell Dev
Biol, 2004;20:87–123.
FUTURE DIRECTIONS 13. Lee, M.C., Orci, L., Hamamoto, S. et  al. Sar1p N‐terminal helix initiates
membrane curvature and completes the fission of a COPII vesicle. Cell,
2005;122(4):605–17.
Major hepatocellular functions include secretion of nascent 14. Barlowe, C. and Helenius, A. Cargo capture and bulk flow in the early secre­
proteins and VLDL particles into the circulation, and the forma­ tory pathway. Annu Rev Cell Dev Biol, 2016;32:197–222.
tion of bile, processes for which protein and lipid trafficking are 15. Brown, H.A., Gutowski, S., Moomaw, C.R., Slaughter, C., and Sternweis,
essential. This chapter has described some of the molecular P.C. ADP‐ribosylation factor, a small GTP‐dependent regulatory protein,
stimulates phospholipase D activity. Cell, 1993;75(6):1137–44.
components required to direct nascent proteins and more 16. De Matteis, M.A., Wilson, C., and D’Angelo, G. Phosphatidylinositol‐
­complex lipoprotein particles (VLDL) from the ER to the Golgi 4‐phosphate: the Golgi and beyond. BioEssays, 2013;35(7):612–22.
apparatus and then to their final destination. The core machin­ 17. Ktistakis, N.T., Brown, H.A., Waters, M.G., Sternweis, P.C., and Roth, M.G.
ery is evolutionarily conserved and operates in all cells, but is Evidence that phospholipase D mediates ADP ribosylation factor‐dependent
modified for the specific trafficking processes performed by the formation of Golgi coated vesicles. J Cell Biol, 1996;134(2):295–306.
18. Ambroggio, E., Sorre, B., Bassereau, P. et al. ArfGAP1 generates an Arf1
liver, such as secretion of VLDL. Attaining a better understand­ gradient on continuous lipid membranes displaying flat and curved regions.
ing of how these processes are utilized by the liver under nor­ EMBO J, 2010;29(2):292–303.
mal and pathophysiological conditions is a current and future 19. Bannykh, S.I., Rowe, T., and Balch, W.E. The organization of endoplasmic
challenge for liver cell biologists. An important development reticulum export complexes. J Cell Biol, 1996;135(1):19–35.
20. Cosson, P. and Letourneur F. Coatomer interaction with di‐lysine endoplas­
over the past few years has been elucidation of the intimate con­
mic reticulum retention motifs. Science, 1994;263(5153):1629–31.
nection between membrane trafficking pathways and lipid 21. Letourneur, F., Gaynor, E.C., Hennecke, S. et al. Coatomer is essential for
transport and metabolism. Both production of cytosolic lipid retrieval of dilysine‐tagged proteins to the endoplasmic reticulum. Cell,
droplets and luminal VLDL particles is closely coupled to 1994;79(7):1199–207.
vesicular trafficking from the ER to the Golgi. An important 22. Bonifacino, J.S. Adaptor proteins involved in polarized sorting. J Cell Biol,
2014;204(1):7–17.
area for future work is in understanding this crosstalk as it
23. Guo, Y., Sirkis, D.W., and Schekman R. Protein sorting at the trans‐Golgi
occurs in liver cells. Elucidating the mechanisms involved will network. Annu Rev Cell Dev Biol, 2014;30:169–206.
provide avenues for therapeutic intervention not only in liver 24. Manolea, F., Chun, J., Chen, D.W. et  al. Arf3 is activated uniquely at the
diseases such as alcoholic and non‐alcoholic FLD, but also in trans‐Golgi network by brefeldin A‐inhibited guanine nucleotide exchange
countering infection by viruses (such as hepatitis C and E) factors. Mol Biol Cell, 2010;21(11):1836–49.
25. Devos, D., Dokudovskaya, S., Alber, F. et al. Components of coated vesicles
which subvert these processes for their propagation. and nuclear pore complexes share a common molecular architecture. PLoS
Biol, 2004;2(12):e380.
26. Duden, R., Griffiths, G., Frank, R., Argos, P., and Kreis, T.E. Beta‐COP,
a 110 kd protein associated with non‐clathrin‐coated vesicles and the
ACKNOWLEDGMENTS Golgi complex, shows homology to beta‐adaptin. Cell, 1991;64(3):
649–65.
This chapter is based on the chapter by Susan Chi and Mark 27. Serafini, T., Stenbeck, G., Brecht, A. et al. A coat subunit of Golgi‐derived
non‐clathrin‐coated vesicles with homology to the clathrin‐coated vesicle
McNiven in the previous edition. CLJ was supported by grants
coat protein beta‐adaptin. Nature, 1991;349(6306):215–20.
from the ANR (ANR-13-BSV2-0013) and the “Fondation pour 28. Beck, R., Rawet, M., Wieland, F.T., and Cassel D. The COPI system:
la Recherche Médicale” (DEQ20150934717), France. ­molecular mechanisms and function. FEBS Lett, 2009;583(17):2701–9.
84 THE LIVER:  REFERENCES

29. Orci, L., Stamnes, M., Ravazzola, M. et al. Bidirectional transport by distinct 53. Zolov, S.N. and Lupashin, V.V. Cog3p depletion blocks vesicle‐mediated
populations of COPI‐coated vesicles. Cell, 1997;90(2):335–49. Golgi retrograde trafficking in HeLa cells. J Cell Biol, 2005;168(5):
30. Bui, Q.T., Golinelli‐Cohen, M.P., and Jackson, C.L. Large Arf1 guanine nucle­ 747–59.
otide exchange factors: evolution, domain structure, and roles in membrane 54. Guo, Y., Punj, V., Sengupta, D., and Linstedt, A.D. Coat‐tether interaction in
trafficking and human disease. Mol Genet Genomics, 2009;282(4):329–50. Golgi organization. Mol Biol Cell, 2008;19(7):2830–43.
31. Casanova, J.E. Regulation of Arf activation: the Sec7 family of guanine 55. Goueslain, L., Alsaleh, K., Horellou, P. et  al. Identification of GBF1 as a
nucleotide exchange factors. Traffic, 2007;8(11):1476–85. cellular factor required for hepatitis C virus RNA replication. J Virol,
32. Novick, P., Field, C., and Schekman R. Identification of 23 complementation 2010;84(2):773–87.
groups required for post‐translational events in the yeast secretory pathway. 56. Farhat, R., Ankavay, M., Lebsir, N. et al. Identification of GBF1 as a cellular
Cell, 1980;21(1):205–15. factor required for hepatitis E virus RNA replication. Cell Microbiol,
33. Peyroche, A., Paris, S., and Jackson, C.L. Nucleotide exchange on ARF 2018;20(1).
mediated by yeast Gea1 protein. Nature, 1996;384(6608):479–81. 57. Lavie, M. and Dubuisson J. Interplay between hepatitis C virus and lipid
34. Claude, A., Zhao, B.P., Kuziemsky, C.E. et al. GBF1: a novel Golgi‐associ­ metabolism during virus entry and assembly. Biochimie, 2017;141:62–9.
ated BFA‐resistant guanine nucleotide exchange factor that displays speci­ 58. Farhat, R., Seron, K., Ferlin, J. et al. Identification of class II ADP‐ribosylation
ficity for ADP‐ribosylation factor 5. J Cell Biol, 1999;146(1):71–84. factors as cellular factors required for hepatitis C virus replication. Cell
35. Schlacht, A., Mowbrey, K., Elias, M., Kahn, R.A., and Dacks, J.B. Ancient Microbiol, 2016;18(8):1121–33.
complexity, opisthokont plasticity, and discovery of the 11th subfamily of 59. Altan‐Bonnet, N., Sougrat, R., and Lippincott‐Schwartz, J. Molecular basis
Arf GAP proteins. Traffic, 2013;14(6):636–49. for Golgi maintenance and biogenesis. Curr Opin Cell Biol,
36. Schlacht A. Evolution of the vesicle formation machinery. PhD thesis, 2004;16(4):364–72.
Department of Cell Biology, University of Alberta, Edmonton, Canada, 60. Jamieson, J.D. and Palade, G.E. Synthesis, intracellular transport, and
director Joel Dacks, 2015. ­discharge of secretory proteins in stimulated pancreatic exocrine cells. J Cell
37. Lowery, J., Szul, T., Styers, M. et al. The Sec7 guanine nucleotide exchange Biol, 1971;50(1):135–58.
factor GBF1 regulates membrane recruitment of BIG1 and BIG2 guanine 61. Presley, J.F., Cole, N.B., Schroer, T.A. et al. ER‐to‐Golgi transport ­visualized
nucleotide exchange factors to the trans‐Golgi network (TGN). J Biol Chem, in living cells. Nature, 1997;389(6646):81–5.
2013;288(16):11532–45. 62. Barlowe, C., Orci, L., Yeung, T. et al. COPII: a membrane coat formed by
38. Bonifacino, J.S. and Lippincott‐Schwartz, J. Coat proteins: shaping mem­ Sec proteins that drive vesicle budding from the endoplasmic reticulum.
brane transport. Nat Rev Mol Cell Biol, 2003;4(5):409–14. Cell, 1994;77(6):895–907.
39. Deng, Y., Golinelli‐Cohen, M.P., Smirnova, E., and Jackson, C.L. A COPI 63. Mizuno, M. and Singer, S.J. A soluble secretory protein is first concentrated
coat subunit interacts directly with an early‐Golgi localized Arf exchange in the endoplasmic reticulum before transfer to the Golgi apparatus. Proc
factor. EMBO Rep, 2009;10(1):58–64. Natl Acad Sci U S A, 1993;90(12):5732–6.
40. Sato, K. and Nakano A. Dissection of COPII subunit‐cargo assembly and 64. Orci, L., Ravazzola, M., Meda, P. et al. Mammalian Sec23p homologue is
disassembly kinetics during Sar1p‐GTP hydrolysis. Nat Struct Mol Biol, restricted to the endoplasmic reticulum transitional cytoplasm. Proc Natl
2005;12(2):167–74. Acad Sci U S A, 1991;88(19):8611–15.
41. Tanigawa, G., Orci, L., Amherdt, M. et al. Hydrolysis of bound GTP by ARF 65. Bielli, A., Haney, C.J., Gabreski, G. et al. Regulation of Sar1 NH2 terminus
protein triggers uncoating of Golgi‐derived COP‐coated vesicles. J Cell Biol, by GTP binding and hydrolysis promotes membrane deformation to control
1993;123(6 Pt 1):1365–71. COPII vesicle fission. J Cell Biol, 2005;171(6):919–24.
42. Forster, R., Weiss, M., Zimmermann, T. et  al. Secretory cargo regulates the 66. Aguilera‐Gomez, A., Zacharogianni, M., van Oorschot, M.M. et al. Phospho‐
turnover of COPII subunits at single ER exit sites. Curr Biol, rasputin stabilization by Sec16 is required for stress granule formation upon
2006;16(2):173–9. amino acid starvation. Cell Rep, 2017;20(4):935–48.
43. Kliouchnikov, L., Bigay, J., Mesmin, B. et  al. Discrete determinants in 67. Hanna, M.G., Block, S., Frankel, E.B. et  al. TFG facilitates outer coat
ArfGAP2/3 conferring Golgi localization and regulation by the COPI coat. ­disassembly on COPII transport carriers to promote tethering and fusion
Mol Biol Cell, 2009;20(3):859–69. with ER‐Golgi intermediate compartments. Proc Natl Acad Sci U S A,
44. Weimer, C., Beck, R., Eckert, P. et  al. Differential roles of ArfGAP1, 2017;114(37):E7707–E16.
ArfGAP2, and ArfGAP3 in COPI trafficking. J Cell Biol, 2008;183(4): 68. Saito, K., Chen, M., Bard, F. et  al. TANGO1 facilitates cargo loading at
725–35. endoplasmic reticulum exit sites. Cell, 2009;136(5):891–902.
45. Bigay, J., Casella, J.F., Drin, G., Mesmin, B., and Antonny, B. ArfGAP1 69. Cai, H., Yu, S., Menon, S. et al. TRAPPI tethers COPII vesicles by binding
responds to membrane curvature through the folding of a lipid packing sen­ the coat subunit Sec23. Nature, 2007;445(7130):941–4.
sor motif. EMBO J, 2005;24(13):2244–53. 70. Santos, A.J., Nogueira, C., Ortega‐Bellido, M., and Malhotra V. TANGO1
46. Bigay, J., Gounon, P., Robineau, S., and Antonny, B. Lipid packing sensed and Mia2/cTAGE5 (TALI) cooperate to export bulky pre‐chylomicrons/
by ArfGAP1 couples COPI coat disassembly to membrane bilayer curvature. VLDLs from the endoplasmic reticulum. J Cell Biol, 2016;213(3):343–54.
Nature, 2003;426(6966):563–6. 71. Shimoi, W., Ezawa, I., Nakamoto, K. et al. p125 is localized in endoplasmic
47. Barrowman, J., Bhandari, D., Reinisch, K., and Ferro‐Novick S. TRAPP reticulum exit sites and involved in their organization. J Biol Chem,
complexes in membrane traffic: convergence through a common Rab. Nat 2005;280(11):10141–8.
Rev Mol Cell Biol, 2010;11(11):759–63. 72. Melero, A., Chiaruttini, N., Karashima, T. et al. Lysophospholipids facilitate
48. Schroeter, S., Beckmann, S., and Schmitt, H.D. Coat/tether interactions‐ COPII vesicle formation. Curr Biol, 2018;28(12):1950–8 e6.
exception or rule? Front Cell Dev Biol, 2016;4:44. 73. Brodsky, J.L., Gusarova, V., and Fisher, E.A. Vesicular trafficking of hepatic
49. Andag, U. and Schmitt, H.D. Ds11p, an essential component of the Golgi‐ apolipoprotein B100 and its maturation to very low‐density lipoprotein par­
endoplasmic reticulum retrieval system in yeast, uses the same sequence ticles; studies from cells and cell‐free systems. Trends Cardiovasc Med,
motif to interact with different subunits of the COPI vesicle coat. J Biol 2004;14(4):127–32.
Chem, 2003;278(51):51722–34. 74. Zanetti, G., Pahuja, K.B., Studer, S., Shim, S., and Schekman R. COPII and the
50. Reilly, B.A., Kraynack, B.A., VanRheenen, S.M., and Waters, M.G. Golgi‐ regulation of protein sorting in mammals. Nat Cell Biol, 2011;14(1):20–8.
to‐endoplasmic reticulum (ER) retrograde traffic in yeast requires Ds11p, a 75. Gluchowski, N.L., Becuwe, M., Walther, T.C., and Farese, R.V., Jr. Lipid
component of the ER target site that interacts with a COPI coat subunit. Mol droplets and liver disease: from basic biology to clinical implications. Nat
Biol Cell, 2001;12(12):3783–96. Rev Gastroenterol Hepatol, 2017;14(6):343–55.
51. Ren, Y., Yip, C.K., Tripathi, A. et  al. A structure‐based mechanism for 76. Gusarova, V., Brodsky, J.L., and Fisher, E.A. Apolipoprotein B100 exit
vesicle capture by the multisubunit tethering complex Ds11. Cell,
­ from the endoplasmic reticulum (ER) is COPII‐dependent, and its lipidation
2009;139(6):1119–29. to  very low density lipoprotein occurs post‐ER. J Biol Chem,
52. Suvorova, E.S., Duden, R., and Lupashin, V.V. The Sec34/Sec35p complex, 2003;278(48):48051–8.
a Ypt1p effector required for retrograde intra‐Golgi trafficking, interacts 77. Shoulders, C.C., Stephens, D.J., and Jones, B. The intracellular transport of
with Golgi SNAREs and COPI vesicle coat proteins. J Cell Biol, chylomicrons requires the small GTPase, Sar1b. Curr Opin Lipidol,
2002;157(4):631–43. 2004;15(2):191–7.
7:  The Hepatocellular Secretory Pathway 85

78. Santos, A.J., Raote, I., Scarpa, M., Brouwers, N., and Malhotra V. TANGO1 89. Henley, J.R. and McNiven, M.A. Association of a dynamin‐like protein with
recruits ERGIC membranes to the endoplasmic reticulum for procollagen the Golgi apparatus in mammalian cells. J Cell Biol, 1996;133(4):761–75.
export. eLife, 2015;4. 90. Cao, H., Weller, S., Orth, J.D. et al. Actin and Arf1‐dependent recruitment
79. Tiwari, S. and Siddiqi, S.A. Intracellular trafficking and secretion of VLDL. of a cortactin‐dynamin complex to the Golgi regulates post‐Golgi transport.
Arterioscler Thromb Vasc Biol, 2012;32(5):1079–86. Nat Cell Biol, 2005;7(5):483–92.
80. Novikoff, P.M. and Yam A. Sites of lipoprotein particles in normal rat hepat­ 91. Renard, H.F., Johannes, L., and Morsomme, P. Increasing diversity of biologi­
ocytes. J Cell Biol, 1978;76(1):1–11. cal membrane fission mechanisms. Trends Cell Biol, 2018;28(4):274–86.
81. Peden, A.A., Rudge, R.E., Lui, W.W., and Robinson, M.S. Assembly and 92. Miserey‐Lenkei, S., Chalancon, G., Bardin, S. et al. Rab and actomyosin‐
function of AP‐3 complexes in cells expressing mutant subunits. J Cell Biol, dependent fission of transport vesicles at the Golgi complex. Nat Cell Biol,
2002;156(2):327–36. 2010;12(7):645–54.
82. Guo, X., Mattera, R., Ren, X. et  al. The adaptor protein‐1 mu1B subunit 93. Jackson, C.L., Walch, L., and Verbavatz, J.M. Lipids and their trafficking:
expands the repertoire of basolateral sorting signal recognition in epithelial an integral part of cellular organization. Dev Cell, 2016;39(2):139–53.
cells. Dev Cell. 2013;27(3):353–66. 94. Vance, J.E. Phospholipid synthesis and transport in mammalian cells.
83. Doray, B., Bruns, K., Ghosh, P., and Kornfeld, S. Interaction of the cation‐ Traffic, 2015;16(1):1–18.
dependent mannose 6‐phosphate receptor with GGA proteins. J Biol Chem, 95. Walther, T.C., Chung, J., and Farese, R.V., Jr. Lipid droplet biogenesis.
2002;277(21):18477–82. Annu Rev Cell Dev Biol, 2017;33:491–510.
84. Puertollano, R., Randazzo, P.A., Presley, J.F., Hartnell, L.M., and Bonifacino, 96. Gaspar, M.L., Hofbauer, H.F., Kohlwein, S.D., and Henry, S.A.
J.S. The GGAs promote ARF‐dependent recruitment of clathrin to the TGN. Coordination of storage lipid synthesis and membrane biogenesis: evidence
Cell, 2001;105(1):93–102. for cross‐talk between triacylglycerol metabolism and phosphatidylinositol
85. Collins, B.M., Praefcke, G.J., Robinson, M.S., and Owen, D.J. Structural synthesis. J Biol Chem, 2011;286(3):1696–708.
basis for binding of accessory proteins by the appendage domain of GGAs. 97. Siniossoglou, S. Phospholipid metabolism and nuclear function: roles of
Nat Struct Biol, 2003;10(8):607–13. the lipin family of phosphatidic acid phosphatases. Biochim Biophys Acta,
86. Puertollano, R., Aguilar, R.C., Gorshkova, I., Crouch, R.J., and Bonifacino, 2013;1831(3):575–81.
J.S. Sorting of mannose 6‐phosphate receptors mediated by the GGAs. 98. Wang, Y., Viscarra, J., Kim, S.J., and Sul, H.S. Transcriptional regulation of
Science, 2001;292(5522):1712–16. hepatic lipogenesis. Nat Rev Mol Cell Biol, 2015;16(11):678–89.
87. Zhen, Y. and Stenmark H. Cellular functions of Rab GTPases at a glance. 99. You, M., Jogasuria, A., Lee, K. et al. Signal transduction mechanisms of
J Cell Sci, 2015;128(17):3171–6. alcoholic fatty liver disease: emerging role of lipin‐1. Curr Mol Pharmacol,
88. Wang, J., Morita, Y., Mazelova, J., and Deretic, D. The Arf GAP ASAP1 2017;10(3):226–36.
provides a platform to regulate Arf4‐ and Rab11‐Rab8‐mediated ciliary 100. Peterson, T.R., Sengupta, S.S., Harris, T.E. et al. mTOR complex 1 ­regulates
receptor targeting. EMBO J, 2012;31(20):4057–71. lipin 1 localization to control the SREBP pathway. Cell, 2011;146(3):408–20.
Mitochondrial Function,
8 Dynamics, and Quality Control
Marc Liesa, Ilan Benador, Nathanael Miller, and Orian S. Shirihai
Department of Medicine, Division of Endocrinology and Department of Molecular and Medical Pharmacology,
David Geffen School of Medicine at UCLA, Los Angeles, CA, USA

INTRODUCTION MITOCHONDRIAL OXIDATIVE


PHOSPHORYLATION (OXPHOS)
The word “mitochondrion” originates from the fusion of two
Greek words: mitos, meaning thread and chondros, meaning Mitochondria are organelles formed by two membrane layers
grain [1]. This term was coined by Benda after visual inspection with different compositions of lipids and proteins, separated
of these organelles. It is worth noting that thread and grain are by an intermembrane space (IMS). Over 1500 proteins are esti-
the most common shapes of bacteria: coccus was coined for mated to reside in human mitochondria, but only 13 are encoded
grain‐shaped bacteria and bacillus for thread‐shaped bacteria. by the mitochondrial DNA (mtDNA). This low number of
Later, it was demonstrated that mitochondria originated from ­proteins means that protein import through the mitochondrial
bacteria engulfed by the ancestors of eukaryotic cells. The intra- membrane layers is a central process for mitochondrial biogen-
cellular parasite Rickettsia is widely accepted as the α‐proteo- esis, function, and turnover.
bacteria order from which mitochondria originated and is The outer mitochondrial membrane is bidirectionally perme-
morphologically classified as a cocco‐bacillus bacterial species. able to small solutes, including small peptides (<5 kDa), thanks
Therefore, the endosymbiotic theory has both anatomical and to the presence of a protein that acts as a channel, named porin/
DNA conservation evidence in its favor. In this regard, the main VDAC. The family of TOM (translocase of the outer membrane)
features of mitochondria that are present among all eukaryotic proteins is responsible for protein transport across the outer
species are conserved in bacteria, namely: (i) Mitochondria con- mitochondrial membrane [2, 3].
sume oxygen to synthesize ATP, using a process named oxida- The permeability of the inner mitochondrial membrane is
tive phosphorylation (OXPHOS), similar to bacterial aerobic more stringent, which limits the bidirectional diffusion of very
ATP synthesis. (ii) Mitochondria have their own genome, which small molecules, including protons (H+, 1 Da). This stringency
encodes for 13 subunits of the respiratory complexes executing means that the transport across the inner membrane of nutrients,
OXPHOS, tRNA, and rRNAs required for their translation. (iii) fuels, metabolites, and cofactors of polar nature is strictly
Mitochondria are motile organelles that can fuse or divide into dependent on transporter proteins and their regulatory mecha-
smaller organelles, similar to bacteria. nisms. The inner membrane is the location of proteins that exe-
In this chapter, we will provide a summary of these three cute oxidative phosphorylation (OXPHOS), a process briefly
conserved features of mitochondria, with a focus on hepato- defined as a chain of redox events involving electron transport
cyte mitochondrial OXPHOS function and the role of across protein multimers (complexes) that extrude protons to
­mitochondrial dynamics in hepatocytes determining OXPHOS the IMS, with proton re‐entry through ATP synthase producing
function. ATP. The stringent permeability to protons, together with high

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
8:  Mitochondrial Function, Dynamics, and Quality Control 87

proton concentrations within the IMS limiting nutrient oxida- Oxidation products from different nutrients converge
tion rates, allows ATP synthase and thus ATP demand to control in the TCA cycle
OXPHOS rates (i.e. how much nutrient is oxidized and how
Glucose‐derived pyruvate, amino acid, and fatty acid oxidation in
much oxygen is consumed). Moreover, protein import across
the mitochondria result in intermediate and final metabolites that
the inner membrane to build mitochondria is executed by the
can enter the citric acid/Krebs/TCA cycle. This redox cycle con-
family of TIM proteins (translocase of the inner membrane),
sists of eight reactions catalyzed by: (i) four dehydrogenases (two
with energy‐demanding translocation of proteins also being
of which are decarboxylases), (ii) two synthetases, (iii) one
fueled by the difference in charge and pH across the inner mem-
isomerase, and (iv) one hydratase that interconnects eight carbox-
brane generated by the proton gradient [4]. The dependency of
ylic acids, with four to six carbons in their structure. The cycle
these two crucial processes on the proton gradient illustrate that
starts with the synthesis of citrate (six carbons) from the condensa-
the stringent permeability of the inner membrane is essential for
tion of oxaloacetate (four carbons) and acetyl‐CoA (two carbons)
proper mitochondrial function.
catalyzed by citrate synthase. Importantly, two carbons are lost as
In this context, mitochondrial OXPHOS is constituted by
two CO2 in one turn of the cycle, while generating three NADH,
three major processes acting in coordination and ultimately
one FADH2, one ATP, or GTP. This means that a minimal entry of
determined by ATP synthase activity: (i) nutrient and fuel oxida-
two carbons is required in each turn to replace the two carbons lost
tion; (ii) electron transport chain and cytochrome c oxidase
as CO2, as no carboxylases are part of the core enzymes constitut-
activity; and (iii) mitochondrial ATP synthase activity.
ing the TCA cycle itself. As these eight TCA cycle reactions occur
within the mitochondrial matrix and the eight TCA acids are nega-
Nutrient and fuel oxidation tively charged, their concentrations in the matrix need to be main-
Ninety percent of the oxygen that we breathe and transform to tained within a range. This maintenance is not only required to
water and CO2 is consumed by mitochondria. This can explain sustain the TCA cycle “spinning” to provide transferable electrons
why some nutrients are fully oxidized inside the mitochondria and ATP, but also to preserve the energy contained in the differ-
(some amino acids and fatty acids). Moreover, nutrients whose ence in charge and pH generated by the proton gradient across the
oxidation does not occur inside mitochondria can provide fuels/ inner membrane. The need for carbon supply in a context of a
intermediates to be oxidized inside mitochondria (i.e. glucose‐ limited capacity to change TCA cycle metabolite concentrations
derived glycerol‐3‐phosphate and pyruvate). Nutrient oxidation in the matrix is widely accepted to explain the requirement for
reactions in the mitochondria mostly occur in the matrix, with a anaplerotic and cataplerotic processes.
few in the IMS (glycerol‐3‐phosphate dehydrogenase). These
Anaplerosis and cataplerosis are required to sustain TCA
reactions are catalyzed by enzymes named dehydrogenases that
strip electrons from nutrients and fuels and transfer them to cycle rates and mitochondrial function
nicotine and flavin adenine dinucleotides (NAD+ and FAD+). The carboxylic acids constituting the TCA cycle can be consumed
These adenine dinucleotides can accept two electrons each and for de novo synthesis of multiple molecules, including glucose,
carry them for their transfer to the electron transport chain. In fatty acids, and heme, among others. The processes that replenish
the case of hepatocytes, mitochondrial nutrient and fuel oxida- TCA cycle intermediates after their consumption are named “ana-
tion is highly regulated and complex. This complexity and high plerotic.” In contrast, “cataplerosis” refers to the processes that
regulation means that nutrient oxidation in the mitochondria is consume TCA cycle intermediates, which can be linked to biosyn-
not exclusively determined by nutrient availability outside or thetic processes. Hepatocytes have a strong demand for anaplero-
inside hepatocytes, or even in the capacity of OXPHOS com- sis to sustain cataplerotic processes that are regulated by fed and
plexes to transfer electrons coupled to ATP synthesis. There are fasting states, including synthesis of fatty acids, glucose, and
four underlying reasons for this complexity as describing in the heme. Thus, anaplerosis and cataplerosis are tightly and hormo-
following. nally regulated in hepatocytes. The perfect example for an anaple-
rotic process in hepatocytes is the reaction catalyzed by
Mitochondrial nutrient oxidation can sustain nutrient mitochondrial pyruvate carboxylase. This enzyme uses ATP
synthesis in hepatocytes hydrolysis to add one carbon from bicarbonate (which can be
Hepatocytes are responsible for synthesis and supply nutrients originated from CO2) to pyruvate. Under the fed state, pyruvate
to other tissues (preferentially the brain) under organismal fast- derived from glucose is used by both pyruvate carboxylase and
ing and starvation, mostly glucose and ketone bodies. Under pyruvate dehydrogenase to generate oxaloacetate and acetyl‐CoA,
fasting, hepatocytes do not oxidize glucose, they oxidize fatty respectively. In this way, pyruvate carboxylase activity replenishes
acids mostly in the mitochondria to sustain their ATP demand the oxaloacetate that left the TCA cycle, as citrate is used for fatty
and produce ketone bodies [5]. In the case of amino acids, their acid synthesis or that left to be a carbon source for glucose synthe-
oxidation during fasting can be used to provide ATP and carbon sis (cataplerosis). Another major cataplerotic process in liver is
intermediates for gluconeogenesis. Under the fed state, nutrient heme synthesis, as the first step in its synthesis involves the con-
oxidation preferences change, with mitochondria oxidizing sumption of succinyl‐CoA (one of the eight TCA intermediates).
pyruvate resulting from glucose oxidation, a process that sup-
Glucose‐derived pyruvate and fatty acids compete
ports fatty acid synthesis. Thus, mitochondrial nutrient oxida-
to be oxidized by mitochondrial OXPHOS
tion and fuel/nutrient transport is specifically tuned to support
these context‐dependent hepatocyte functions, being sensitive This competition facilitates the transition and therefore the
to hormonal regulation (which respond to fasting and feeding). execution of the opposite roles in fat metabolism of hepatocyte
88 THE LIVER:  MITOCHONDRIAL OXIDATIVE PHOSPHORYLATION (OXPHOS)

mitochondria in the fed versus the fasting state. In the post- • Ubiquinone reduction: Two electrons are used to reduce
prandial state, glucose‐derived pyruvate is oxidized in the ubiquinone to ubiquinol, which is inserted in the inner
mitochondrial matrix to generate citrate to be used for de novo ­mitochondrial membrane. Ubiquinol transfers the electrons
fatty acid synthesis, whereas during fasting, mitochondrial fatty to complex III and can freely diffuse within the inner mito-
acid oxidation fuels glucose synthesis and ketone body produc- chondrial membrane. This diffusion allows the transfer from
tion. Key molecular mechanisms allowing this transition based the accepting electron complexes to complex III.
on competition can be summarized in two points: • Complex III‐mediated reduction of cytochrome c: Complex III
is also named ubiquinol:cytochrome c oxidoreductase. Its role
• Product inhibition of TCA dehydrogenases and pyruvate
is to transfer the electrons from ubiquinol to cytochrome c.
dehydrogenase (PDH): High levels of ATP, acetyl‐CoA, and
Cytochrome c is a small protein (104 amino acids) containing
NADH during fatty acid oxidation in the matrix can slow
a heme group and associated with the inner membrane. The
down PDH and TCA cycle oxidation rates by product inhibi-
electrons are accepted by the Fe3+ within the heme group to be
tion, decreasing CO2 release [6, 7]. Slowing down the TCA
reduced to Fe2+. Electrons from Fe2+‐cytochrome c molecules
cycle allows balancing anaplerosis resulting from fatty acid
are stripped by complex IV, to regenerate Fe3+‐cytochrome c
and amino acid oxidation, with cataplerosis consuming
and transfer the electrons to O2.
oxaloacetate and acetyl‐CoA for glucose and ketone body
• Proton translocation: Complexes I, III, and IV are the multi-
synthesis, respectively. This competitive mechanism in liver
protein complexes. Complex I comprises 44–45 proteins,
and in other tissues such as muscle contributes to the decrease
complex III has 11 proteins, and complex IV has 13–14
in organismal CO2 production rates under fasting or when
­proteins (one protein initially thought to be part of complex
fatty acid oxidation is the major nutrient being oxidized.
I  is now thought to be part of complex IV). The presence
• Inhibition of long‐chain fatty acid entry into the mitochondria:
of multiple proteins allows coupling of the electron transport
When mitochondrial oxidation of glucose‐derived pyruvate
process and redox events to proton pumping to the IMS.
increases in the fed state, cataplerosis mandates that elevated
Accordingly, complexes transferring electrons and with lower
citrate is exported out of the mitochondria. Exported citrate is
numbers of proteins do not pump protons. Complex II con-
the precursor in the cytosol of malonyl‐CoA, a molecule that
tains 4 proteins, ETF/ETF:QO has 2 proteins, and glycerol‐3‐
inhibits activated fatty acid translocation into the mitochondria
phosphate dehydrogenase is a single protein. Although these
by blocking the protein executing this transport, namely CPT1
complexes do not pump protons themselves, the electrons
(located in the outer membrane) [8]. The fact that glucose‐medi-
delivered to ubiquinone will still induce proton pumping when
ated inhibition of fatty acid oxidation occurs by inhibiting the
they reach complexes III and IV. As these complexes can pro-
entry of fatty acids supports the conclusion that, by default, fatty
vide electrons much faster than complex I, it makes sense that
acid oxidation will win the competition with pyruvate oxidation
they do not pump protons to prevent a sudden increase in
in the mitochondrial matrix. It is also likely that it allows an
membrane potential. Such an increase could even stop the
active and faster import of long‐chain fatty acids into the mito-
electron transport process itself, which increases the risk of
chondria to meet ATP demand and ketone body production
these electrons generating reactive oxygen species instead of
under fasting, which might not be possible by diffusion.
reaching complex IV to reduce oxygen and ­generate water.
• Supercomplex formation: The complexes that pump protons (I,
Electron transport chain and cytochrome c III, and IV) can be found in isolation but also form units named
“supercomplexes.” The exact mechanisms behind supercom-
oxidase activity plex assembly and their physiological regulation by feeding and
The electron transport chain is constituted by four multiprotein fasting in mitochondria have only been characterized recently,
complexes, named complexes I, II, III, and IV. Electrons carried and they are still subject to intense research [9]. Evidence sug-
by NADH enter through complex I, a multiprotein complex also gests that supercomplex regulation and composition might dif-
known as NADH:ubiquinone oxidoreductase, whereas electrons fer in mitochondria from different tissues (i.e. heart is different
carried by FADH2 can enter via complex II, also known as the from liver). However, there is agreement that supercomplex
TCA cycle enzyme succinate dehydrogenase. There are two formation can increase the efficiency of electron transfer by
additional sites for electron entry from FADH2: inner mem- cytochrome c from complex III to complex IV. Because
brane‐bound glycerol‐3‐phosphate dehydrogenase and the elec- cytochrome c is a small protein loosely bound to the inner
tron‐transferring flavoprotein (ETF) and ETF:ubiquinone membrane, the formation of a unit containing complexes III and
oxidoreductase (ETF:QO). Of note, glycerol‐3‐phosphate is an IV allows a pool of cytochrome c to be trapped in between
intermediate generated in the cytosol by glycolysis, while ETF them. This trapped pool would allow electron transfer to be
accepts the electrons from FADH2 generated by 11 different faster, because it would not be solely dependent on random dif-
mitochondrial dehydrogenases, mostly involved in mitochon- fusion of cytochrome c from complex III to complex IV. A simi-
drial fatty acid and amino acid oxidation processes. The destina- lar explanation has been given to justify the interaction between
tion of the electrons stripped from NADH and FADH2 is complexes I and III, but in this case it facilitates electron trans-
complex IV, also named cytochrome c oxidase. Complex IV fer mediated by membrane‐inserted but diffusible ubiquinol.
uses these electrons to reduce O2 to water and, therefore, it is
directly responsible for respiration.
The process of electron transfer from the entry sites to com-
Mitochondrial ATP synthase
plex IV involves four steps, which are highly conserved in mito- Mitochondrial ATP synthase is a multiprotein complex formed
chondria from all different tissues and are shared: by 13 proteins divided into two structurally defined regions: a
8:  Mitochondrial Function, Dynamics, and Quality Control 89

hydrophilic, matrix‐located region called F1 and a hydrophobic transport chain in maintaining the difference in charge generated
region called F0 inserted in the inner membrane. ATP synthase is by the proton gradient. This reverse activity is thought to allow
also known as complex V, although it cannot transport electrons protein import to replace dysfunctional components of the elec-
itself. The rationale behind being named complex V is the func- tron transport chain. Moreover, it still allows substrate and fuel
tional connection of ATP synthesis with the electron transport import for cataplerosis.
chain, usually referred as “coupling.” This coupling is achieved
by the need for ATP synthase to translocate protons to harness
the energy required to phosphorylate ADP to ATP. This energy
transfer starts in the F0 region, which forms the proton pore MITOCHONDRIAL DYNAMICS
across the inner membrane and constitutes a motor that rotates AND TURNOVER, AND THEIR
when protons translocate. The F0 motor directly interacts with
F1 region proteins, which as a result also rotate. The rotation of
RELATIONSHIP TO OXPHOS
F1 proteins causes conformational changes in other F1 proteins
In 1915, Lewis and Lewis showed that mitochondria inside cells
to catalyze the phosphorylation of ADP to ATP in the matrix.
can move from one location to another, and even visualized
Mitochondrial ATP synthase activity and, therefore, cellular
events supporting the existence of mitochondrial fusion and fis-
ATP demand can determine and control nutrient oxidation and
sion/division [11].
respiration rates. The electron transport chain and ATP synthase
evolved in such a way that a large difference in the proton gradient
slows down electron transfer and high ATP levels in the mitochon- Regulation of mitochondrial dynamics
drial matrix inhibit complex V‐mediated proton translocation
and quality control
back to the matrix. Therefore, if the ATP/ADP ratio is high in the
matrix, protons cannot translocate, and electron transport will Mitochondrial fusion is regulated by fusion proteins: mito-
stop, which will also prevent the complexes stripping electrons fusins 1 and 2 (Mfn1 and Mfn2) fuse the outer mitochondrial
from NADH and FADH2. The accumulation of NADH and membrane, and optic atrophy 1 (OPA1) fuses the inner mito-
FADH2 will then decrease nutrient oxidation rates. chondrial membrane (Figure  8.1). Conversely, mitochondrial
Importantly though, the mitochondrial proton gradient does fission factor (Mff) and dynamin related protein 1 (Drp1) are
not exclusively serve ATP synthesis activity. In agreement with responsible for mitochondrial fission. These proteins tightly
this, nutrient and fuel oxidation in the mitochondria do not regulate mitochondrial dynamics and morphology via hydroly-
exclusively serve either proton gradient maintenance or ATP sis of GTP to power their activities [12–14]. Mitochondrial
synthesis. Indeed, the intermediates or end products of mito- dynamics control mitochondrial architecture. Whereas fusion
chondrial nutrient/fuel oxidation are used to synthesize other and fission are ­controlled at the level of each mitochondrion in
nutrients or essential cofactors such as heme, which may be the cell, an overriding signal may result in cellular reduction in
needed under conditions of high ATP/ADP ratio. Therefore, fusion or fission activity. Increased mitochondrial fission and
other mechanisms are recruited to serve these additional pur- decreased fusion lead to fragmentation of the mitochondrial
poses of the proton gradient and mitochondrial nutrient oxida- network. An example of an overriding signal – the effect of cell
tion that are independent of changes in ATP demand. One of cycle on fusion and fission activity – is illustrated in Figure 8.2.
these mechanisms is proton leak: some protons can still cross Mfn2 expression is decreased in different tissues by diet‐
the inner membrane when ATP synthase activity (i.e. ATP induced obesity and contributes to tissue dysfunction [17–20].
demand) is low, but ATP synthase is still the major mechanism Mfn2 has been implicated in the regulation of mitochondrial
of proton re‐entry to the mitochondria. This is indeed the reason function in addition to its role in mitochondrial fusion. Knockdown
why it is called proton leak, as high ATP levels can still reduce of Mfn2 results in altered mitochondrial bioenergetic function, as
respiration rates, and the low respiration and nutrient oxidation reduced glucose oxidation and decreased oxygen consumption in
rates are possible thanks to the leak of protons (i.e. a small num- primary rodent muscle and muscle‐derived cell lines is evident
ber). Uncoupling happens when respiration and nutrient oxida- [17]. Removal of Mfn2 altered substrate preference and impaired
tion is increased because of another mechanism allowing proton thermogenesis in brown adipose tissue [21, 22]. These results col-
re‐entry at a higher capacity than the ATP synthase itself. This lectively demonstrate that mitochondrial morphology and tissue
uncoupling is not common in physiology; it is used to generate function are intimately connected.
heat in specialized tissues, such as brown adipose tissue. Mitochondrial morphology and dynamics control mitochon-
drial degradation via mitochondria‐specific autophagy (or
mitophagy) [23, 24], in part by constantly mixing and reorgan-
Reverse ATP synthase activity
izing the mitochondrial content distribution, but also by gener-
The reverse activity of mitochondrial ATP synthase reveals the ating small fission products that can be eliminated by mitophagy.
importance of the proton gradient beyond energizing ATP syn- Fission events generate daughter mitochondria that can depo-
thesis. Reverse activity is the “default activity” of ATP synthase, larize, leading to PINK1‐mediated phosphorylation of Mfn2
which is to hydrolyze ATP and extrude protons when the differ- [25–28]. PINK1, which is normally degraded in functional,
ence in membrane potential (proton gradient) is not large enough polarized mitochondria, is stabilized on depolarized mitochon-
[10]. This means that if the electron transport chain is damaged dria [29] due to inhibition of its protease. Stabilized PINK1
and cannot sustain the proton gradient, the ATP synthase can phosphorylates Mfn2, which is targeted by Parkin. Parkin ubiq-
extrude protons to the IMS, energized by ATP hydrolysis. In this uitinates Mfn2, whose ubiquitination designates mitochondria
way ATP synthase substitutes for the role of the electron destined for removal by mitophagy.
90 THE LIVER:  MITOCHONDRIAL DYNAMICS AND TURNOVER, AND THEIR RELATIONSHIP TO OXPHOS

Mitophagy

opa1↓
Fission mfn1↓
PINK1↑
parkin↑
2

Pre-autophagic pool
Fusion 3 4
Sustained
depolarization

1
Transient
depolarization

Figure 8.1  The mitochondria life cycle. Mitochondria go through continuous cycles of fusion and fission. Fusion is brief (1). It triggers a fission
event within a period of seconds to a few moments (2). A daughter mitochondrion may maintain an intact membrane potential (orange) or depolarize
(3, green). When depolarized, a subsequent fusion event is unlikely to take place, unless the mitochondrion repolarizes. Depolarized and solitary
mitochondria (4) remain for several hours in a pre‐autophagic pool (5) prior to removal by autophagy. Adapted from [15] with permission of Elsevier.

Initiation of
mitochondrial Hyperfusion of
network fusion mitochondrial
network
GLOBAL
Autophagy CONTROL

Mitochondrial
network begins
fragmentation
Recovery of
membrane
potential
Solitary
period

De- Intact
polarization membrane
potential Mitochondrial
fragmentation
and dispersion
to opposite
telomeres

Fission

LOCAL
CONTROL Fusion

Reorganization
Sequestration
@ 2009 VisuallyMedical.com

Figure 8.2  Mitochondrial life in the cell cycle. This diagram depicts the normal life cycle of an individual mitochondrion during the G0 phase of
the cell cycle. The mitochondrion undergoes fusion, fission, depolarization, and degradation by autophagy. This process is depicted as one of local
control whereby mitochondrial events are largely dictated by the local energetic status and associated local signals. During the cell cycle global
signals cause concerted changes in the mitochondrial population, as noted by hyperfusion in G1–S and fragmentation during M phase. These global
population effects are governed by the cell’s demand for energy required by cell division and the need for homogenization and sequestration of
cellular components during met‐phase. The cell cycle serves as an elegant example of the parities of local and global control. Adapted from [16]
with permission of Elsevier.
8:  Mitochondrial Function, Dynamics, and Quality Control 91

The selective elimination of defective mitochondria, or fusion, with the intracellular architecture looking like an inter-
mitophagy, is a critical housekeeping process within a cell [25]. connected network in some cell lines, supported the analogy
Mfn2 is also the target of ubiquitination by a second, redundant that OXPHOS is executed in all mitochondria as a choir of
pro‐mitophagy ubiquitin ligase, MUL1, whose loss exacerbates 88–2000 people singing a song in unison.
the Parkin‐null phenotype [28]. This is a tightly regulated process, However, because different cellular processes with different
where deubiquitinases such as ubiquitin C‐terminal hydrolase L1 ATP demands can occur in different intracellular locations, it
(UCH‐L1) act in opposition to the ubiquintation of Mfn2 and other would make physiological sense that mitochondria in different
proteins [30, 31]. Mitophagy is well‐controlled but can fail at any locations might be functioning differently or executing different
of the above steps. Although the above mechanisms describe the anabolic processes. A classic example is the different mitochon-
targeting of defective mitochondria to the mitophagic process, drial subpopulations identified in muscle, with mitochondria
mitophagy also requires proper lysosomal function, which has located between muscle fibers having a higher capacity to syn-
been shown to be impaired in conditions of excess nutrients [32]. thesize ATP to support contraction compared to mitochondria
Loss of mitochondrial dynamics proteins leads to diverse located below the fiber membrane (subsarcolemmal). In the adi-
changes in mitochondrial function. For example, tissue-specific pose tissue, a subpopulation of mitochondria that are uniquely
ablation of Drp1 in liver protected against diet‐induced obesity adherent to the lipid droplet has recently been identified
and glucose intolerance while increasing energy expenditure in (Figure  8.3). These were named peridroplet mitochondria
mice [33]. Similarly, liver-specific loss of mitofusin 1 increased (PDM) and have also been found in the liver (Figure 8.4).
mitochondrial mass and shifted their fuel preference to the burn-
ing of fatty acids, while protecting from the effects of diet‐induced What are the peridroplet mitochondria?
obesity [34]. Conversely, loss of mitofusin 2 resulted in impaired
glucose metabolism, decreased lipid accumulation, glucose toler- PDM are mitochondria that reside within 0.5 μm of lipid drop-
ance in liver and other tissues, and increased hepatic glucose lets (LDs). Electron microscopy studies of adipose cells have
­production [20, 35]. Taken together, these data suggest a role of shown that PDM are characterized by an elongated morphol-
morphology in controlling fuel preference of mitochondria. ogy, increased electron density in the mitochondria–LD contact
site, and cristae oriented perpendicular to the axis of mitochon-
dria–LD contact site. This unique morphology supported the
concept that PDM play a specialized role in fat metabolism.
SUBCELLULAR INDIVIDUALISM However, it was unclear whether PDM promote fat oxidation,
AND EXCLUSIVE CLUBS fat storage, or both.
OF MITOCHONDRIA Recent studies have demonstrated that PDM are relatively
immobile and do not fuse with cytoplasmic mitochondria (CM)
The total number of mitochondria per cell varies in different tis- and thereby maintain a distinct proteome and bioenergetic
sues and cell types, with reports estimating between 88 and ­function. Compared to CM, PDM have higher levels of
2000 mitochondria per cell. The discovery of mitochondrial ATP  synthase and cytochrome c oxidase protein levels and

BAT Low-speed High-speed


Dissection centrifugation centrifugation Peridroplet
Fat cake mitochondria (PDM)
(900 × g) (9,000 × g)

Supernatant Cytoplasmic
mitochondria (CM)

Nuclei and (Classical isolation)


unbroken cells

Figure 8.3  Isolation of peridroplet mitochondria by differential centrifugation. Schematic representation of peridroplet mitochondrial (PDM)
isolation technique. Tissue is dissected and homogenized with a glass–Teflon dounce homogenizer. Low‐speed centrifugation separates the fat cake
containing PDM from supernatant containing cytoplasmic mitochondria (CM). High‐speed centrifugation strips PDM from lipid droplets (LDs)
and pelleted CM mitochondria from the supernatant. Note that some mitochondrial isolation protocols discard the fat cake and/or begin with high‐
speed centrifugation step. LDs were marked by the neutral BODIPY 493/503 fluorescent dye (BODIPY) and mitochondria by MitoTracker deep
red dye (MitoTracker). Note the tubular structures staining positively for MitoTracker on LDs. Adapted from [36] with permission of Elsevier.
92 THE LIVER:  REFERENCES

Lean HFD ob/ob

Figure 8.4  Mitochondrial association with lipid droplets can be observed in the liver. The association is particularly noticeable in livers from mice
on a high‐fat diet (HFD) and in ob/ob mice, although this has not been statistically validated. Adapted from [37] with permission of Springer Nature.

enzymatic capacity. Importantly, PDM have higher pyruvate 4. Schmidt, O., Pfanner, N., and Meisinger, C.. Mitochondrial protein import:
oxidation capacity but lower fatty acid oxidation capacity com- From proteomics to functional mechanisms. Nat Rev Mol Cell Biol,
2010;11(9):655–67.
pared to CM. Furthermore, cells with increased levels of PDM 5. Boyd, M.E., Albright, E.B., Foster, D.W., and McGarry, J.D. In vitro reversal
have higher levels of LD expansion and triacyglyceride synthe- of the fasting state of liver metabolism in the rat. Reevaluation of the roles of
sis. The current evidence supports the concept that PDM pro- insulin and glucose. J Clin Invest, 1981;68(1):142–52.
mote LD expansion through ATP‐dependent triacylglyceride 6. Garland, P.B., Shepherd, D., Nicholls, D.G., and Ontko, J. Energy‐dependent
control of the tricarboxylic acid cycle by fatty acid oxidation in rat liver
synthesis, which may protect mitochondria from lipotoxicity in
mitochondria. Adv Enzyme Regul, 1968;6:3–30.
periods of lipid excess. 7. Konig, T., Nicholls, D.G., and Garland, P.B. The inhibition of pyruvate and
Ls+‐isocitrate oxidation by succinate oxidation in rat liver mitochondria.
Biochem J, 1969;114(3):589–96.
Peridroplet mitochondria in the liver 8. McGarry, J.D., Takabayashi, Y., and Foster, D.W. The role of malonyl‐CoA
in the coordination of fatty acid synthesis and oxidation in isolated rat
The liver is tasked with the rapid packaging of lipids into lipo-
­hepatocytes. J Biol Chem, 1978;253(22):8294–300.
proteins for distribution to other tissues in the body. However, 9. Lapuente‐Brun, E., Moreno‐Loshuertos, R., Acin‐Perez, R. et  al.
under conditions of nutrient excess, lipids accumulate within Supercomplex assembly determines electron flux in the mitochondrial elec-
hepatocytes, resulting in a chronic state of inflammation and tron transport chain. Science, 2013;340(6140):1567–70.
cirrhosis. The need for rapid and effective lipid synthesis in 10. Campanella, M., Casswell, E., Chong, S. et al. Regulation of mitochondrial
structure and function by the F1Fo‐ATPase inhibitor protein, IF1. Cell
hepatocytes suggests a potential role of PDM in maintaining Metab, 2008;8(1):13–25.
lipid homeostasis of hepatocytes. 11. Lewis, M.R. and Lewis, W.H. Mitochondria (and other cytoplasmic
Although PDM have not been specifically studied in liver, ­structures) in tissue cultures. Am J Anat, 1915;17(3):339–401. https://doi.
evidence suggests the existence of PDM in hepatocytes. For org/10.1002/aja.1000170304
example, published electron microscopy images show levels of 12. Sesaki, H. and Jensen, R.E. Division versus fusion: Dnm1p and Fz01p antag-
onistically regulate mitochondrial shape. J Cell Biol, 1999;147(4):699–706.
PDM in both genetic and high‐fat diet mouse models of obesity https://doi.org/10.1083/jcb.147.4.699
[37] (Figure 8.4). In addition, high‐fat feeding in both mice and 13. Shaw, J.M. and Nunnari, J. Mitochondrial dynamics and division in budding
humans has been consistently associated with increased expres- yeast. Trends Cell Biol, 2002;12(4):178–84. https://doi.org/10.3174/ajnr.
sion of perilipin5 (Plin5), an LD coat protein that strongly A1256.Functional
14. Yaffe, M.P. The machinery of mitochondrial inheritance and behavior.
recruits mitochondria to LDs [38]. Cell and animal models have
Science (New York), 1999;283(5407):1493–7.
confirmed that overexpression of Plin5 in the liver expands LD 15. Liesa, M. and Shirihai, O.S. Mitochondrial dynamics in the regulation of
mass and protects against lipotoxic liver injury. This provides nutrient utilization and energy expenditure. Cell Metab, 2013;17(4):
further support for the concept that PDM support lipid homeo- 491–506.
stasis in hepatocytes. 16. Hyde, B.B., Twig, G., and Shirihai, O.S. Organellar vs cellular control of
mitochondrial dynamics. Semin Cell Dev Biol, 2010;21(6):575–81.
17. Bach, D., Pich, S., Soriano, F.X. et al. Mitofusin‐2 determines mitochondrial
network architecture and mitochondrial metabolism. A novel regulatory
mechanism altered in obesity. J Biol Chem, 2003;278(19):17190–7.
REFERENCES 18. Le Blanc, S., Villarroel, P., Candia, V. et al. Type 2 diabetic patients and their
offspring show altered parameters of iron status, oxidative stress and genes
1. Benda, C. Ueber die Spermatogenese der Vertebraten und höherer related to mitochondrial activity. Biometals, 2012;25(4):725–35.
Evertebraten. II. Theil: Die Histiogenese der Spermien. Verh Berl Physiol 19. Pich, S., Bach, D., Briones, P. et al. The Charcot‐Marie‐Tooth type 2A gene
Ges, 1898:393–8. product, Mfn2, up‐regulates fuel oxidation through expression of OXPHOS
2. Koehler, C.M. Protein translocation pathways of the mitochondrion. FEBS system. Hum Mol Genet, 2005;14(11):1405–15.
Lett, 2000;476(1–2):27–31. 20. Sebastian, D., Hernandez‐Alvarez, M.I., Segales, J. et al. Mitofusin 2 (Mfn2)
3. Perry, A.J., Hulett, J.M., Likić, V.A., Lithgow, T., and Gooley, P.R. Convergent links mitochondrial and endoplasmic reticulum function with insulin signal-
evolution of receptors for protein import into mitochondria. Curr Biol, ing and is essential for normal glucose homeostasis. Proc Natl Acad Sci U S
2006;16(3):221–9. A, 2012;109(14):5523–8.
8:  Mitochondrial Function, Dynamics, and Quality Control 93

21. Boutant, M., Kulkarni, S.S., Joffraud, M. et al. Mfn2 is critical for brown 30. Costes, S., Huang, C.J., Gurlo, T. et al. β‐cell dysfunctional ­ERAD/­ubiquitin/
adipose tissue thermogenic function. EMBO J, 2017;36(11):1543–58. proteasome system in type 2 diabetes mediated by islet amyloid polypeptide‐
https://doi.org/10.15252/embj.201694914 induced UCH‐L1 deficiency. Diabetes, 2011;60(1):227–38.
22. Mahdaviani, K., Benador, I.Y., Su, S. et al. Mfn2 deletion in brown adipose 31. Costes, S., Gurlo, T., Rivera, J.F., and Butler, P.C. UCHL1 deficiency exac-
tissue protects from insulin resistance and impairs thermogenesis. EMBO erbates human islet amyloid polypeptide toxicity in B‐cells. Autophagy,
Rep, 2017;18(7):1123–38. 2014;10(6):1004–14.
23. Dagda, R.K., Cherra, S.J., Kulich, S.M. et  al. Loss of PINK1 function 32. Trudeau, K.M., Colby, A.H., Zeng, J. et al. Lysosome acidification by photo-
­promotes mitophagy through effects on oxidative stress and mitochondrial activated nanoparticles restores autophagy under lipotoxicity. J Cell Biol,
fission. J Biol Chem, 2009;284(20):13843–55. 2016;214(1):25–34.
24. Twig, G., Elorza, A., Molina, A.J.A. et  al. Fission and selective fusion 33. Wang, L., Ishihara, T., Ibayashi, Y. et al. Disruption of mitochondrial fission
­govern mitochondrial segregation and elimination by autophagy. EMBO J, in the liver protects mice from diet‐induced obesity and metabolic deteriora-
2008;27(January), 433–46. tion. Diabetologia, 2015;58(10):2371–80.
25. Chen, Y. and Dorn, G.W.. PINK1‐phosphorylated mitofusin 2 is a Parkin 34. Kulkarni, S.S., Joffraud, M., Boutant, M. et al. Mfn1 deficiency in the liver
receptor for culling damaged mitochondria. Science (New York), protects against diet‐induced insulin resistance and enhances the hypoglyce-
2013;340(6131):471–5. mic effect of metformin. Diabetes, 2016;65(12):3552–60.
26. Pickrell, A.M. and Youle, R.J. The roles of PINK1, parkin, and mitochon- 35. Chen, X. and Xu, Y. Liver‐specific reduction of Mfn2 protein by RNAi
drial fidelity in Parkinson’s disease. Neuron, 2015;85(2):257–73. https://doi. results in impaired glycometabolism and lipid homeostasis in BALB/c mice.
org/10.1016/j.neuron.2014.12.007 J Huazhong Uni Sci Technol Med Sci 2009;29(6):689–96.
27. Song, M., Mihara, K., Chen, Y., Scorrano, L., and Ii, G.W.D. Mitochondrial 36. Benador, I.Y., Veliova, M., Mahdaviani, K. et al. Mitochondria bound to lipid
fission and fusion factors reciprocally orchestrate mitophagic culling in droplets have unique bioenergetics, composition, and dynamics that support
mouse hearts and cultured fibroblasts. Cell Metab, 2015;21(2):273–85. lipid droplet expansion. Cell Metab, 2018;27(4),869–85.e6.
28. Yun, J., Puri, R., Yang, H. et al. MUL1 acts in parallel to the PINK1/parkin 37. Arruda, A.P., Pers, B.M., Parlakgül, G. et al. Chronic enrichment of hepatic
pathway in regulating mitofusin and compensates for loss of PINK1/parkin. endoplasmic reticulum‐mitochondria contact leads to mitochondrial dys-
ELife, 2014;3:e01958. function in obesity. Nature Med, 2014;20(12):1427–35.
29. Narendra, D.P., Jin, S.M., Tanaka, A. et al. PINK1 is selectively stabilized on 38. Wang, C., Zhao, Y., Gao, X. et al. Perilipin 5 improves hepatic lipotoxicity
impaired mitochondria to activate Parkin. PLoS Biol, 2010;8(1):e1000298. by inhibiting lipolysis. Hepatology, 2015;61(3):870–82.
Nuclear Pore Complex
9 Michelle A. Veronin and Joseph S. Glavy
University of Texas at Tyler, Fisch College of Pharmacy, Tyler, TX, USA

OVERVIEW NUCLEAR PORE COMPLEX


In eukaryotic cells, the separation of nuclear DNA from The NPC is a large macromolecular assembly with an esti-
cytoplasmic organelles and cellular materials is governed by mated size of 110 MDa in vertebrates and 60 MDa in yeast
the nuclear envelope. The nuclear envelope (NE) is defined (Figure 9.1) [11–14]. NPCs possess a symmetric core with an
as a specialized endoplasmic reticulum (ER) membrane with octahedral arrangement across the double membrane of the
a double bilayer containing an inner and outer membrane nuclear envelope, resembling the spokes of a bicycle wheel. They
system. Although this partitioning protects the genome, it are composed of specialized proteins termed nucleoporins or
also is needed to organize nuclear entry and exit for basic Nups. Proteomic studies in yeast and metazoa cataloged the
cellular processes (Figure  9.1). Nuclear pore complexes Nups of the NPC [4, 5]. Mammalian NPCs contained at least
(NPCs), the gateways for macromolecular traffic into and seven additional Nups, including ALADIN, Nup358, Pom210,
out of the nucleus, are set in circular openings of the double Pom121, NDC1, Nup43, and Nup37 (Figure  9.2). So only
membrane of the NE [1, 2]. There are highly regulated path- around 30 types of Nups exist in NPCs but they are present in
ways that control nuclear entry and exit of molecules such as multiple copies, which adds to the puzzle of this large trans-
transcription factors, RNAs, kinases, and viral particles. In porter. Nups can be classified into six groups: (i) Y‐complex
broad spectrum, proteins with their transport signaling (coat) Nups, (ii) adaptor Nups, (iii) channel Nups, (iv) cyto-
sequence to be imported or exported from the nucleus (i) plasmic filament Nups, (v) nuclear basket Nups, (vi) mobile
bind to transport receptors, (ii) are transported through Nup98, and (vii) POM (transmembrane) Nups (Figure  9.2).
NPCs present in the NE, and (iii) translocate from the NPC Nups have structural designs to interlock (coiled coils, β‐pro-
to intranuclear or cytoplasmic target sites [1, 3]. About a pellers or disordered areas) while channel Nups contain flexi-
thousand proteins are needed to build one NPC. NPCs are ble repeating units of phenylalanine‐glycine (FG) (forming sea
composed of proteins termed nucleoporins or Nups, which anemone‐like fingers within the NPC which are implicated in
play a role in the structure of the NPC and in regulating the active nuclear transport) [1, 15–17]. Nups are basically named
translocation of molecules through the NPC [1, 3]. And the by their designated molecular weight and are organized into
composition of NPCs differs from species to species [4–6]. macromolecular assemblies called subcomplexes. Nup sub-
Abnormalities such as chromosomal translocations involv- complexes release from the NPC during open mitosis and are
ing genes encoding Nups have been associated with many the disassembly units of the NPC [18]. In the laboratory, inter-
forms of leukemia [7]. Proteins of the NPC have been asso- phase subcomplexes can be isolated through biochemical
ciated with cancer, Huntington’s disease, micronuclear for- extraction of the NE with low percentage non‐ionic or zwitteri-
mation, viral infection, aging, triple A syndrome, and onic detergent treatment (such as 2% Triton X‐100 and/or 1%
primary biliary cirrhosis [7–10]. CHAPS) [19–21]. A key point of NPCs is that these modular

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
9:  Nuclear Pore Complex 95

(a) (b)

Figure 9.1  (a) Thin‐section electron microscopy illustrating the nuclear pore complex (NPC) in isolated HeLa cell nuclear envelope (NE).
Courtesy of Dr. Samuel Dales. (b) Artist’s rendition of NPC. Courtesy of Grace Glavy.

Cytoplasm

Nup358
Nup358

hCG1 hCG1
ALADIN ALADIN
Rae1 Rae1
Nup214 Nup214
Nup88 Nup98 Nup98 Nup88
Nup43 Nup160 Nup160 Nup43
Nup37 Nup133 Nup133 Nup37
Sec13 Nup107 Nup107 Sec13
Seh1Nup96 Nup96 Seh1
Nup75 Nup75
gp210 Nup205 Nup205 gp210
Nup188 Nup62 Nup62
Nup188
NDC1 Nup155 Nup58 Nup58 Nup155 NDC1
Nup93 Nup54 Nup54 Nup93
Nup Nup
pom121 35 pom121
35
Nup160 Nup160
Seh1 Nup133 Nup133 Seh1
Sec13 Nup107 Nup107 Sec13
Nup43 Nup96 Nup96 Nup43
Nup37 Nup75 Nup75 Nup37
Nup
153

Nup
153

Nup98 Nup98
Rae1 Nup50 Nup50 Rae1

Tpr Tpr

Nucleoplasm
Figure 9.2  Overall schematic of subcomplexes and unique non‐subcomplex proteins: Y‐Nups (dark blue), adaptor Nups (light blue), channel
Nups (violet), cytoplasmic filament Nups (green), pore membrane (POM) Nups (light brown), nuclear basket Nups (orange and dark brown), and
mobile Nup98/Rae1.

units are present in multiple copies arranged around two‐ and estimated at less than 50 nm while the envelope is approxi-
eightfold axes of symmetry and form discrete structures within mately 50 nm [23]. The cytoplasmic rings (CRs) are believed to
the NPC [15]. Overall NPC architecture is conserved between act as a docking site for protein transport and bind to nuclear
yeast and higher eukaryotes with an eightfold symmetry [15, 22]. rings (NRs) [11–13, 23, 24]. An early approach using a combi-
Cryoelectron tomography (cryo‐ET) with three‐dimensional nation of proteomic data with a computational platform has
reconstruction gives a resolution of 23 Å (Figure 9.3) [11, 12]. been applied to the architectures of the macromolecular assem-
The nuclear basket extends nearly 60 nm into the nucleus blies of the NPC [25, 26]. The study concludes that the funda-
(Figure 9.3) [11–13, 23, 24]. The central core inner ring (IR) is mental symmetry unit of the NPC is the spoke [25, 26]. Within
96 THE LIVER:  NUCLEOPORINS (NUPS)

the ringed structure are linked units and flexible repeat units that Early reconstitution experiments in yeast were assembled
both stabilize and are involved in transport [1]. The inner and into a Y‐complex visualized through negative staining electron
outer rings help to facilitate the membrane structure (Figure 9.3). microscopy [35]. The structural modules of the subcomplex are
The complete architecture of the NPC includes some non‐Nup a collection of α‐helical repeats and β‐propellers. X‐ray crystal
proteins – the nuclear membrane proteins and the nuclear lam- structures have been reported for several members. Analysis of
ina [15]. These non‐Nup proteins make up the surrounding Nup133 revealed a seven‐bladed β‐propeller domain of the N‐
regions of the NPC and support its transport function. terminus [36]. The crystal structure of the α‐helical interaction
of the Nup107–133 shows an elongated structure forming a
compact interface in a tail‐to‐tail fashion [36]. This interaction
NUCLEOPORINS (NUPS) is a critical attachment point for Nup133 [36]. Another crystal
structure of Nup96–Sec13 forms a hetero‐octamer [37]. The N‐
terminus of Nup96 invades the six‐bladed β‐propeller structure
Y‐Nups of Sec13, providing a seventh blade [37]. The remaining portion
Y‐Nups are contained within the Y‐complex (Nup107 subcom- of Nup96 forms an antiparallel α‐helical domain. The potential
plex) (Figure 9.2). The cytoplasmic Y‐complex has nine mem- interlocking modules of the subcomplex may be the meshwork
bers: Nup160, Nup133, Nup107, Nup96, Nup75, Nup43, of the overall macromolecule. The potential coating modules
Nup37, Seh1, and Sec 13, whereas the nuclear Nup107 subcom- may form a cylinder layer that apposes the membrane [37].
plex contains a tenth member: ELYS (Figure 9.2) [11, 13, 21, Analysis of the Nup75–Seh1 X‐ray structure led to conclusions
27–30]. This subcomplex has been classified as a keystone of that the scaffold Nups form a membrane‐bordering lattice, pro-
NPC assembly [31]. RNAi knockdown experiments of individ- viding attachment sites for additional Nups such as channel
ual members within the Y‐complex affect select members of the Nups, Nup98, and Nup155 [38].
subcomplex plus some other channel Nups [20, 31]. These find- Through X‐ray crystallography the NPC can be pieced
ings show the interdependence of subcomplex members and the together bit by bit with the Y‐complex as the main component.
overall NPC. Positioned at the curvatures of the membrane Larger crystal approaches has yielded great progress. A yeast
embedding the NPC, the Y‐complex acts to stabilize these bends hexameric Y‐complex was achieved using a single‐domain
in the membrane (Figure  9.3) [32]. In the “protocoatomer” synthetic antibody crystallization chaperone [39]. Among the
hypothesis, components of the Y‐complex serve as membrane‐ information provided was further proof of an evolutionarily
curving modules similar to the members of the COPI, COPII, conserved ring structure formed by the yeast Y‐complex. Our
and clathrin complexes [33]. One member of the Y‐complex, cryo‐ET (approximately 23 Å) work has shown that the human
Sec13, is also found as key member of the COPII complex. Y‐complex forms two reticulated rings on the cytoplasmic and
Furthermore, ArfGAP1 lipid packing sensor (ALPS) domains again on the nuclear side [12–14] (Figure 9.3). Numerous dis-
are found within members of the subcomplex [25, 34]. ALPS crete crystal structures of the Y‐complex fit directly into the
domains are classified as membrane‐binding amphipathic α‐ tomographic map, including the yeast hexameric structure [11–
helix regions. Nup133 contains an ALPS domain shown to bind 14]. In our integrated approach, we coupled electron tomogra-
to isolated membrane [34]. Additional Nups, including Nup107, phy, single‐particle electron microscopy, and cross-linking mass
Nup160, and Nup188, contain predicted ALPS domains [25]. spectrometry to show that 32 copies of the Y‐complex amass

CRs

ONM

LC

INM

NRs

Nup205/ Outer Inner


Nup155 Nup93 Nup62 Nup58 Nup54
Nup188 Y complex Y complex

Figure 9.3  Cross‐section of the nuclear pore complex from the cytoplasm (top) to the nucleoplasm (bottom). Structure is based on cryoelectron
tomography. The curvature of the nuclear envelope (NE) is shown: outer nuclear membrane (ONM), luminal curve (LC), and inner nuclear mem-
brane (INM). The NPC is labeled: cytoplasmic rings (CRs) and nuclear rings (NRs).
9:  Nuclear Pore Complex 97

into two reticulated rings, one at each of the cytoplasmic and Nuclear basket Nups
nuclear faces of the NPC [11]. This twin‐ring organization of
the Y‐complex leaves an explanation for structural plasticity of Nup153 and Nup50, along with Tpr (translocated promoter
the NPC [11]. Flexible and spring‐shaped hinges provide for region protein), make up the nuclear basket and provide the sur-
large‐scale rearrangements that might be relevant for large unit face to utilize binding areas for transport. Tpr is an unusual Nup
transport [11] (Figure 9.3). Furthermore, we connected the cyto- (nucleoporin Tpr) with more filament protein properties than
plasmic filaments Nups to the Y‐complex [11–14] (see the cyto- most yet required for trafficking across the nuclear envelope.
plasmic filaments section). Tpr acts as a framework component in the nuclear segment and
tethering chromatin to begin perinuclear heterochromatin exclu-
sion zones. Tpr is also believed to act as docking site for express-
Adaptor Nups ing genes interacting with select Nups. It participates in both
The adaptor complex (formerly called the Nup93 subcom- nuclear import and export pathways for proteins with or without
plex) includes five members: Nup93, Nup205, Nup188, nuclear export sequences (NES) as well as the export of mRNA
Nup155, and Nup35 [40]. The X‐ray crystal structure of [44, 45]. Tpr coiled‐coil domains help give a reliable support to
Nup93 reveals an elongated, α‐helical structure  [41]. This form and maintain the nuclear basket. Tpr plays a role in mitotic
form is evolutionarily conserved and therefore functionally spindle checkpoint signaling [44, 45]. Within the nuclear
maintained [41]. Members of the adaptor complex contain basket, Nup153 is associated with Tpr and contains four zinc
mainly α‐helical domains [41]. Just like the Y‐complex, fingers. These zinc fingers increase the local Ran concentration
Nup93 is a highly abundant protein with 32 copies within the to assist nuclear transport [46]. Along with Nup50, Nup153
NPC [30]. The Nup93 subcomplex aids in IR stabilization helps to terminate karyopherin‐mediated transport (Figures 9.2
[41] and is needed for correct nuclear pore assembly and and 9.4) [1].
homeostasis of the NPC [41]. RNAi experiments suggest a
functional link between NE transmembrane NDC1, Nup93, Cytoplasmic filament Nups
and Nup205. These results suggest an anchor point function
for the Nup93 subcomplex [42]. This has been confirmed by As in the case of channel Nups, cytoplasmic filament Nups pos-
structure studies that Nup93 is a key component to the IR. sess FG regions. Their floppy tentacle nature is cohesive with
Furthermore, Nup62 (channel Nup) has been shown to inter- crystal formation. Thus far, only non‐FG regions of these Nups
act with Nup93, illustrating interdependence between IR have been reported. The X‐ray crystal structure of the non‐FG
Nups and the channel Nups [42]. Using an integrated repeat N‐terminus of Nup214 reveals a seven‐bladed β‐propel-
approach, we found in our studies that 32 copies of both ler with a segment of its C‐terminus bound to the propeller [47].
Nup188 and Nup205 (Figure  9.3, yellow), Nup93 (red), Furthermore, X‐ray analysis of Nup58/45 revealed a possible
Nup155 (green), and the channel Nups (next section: Nup62, circumferential sliding mechanism to adjust the diameter of the
Nup58, and Nup54) fit into the IR with additional Nup155 central transport channel [43]. The α‐helical region forms dis-
protein reaching up and connecting to the outer ring tinct tetramers with a hydrophobic interface. The residues are
(Figure 9.3, green). As shown in Figure 9.3, the IRs comprise laterally displaced in numerous tetramer confirmations, giving
rings similar to the Y‐complex, confirming their evolutionary the possibility of a sliding structure [43]. Selective knockdown
connection [12]. This resolves the subdomain of Nup205 and of Nup214 followed by cryo‐ET demonstrates that this subcom-
Nup188 and the asymmetric structures of Nup93 into the IR plex protrudes into the cytoplasmic ring region [11]. This posi-
[12] and links channel Nups to the ring structure [12]. tion secures FG repeats onto the framework of the rings to
facilitate nuclear transport [11]. Further application of gene‐
silencing/cryo‐ET uncovered a role for Nup358 to stabilize
Channel Nups solely the cytoplasmic reticulated double ring structure  [13].
Channel Nups have stretches of FG (Phe–Gly) repetitive resi- These findings change the edges between Y and cytoplasmic
dues, which are separated by polar spacer regions of variable filament Nups [13].
lengths (Figure 9.1b) [1, 16]. FG repeat domains form unstruc-
tured regions that form weak interactions with transporting pro-
Mobile Nup: Nup98
teins called karyopherins (kaps) [1]. Channel Nups such as the
Nup62 subcomplex are located in the inner pore region or cen- Nup98 is found both at the nuclear pore complex and within the
tral core IR [23, 24]. The Nup62 subcomplex includes Nup62, nucleus [27]. It has multiple functions and binding partners
Nup58–Nup45, and Nup54 (Figures 9.2 and 9.3) [15, 43]. This [48]. Nup98 has roles in transport, mitotic progression, gene
subcomplex was referred to as the central plug region of the expression, epigenetic changes, and viral infection [9, 48–52].
NPC. Although these transport Nups line the inner NPC, it is Nup98 arises from a Nup98‐–Nup96 precursor form that splits
unlikely that they form a plug against transport, but rather a by a self‐cleavage domain similar to those found in Drosophila
dynamic transport area of the complex. The FG repeat domains hedgehog and Flavobacterium glycosylasparaginase [27, 53]. It
form tentacle‐like structures that emanate from and line the is classified as a non‐subcomplex Nup with multiple locations
channel of the pore (discussed further in the transport section). along both sides of the NPC (Figure  9.3) [27]. Nup98 ferries
In Figure 9.3, they are shown to line the IR region. Not shown between internuclear regions to NPCs while regulating tran-
is  how they extend into the channel and would be ready to scription of select genes implicated in development and growth
receive cargo. [9, 54–57]. The N‐terminus of Nup98 contains FG repeats used
98 THE LIVER:  NUCLEOPORINS (NUPS)

α-kap
D
1 T D
6
Importin T
β-kap
2

3
NLS-cargo

Cytoplasm

Nucleus

T
T RanGTP T
4
D RanGDP T
Exportin
β-kap

Figure 9.4  Karyopherin‐dependent transport cycle. Key components for the cycle are RanGTP, nuclear localization sequence (NLS)–cargo, kaps,
and nuclear pore complex (NPC). Key concept is that the concentration of RanGTP is much higher in the nucleus. (1) Formation of the α‐kap and
β‐kap transport complex. (2) α‐Kap recognition of NLS–cargo with β‐kap. (3) Selective transport through NPC. (4) Once inside, the high levels of
nuclear RanGTP increase the likelihood of interaction and hence, disassembly of the transport complex. (5) GTP‐bound β‐kap transport through
the NPC to the cytosol, while α‐kap forms a complex with GTP‐bound export β‐kap, then exits by way of the nearby NPC. Left behind is the NLS–
cargo. (6) Once back in the cytosol, GTP undergoes hydrolysis and kaps are recycled for the next round of transport.

as a docking site for kaps, but is not classified as an FG Nup but lined with select membrane proteins believed to anchor NPCs
rather as a mobile Nup [27, 58]. Nup98 facilitates mRNA export like the settings of a ring. POM proteins are involved in the ini-
from the nucleus [27, 59]. Mitotic phospho‐Nup98 is a deter- tiation of pore complex formation, stabilization, release and ref-
mining factor in NPC disassembly [56, 57]. One of Nup98’s ormation of the NPC. To date, four POM proteins have been
interacting partners is Rae1; together they act as temporal regu- classified through proteomic and genetic screenings: gp210,
lators of the anaphase‐promoting complex [60]. Vesicular sto- Pom121, NDC1, and Pom33. The largest POM protein, gp210
matitis virus (VSV) matrix (M) protein binds Nup98 and inhibits (Pom210), contains a single transmembrane domain as do the
nuclear export of mRNA [59]. In some instances, inhibition has rest of the POM proteins. NDC1 is found both in the POM
been linked to activation of p53 [61]. Nup98 gene translocations region and the spindle pole body (SPB) during mitosis. In yeast
have been linked to several forms of leukemia (covered in it helps to anchor the SPB at the NE [63], but in human cells its
section on NPC and disease) [9]. Nup98 has also been impli- purpose is still unknown. RNAi experiments suggest a func-
cated in nuclear envelope breakdown (NEBD) (see later NEBD tional link between NE membrane NDC1 and two members of
section) [56, 62]. the adaptor complex: Nup93 and Nup205 [64]. These results
suggest an anchor point function for the Nup93 subcomplex
[42]. Elimination of Pom121 results in failure to assemble NPCs
Pore membrane Nups and formation of continuous nuclear membranes [65]. Cross-
The NE is separated into three domains: the outer nuclear mem- linking mass spectrometry (XL‐MS) reveals a direct linkage
brane (ONM), the pore membrane (POM), and the inner nuclear between Nup133 and Pom121 at the POM interface [11].
membrane (INM) [2]. The POM region is the sector of the NE Although no single POM or combination of POMs is com-
where INM and ONM fusion occurs. The POM curvatures are pletely indispensable for survival, implying structural/
9:  Nuclear Pore Complex 99

functional redundancy of diverse transmembrane and soluble that contributes to nuclear stability [79]. The inner membrane of
Nups, nuclear import and soluble Nup localization is most the NE in mammalian cells contains a unique array of integral
­perturbed in cells lacking NDC1 and Pom121 [64, 66, 67]. membrane proteins including lamina‐associated polypeptides 1
POM‐depletion studies point to the need for additional compo- and 2 (LAP1 and LAP2), emerin, MAN1, nesprin‐1, nesprin‐2,
nents to restore overall POM function. NDC1 was shown to and lamin B receptor (LBR) [2]. These proteins contain an LEM
form a linkage between the NE and soluble Nups [64, 68, 69] domain (LAP2, emerin, and Man1 domain), which is a 40‐residue
and depletion of NDC1 shows markedly reduced channel Nup motif that interacts specifically with lamin A/C [80].
staining compared to NDC1/control siRNA treatment [64, 69]. When the NE breaks down during mitosis, the nuclear lamina
These results point to a crucial role for NDC1 in structure, func- also disassembles, which occurs when lamins are phosphoryl-
tion, NEBD, and assembly. POM protein “put‐back” experi- ated by protein kinase C. During this disassembly, A‐type
ments fail to restore function, implying the need for additional lamins are dispersed throughout the cytoplasm while B‐type
components as well as POM proteins [64, 69, 70]. We recognize lamins remain attached to the nuclear membrane. Because lamin
Pom121 is required for interphase assembly and localized with A is not farnesylated, it is more soluble than lamin B [78].
Nups at forming pores [71]. In fact, a fragment of Pom121 has a During NE reassembly, lamins are dephosphorylated by type 1
dominant‐negative effect on pore assembly [72]. Thus Pom121 protein phosphatase to facilitate reassembly of the lamina, and
plays an important role in assembly and nuclear pore biogene- A‐type and B‐type lamins are transported into the nucleus [78].
sis. The protein’s repeat‐containing POM domain is involved in Lamins have also been shown to play an important structural
anchoring components of the pore complex to the pore mem- role in determining nuclear morphology and stability. Multiple
brane (Figure 9.2) and when overexpressed in cells induces the studies have demonstrated that lowered expression of functional
formation of cytoplasmic annulate lamellae [73, 74]. Interactions lamins results in nuclei that are more fragile and prone to defor-
of Pom121 with Nup155 and Nup160 are foreseen to contribute mation [78, 81]. Evidence for the platform role of lamins in the
to the formation of the nuclear pore as well as the fastening of formation of multiprotein complexes that function within the
the NPC to the pore membrane [67]. nucleus is supported by their wide variety of interaction
The newest POM protein, Pom33, is required for proper NPC partners, such as retinoblastoma 1 and c‐Fos [78]. Lamins are
distribution [75, 76] as well as assembly. But a percentage of also involved in facilitating the connection between the nucleus
Pom33 resides in the ER [75, 76]. Although it is an element of and the cytoplasm through the LINC complex, which contains
the tubular ER network, Pom33 (TMEM33) also acts as a POM lamin‐interacting proteins, as well as transducing signals from
protein at NPCs [75, 76]. The C‐terminus of Pom33 contains the cytoskeleton into the nucleus through contact with NPCs
several amphipathic regions that both bind to Kap123 and serve [78, 82].
as attachment points, aiding in positioning the POM region In addition, it has been shown that lamins may be involved in
while N‐terminus transmembrane regions anchor the POM [76]. DNA repair. One study demonstrated that expression of disease‐
Phosphorylation of gp210 may be mediated by cyclin‐B– causing mutant lamins alters localization and expression of ATR
cdc2, and depletion of cyclin B from C. elegans embryos also kinase, a regulator of DNA repair, which in turn may adversely
leads to a nuclear‐twinning phenotype [77]. It has been shown affect DNA repair pathways requiring the function of ATR as
that gp210 is important for effective NPC disassembly and sug- well as the additional repair regulator ATM [83]. Furthermore,
gests that phosphorylation of gp210 is an early event in NEBD the presence of mutant lamins was shown to reduce recruitment
[76]. Yet, the role of this set of POM proteins at the onset of of DNA repair factor 53BP1, a p53‐binding protein, in response
disassembly is undetermined. to DNA damage [83]. This suppression of 53BP1 may be con-
nected to degradation by cysteine protease cathepsin‐L (CTSL),
whose upregulation is associated with insufficient expression of
lamin A/C [79]. Substantial DNA damage, as well as alterations
NUCLEAR LAMINA in the genome including breaks in chromatids and chromo-
somes, increased aneuploidy, and telomere defects, have also
In general, lamins fall into the category of either A‐type or B‐ been observed in LMNA‐knockout mouse embryonic fibro-
type lamins, and these classifications are based on differing fea- blasts [79]. A‐type lamins have also been shown to participate in
tures including protein structure and expression patterns [78]. DNA repair of double‐strand breaks by non‐homologous end
For example, A‐type lamins contain an additional C‐terminal joining and homology‐directed repair [79].
domain consisting of a unique 90 amino acid sequence not pre-
sent in B‐type lamins. Additionally, B‐type lamins are expressed
in most cells, whereas A‐type lamins are expressed primarily in
differentiated cells [78]. Three lamin genes are encoded within NUCLEAR TRANSPORT CYCLE
mammalian genomes: LMNA, LMNB1, and LMNB2. At least
seven protein isoforms are produced from alternative splicing; Molecules smaller than 40 kDa can passively diffuse across the
the two major isoforms expressed from LMNA are lamin A and NPC. With higher molecular weight proteins, an orchestrated
C, which are ubiquitous in most differentiated cells [78]. Lamins receptor‐mediated transport is needed to facilitate crossing into
polymerize to form higher order structures, a process involving the nucleus (Figure 9.4) [1, 16, 84, 85]. This process is highly
dimerization from their α‐helical domain followed by head‐to‐ efficient, with evidence to support a potential import rate capac-
tail parallel association between dimers [78]. These higher order ity of 1000 molecules per second per NPC [1, 86, 87]. For
structures form the structural foundation of the nuclear lamina nucleocytoplasmic transport to be carried out via the NPC, the
100 THE LIVER:  NUCLEAR TRANSPORT CYCLE

NPC itself must be dynamic and undergo structural changes, implies an interaction between the integral protein and the
which may occur during varying stages of transport [1, 86]. The channel Nups [93]. The kap, along with its inner membrane
NPC also constitutes a selective transport system between the protein cargo, interacts with Nups across or through scaffolding
nucleus and cytoplasm composed of unstructured regions com- Nup barriers. Fully understanding this inner membrane transport
prising FG repeats, which also act as docking sites for protein represents a significant challenge. Furthermore, the targeted
cargo recognition molecules [88]. These recognition molecules, transport of POM membrane could occur through this cargo‐
known as karyopherins (kaps), bind to cargo in the cytosol and based system. This may require a modification of the NLS or
carry it into the nucleus; conversely, they bind cargo in the additional adaptor proteins to abort transport.
nucleus and deliver it to the cytoplasm. Kaps recognize their In higher eukaryotes, integral membrane proteins can be
cargo by binding short amino acid sequence segments of the targeted to the inner nuclear membrane during mitotic NE
protein [89]. Cargos have nuclear localization sequences (NLS) breakdown [94]. During reformation of the NE in telophase,
that are classically rich in basic residues for import [1, 84, 89, integral membranes may become trapped within the inner
90], and nuclear export sequences (NES) that are classically leu- nuclear membrane, and thus do not fully traverse the NPC.
cine‐rich for export [1, 84]. An NLS or NES may be masked or Targeting mechanisms vary during interphase and for organisms
modified to allow for regulation of protein localization [90]. that undergo a closed mitosis [94]. One model proposed to
Cargo binding occurs through either a direct interaction between explain the process of membrane protein transport across the
β‐kaps and the cargo, or indirect interactions through α‐kap, an NPC is the diffusion–retention model. This model underlines
adaptor protein, and β‐kap [88]. the importance of interactions between membrane proteins
The majority of NPC traffic is achieved by coupling kap‐ and components of the nucleus in membrane protein localiza-
cargo binding to a cycle of GTP‐binding and hydrolysis con- tion, which has been observed in multiple organisms. During
ducted through the protein Ran [91]. Ran plays a key role in transport, transmembrane domains are retained within the pore
maintaining proper transport directionality, which is determined membrane while soluble domains cross lateral channels, regions
by a concentration gradient of RanGTP [88]. Kaps deliver NLS‐ between the pore membrane and NPC scaffold [94–97]. It has
cargo and transport it through the NPC through weak transient been shown that membrane proteins with soluble domains larger
interactions with channel Nups that extend into the NPC. than 60–75 kDa are unable to be transported to the inner nuclear
Because it is a small monomeric G protein, Ran has a low inher- membrane via the NPC, a limitation attributed to the size of the
ent rate of GDP–GTP exchange and GTP hydrolysis [16, 84, 85, narrow lateral channel within the NPC [94, 96, 98]. Another
89, 92], and it relies on interacting proteins to regulate its nucle- model, supported by evidence of diminished accumulation of a
otide state. Binding partners are separated according to their membrane reporter in the inner nuclear membrane following
role in the cycle. Specifically, guanine nucleotide exchange fac- depletion of ATP, suggests that transport occurs based on
tor (GEF) for Ran is concentrated in the nucleus while the continuous structural changes of the NPC that are energy
GTPase‐activating protein (GAP) for Ran is concentrated in the dependent [94, 99].
cytoplasm. As a result, nuclear Ran is in the GTP‐bound form It has been shown through studies using yeast proteins Heh1
while the cytoplasm contains Ran that is predominantly in the and Heh2 that these proteins are targeted to the inner nuclear
GDP‐bound form (Figure 9.4) [85, 89]. This established differ- membrane in a process involving transport receptors Kap60 and
ence is then coupled to import through nucleotide‐dependent Kap95. Once the complex is formed between cargo, Kap60, and
binding of Ran to the kaps, of which the GTP‐bound form binds Kap95, transport across the NPC is facilitated by interactions
α‐kaps. For import, RanGTP will dissociate its cargo, releasing with channel Nups. Once inside the nucleus, the release of cargo
it into the nucleus [92]. Additionally, RanGTP facilitates the from transport receptors is promoted by RanGTP [94]. Three
formation of export complexes within the nucleus [88]. The other membrane proteins have been discovered to contain an
kap–RanGTP complex can then travel back through the NPC to NLS: human Pom121 and SUN2, and C. elegans Unc‐84 [73,
the cytoplasm where RanGTP is hydrolyzed and releases the 94, 100–102]. However, there have been some difficulties with
carrier protein for another cycle with NLS‐cargo binding and these findings, such as the connection between frequency of
import [92]. For export, this process occurs in the converse NLSs in these membrane proteins and their chromatin‐binding
manner. function, the necessity of multiple targeting signals and addi-
Exporting kaps have the ability to bind when already bound tional factors for localization of Unc‐84 and SUN2, respectively
to RanGTP [92]. GTP hydrolysis causes dissociation of the ter- [94, 102, 103]. In addition, it has been shown that varying trans-
nary complex in the cytoplasm, allowing the kaps to travel back membrane proteins exhibit different responses to decreased lev-
through the NPC for another cycle of export and bind to an els of ATP and Ran [94, 102].
NLS‐cargo for import. This process is illustrated in Figure 9.4 Heh1 and Heh2 have been shown to possess NLSs with a
[84, 85, 89]. The key concept is the Ran gradient and the con- high affinity for Kap60, known respectively as h1NLS and
centration of RanGTP is much higher in the nucleus. Therefore, h2NLS [94, 104]. These NLSs are separated from the trans-
once the complex has entered the nucleus, the high concentra- membrane domain by an intrinsically disordered linker sequence
tion allows for the disassembly and release of the components comprising 180 or 235 amino acids [94]. It has been demon-
(Figure 9.4) [84]. strated that h2NLS, in combination with this linker sequence,
Inner nuclear membrane proteins are thought to use a similar acts as a signal for a synthetic transmembrane segment as well
mechanism for movement across the POM and eventual target- as for Sec61 to accumulate within the inner nuclear membrane.
ing [2, 93]. NLSs of the integral proteins signal their transport Furthermore, the length of the linker sequence is associated
through the cargo‐based transport system [2]. This mechanism with inner nuclear membrane accumulation [94, 105]. A model
9:  Nuclear Pore Complex 101

has been proposed based on a study that captured a reporter pro- genome‐wide analysis screening [118, 119]. However it is not
tein, Nsp1, at the anchor domain of a channel Nup. This model yet known if every gene is gated, so more extensive work is
states that upon binding to Kap60/95, h2NLS will travel through necessary to fully prove this hypothesis [105, 109].
the central channel of the NPC as interactions with channel
Nups facilitated by a linker sequence of adequate length occur.
In addition, transmembrane segments will diffuse through the
pore membrane [94]. It has also been shown through work with NUCLEAR ENVELOPE BREAKDOWN
Sec61 fusions that membrane reporters are co-translationally
inserted in the ER membrane rather than transported through the Cell division comprises two processes – mitosis and cytokine-
NPC as soluble proteins. Accumulation of these reporters at the sis. Mitosis is the division of the genetic material into two equal
inner nuclear membrane associated with an NLS and linker halves, whereas cytokinesis is the process by which the cell
sequence was observed [94]. divides itself into two daughter cells. Mitosis is achieved by
condensed chromosomes becoming attached in a bipolar fash-
ion to the microtubule‐based mitotic spindle and subsequent
pulling apart of the sister chromatids to opposite poles of the
GENES, TRANSPORT, AND THE NUCLEAR cell. During interphase, NE surrounds the DNA but this barrier
PORE COMPLEX may not remain throughout mitosis. It is a multifaceted barrier.
Because the mitotic spindles form outside the NE in higher
Proposed by Günter Blobel, the “gene gating” hypothesis pos- eukaryotes, the whole NE must disintegrate to progress through
tulates the idea that transcribed genes are positioned close to the cell cycle [18]. This process is termed nuclear envelope
NPCs, thereby streamlining mRNA export [106]. Due to this breakdown (NEBD) [18, 120]. In some lower eukaryotes, such
vicinity, NPCs potentially play a role in the adjustment of as yeast, the mitotic spindles embed in the inner NE and the
genomic structure corresponding to different stages in devel- process occurs within an intact envelope yielding a “closed”
opment, differentiation, and the cell cycle. It is an important mitosis, while higher eukaryotic NEBD gives an “open” mitosis
assumption that the genome of higher eukaryotes is organized [18]. The difference between the two forms has important impli-
into multiple, distinct three‐dimensional structures. Each of cations in the events triggering the processes in both.
these structures reflects a given differentiated state of the cell. NEBD occurs with the G2/M phase transition at the very onset
Although DNA contains the information necessary for the for- of open mitosis and is a phosphorylation‐dependent process [18,
mation of these structures, the DNA alone would not be suffi- 45, 121, 122]. A pivotal feature of NEBD leading to the opening
cient for the entirety of such a process. Thus, NPCs contribute of the NE and the blending of cellular components is disassem-
to the interpretation of this information and subsequent struc- bly of the NPC. Evidence indicates the mitotic disassembly of
tural organization. Additionally, the nuclear lamina is pro- the NE and the NPC is driven by reversible phosphorylation of a
posed to participate in organization of compacted regions of subset of proteins [21, 42, 45, 123, 124], Nups, lamina, and NE
the genome [106]. NPCs are envisioned as gene gating orga- membrane proteins, which may disrupt structurally significant
nelles capable of interacting specifically with expressing genes interactions. It has been shown that mitosis‐activated cyclin‐
[106]. All transcripts of a particular gene are proposed to exit dependent kinase 1 (CDK1) activity is required for keeping
the nucleus via the NPC to which the gene is gated [106]. The NPCs dissociated during mitosis, while reassembly shows a
nonrandom distribution of the NPC may mirror the underlying phosphatase dependence [125]. In addition to kinase activity, a
genomic organization. Given that the specific conformation of microtubule‐based tearing process assists NE disassembly in
the genome is characteristic for each differentiated state, the cells [126, 127]. A key component in this process is dynein,
distribution of NPCs should be both unique to each differenti- which is recruited to the NE at the beginning of prophase, and
ated state as well as identical for all cells of the same differen- interacts with spindle microtubules [128]. This interaction and
tiated state and cell cycle stage [106]. Furthermore, it is resulting tension placed on the NE leads to its rupture [127, 128].
proposed that up to eight genes could be positioned due to the Microtubule‐targeting drugs have been shown to interfere with
eightfold symmetry of the structure [106]. Gene attachment is the onset of this effect and thereby delay NEBD [127].
suggested to be mediated by a DNA‐binding subunit that is an The combination of phosphorylation and microtubule‐based
identical feature of all NPCs [106]. tearing are believed to lead to the resulting NEBD [122]. Several
NPCs have been implicated in the sheltering of euchromatin kinases lead to the hyperphosphorylation of Nup98 and Nup35
regions from repressing factors [107–110]. Inside the nucleus, that may be contributing factors to NPC disassembly [56, 62,
the nuclear basket interacts with proteins involved in trafficking 129]. In addition, nuclear basket protein TPR may contribute to
of tRNA and mRNA [111, 112]. One example is the nuclear NPC disassembly and localization through its phosphorylation
basket protein Nup2p, which has been shown to interact with by mitotic kinases [45]. Although a general description of the
the histone variant H2AZ part of the euchromatin boundary for- dynamic process of NEBD is starting to emerge, very little is
mation and may position euchromatin in closer proximity to the known about the molecular machinery behind it. Multiple sign-
NPC [112–114]. Furthermore, studies combining peripheral aling pathways may be involved in the final triggering. For
positioning data with quantitative polymerase chain reaction example, it has been shown that high levels of RanGTP affect
(PCR) of mRNA levels correlate with expression of some genes the dynamics of the late steps of NEBD and may play a pivotal
coupled to NPCs [115–117]. Studies have also found that role in mitotic entry [120]. The full extent of molecular compo-
DNA and Nups immunoprecipitated with NPC components in nents required is not fully understood.
102 THE LIVER:  NUCLEAR PORE COMPLEX AND DISEASES

NUCLEAR PORE COMPLEX hundreds of rearranged DNA fragments, is highly correlated


AND DISEASES with NE rupture in micronuclei and chromatin bridges [141,
144, 145]. Such rearrangements may lead to circularization of
Nups have emerged as factors associated with human diseases. DNA, and these circular fragments may become amplified if
It has been well documented that chromosomal translocations they carry oncogenes [141, 146].
involving genes encoding Nups have been associated with many
forms of leukemia, including acute myelogenous leukemia Micronuclei and cancer
(AML), chronic myeloid leukemia (CML), myelodysplastic Micronuclei are small nuclei that form separately from the
syndrome (MDS), and T‐cell acute lymphoblastic leukemia ­primary nucleus following the formation of an NE around mis‐
(ALL) [7, 9, 10, 130–133]. segregated chromosomes [147]. Some micronuclei remain
intact upon formation, but there is also evidence that a substan-
tial proportion of micronuclei experience NE rupture, which is
Cancer less likely to be repaired than NE rupture that occurs in primary
nuclei. This subsequently may lead to improper functioning of
The involvement of the NPCs in mitosis, particularly in assem-
the micronuclei as well as extensive DNA damage, which has
bly and function of structures including kinetochores, mitotic
been observed in both cultured cancer cells as well as sections
spindles, and centrosomes as well as the role they play in proper
of non‐small cell lung cancer (NSCLS) tumors [147]. Disrupted
chromosome segregation, indicates their importance in main-
micronuclei have also been observed in non‐transformed cell
taining genome integrity. Thus, changes that affect the ability of
lines, which indicates that micronuclei could potentially be use-
the NPC to function in mitosis could contribute to cancer devel-
ful for detection in early cancer development [147–149]. There
opment [134]. It has been shown that Nup358 (also known as
is also evidence to show that defects are present within intact
RanBP2), a component of the cytoplasmic filaments of the NPC
micronuclei as well, including greater susceptibility to DNA
which functions in multiple cellular processes, including mito-
damage [147–149]. Disruption of micronuclei is correlated with
sis, is implicated in the development of colon cancer, the third
the development of some autoinflammatory diseases due to the
most common cancer worldwide. Specifically, Nup358 pro-
activation of cGAS, a DNA sensor, which is proposed to accu-
motes survival of colon cancer cells by contributing to the pre-
mulate on chromatin‐containing micronuclei following NE rup-
vention of mitotic cell death [134, 135]. A mutation that
ture [141, 150–152].
generally correlates with a less favorable prognosis and is
Micronuclear disruption, specifically loss of compartmentali-
observed in about 8–10% of patients with colon cancer is the
zation, does not appear to be caused by loss of the NPC [147]. It
BRAF (V600) mutation [134–136]. It has been shown that
has been shown that levels of core Nups detected by mAb414
Nup358 ameliorates the effects of mitotic defects present within
did not greatly differ between intact and disrupted micronuclei.
BRAF‐like colon cancer cell lines. Furthermore, it has been
However, there is evidence that Nup153 and TPR levels (both
shown that knockdown of RANBP2 leads to mitotic defects in
basket Nups) are lower in disrupted micronuclei, although this
colon cancer cells with this mutation, eventually resulting in cell
evidence is not conclusive regarding the connection between
death specifically due to prolonged mitosis [134, 135, 137].
loss of compartmentalization in micronuclei and stability of the
An important feature of nucleocytoplasmic transport through
NPC [147].
the NPC is the recognition of NLSs and NESs on large proteins
by transport receptors [9]. Proteins that contain NLSs and NESs
include oncogenes and tumor suppressors that have nuclear
Nup98
functions, such as p53 and FoxO [9]. Disruption of nucleocyto-
plasmic transport involving these proteins is associated with Nup98 regulates transcription of genes that have functions relat-
tumor formation [9]. These proteins bind with Crm1, an expor- ing to development and the cell cycle [9, 54, 55]. Chromosomal
tin that has been shown to exhibit overexpression in leukemias translocations involving Nup98 also alter expression of Nup96,
as well as gliomas and osteosarcomas [9]. It is believed that this as both proteins are encoded by the same mRNA [9, 27].
overexpression promotes excessive export of tumor suppressors Disrupted expression of Nup96 may serve as an additional con-
out of the nucleus, thus decreasing their function [9, 138]. In tributing factor to disease phenotypes that arise from Nup98
addition, a mutation in BRCA2, a tumor suppressor whose translocations with transcription factors, as Nup96 plays a role
mutant form is associated with the development of breast, ovar- in the regulation of export of mRNAs that are associated with
ian, and pancreatic cancers, has been connected to disruption of immunity and cell cycle regulation [9, 49, 153]. Nup98 translo-
nucleocytoplasmic transport. Interaction of a mutant BRCA2 cations are most frequently observed in AML, chronic myeloid
and the 26S proteasome complex subunit DSS1 has been shown leukemia in blast crisis (CML‐bc), and myelodysplastic syn-
to mask the NES of BRCA2 and allow recognition by Crm1, drome (MDS) [154]. Several mechanisms have been proposed
which mislocalizes it to the cytoplasm [9, 139]. to explain the role of Nup98 fusion proteins in leukemic trans-
Several types of cancer are associated with intricate mecha- formations, including upregulation of HOXA genes, suppressed
nisms known as chromothripsis and kataegis that occur in a sin- differentiation, and increased self‐renewal [154, 155]. A study
gle event and catastrophically alter the genome [140–143]. focused on Nup98–HOXA9 revealed a biphasic effect of the
These mechanisms may be connected to DNA damage that fusion protein on the growth of CD34+ hematopoietic cells, with
occurs as a result of NE rupture. For example, chromothripsis, a growth initially inhibited before continuous, long‐term prolif-
cataclysmic incidence in the genome that leads to upwards of eration of primitive cells was observed [156]. This finding
9:  Nuclear Pore Complex 103

suggests the development of AML from MDS, which has been aging brains of rats revealed that scaffold Nups have very slow
shown in transgenic mouse models containing the fusion protein turnover rates compared to peripheral Nups, and additionally
Nup98–HOXD13. In addition, Nup98–HOXA9 has been shown showed that relative Nup composition of NPCs changed
to suppress hematopoietic differentiation and to increase primi- throughout the aging process [155, 168–170]. This finding sug-
tive self‐renewing cells [156]. gests that slow Nup turnover rate may contribute to the accumu-
Nup98 has been implicated in infection by influenza virus. lation of damaged NPCs over time [155, 168].
Specifically, downregulation of Nup98 by the virus nonstruc- A very rare and fatal premature aging disease called
tural protein 1 (NS1) is associated with increased viral replica- Hutchinson–Gilford progeria syndrome has been associated
tion [9, 157]. This is correlated with evidence that viruses target with the malformation of the protein lamin A [171–173].
nuclear mRNA export in order to enhance replication, which Specifically, single‐base mutations of the LMNA gene lead to
includes the export of mRNAs that encode antiviral factors. For activation of a cryptic splice site and subsequent farnesylation
example, interaction occurs between Nup98 and the mRNA of lamin A, producing a variant of the protein known as progerin
export factor Rae1, which is targeted by the VSV matrix protein. [172, 173]. Continuous farnesylation allows insertion of prog-
This inhibition of mRNA export has been shown to be reversible erin into the inner nuclear membrane, in which it accumulates,
through upregulation of Rae1 and Nup96–Nup98, which may inflicting damage upon aging cells [173, 174].
be achieved by treatment with interferons [157]. It has been
demonstrated that upon infection with the influenza virus in
293T and MDCK cells, Nup98 is actively degraded. This poten-
Huntington’s disease
tially acts as a contributing factor to the suppression of nuclear HD, which is the most common heritable neurodegenerative
mRNA export [157]. It has also been shown that Nup98 is tar- disorder, has been associated with mislocalization and aggrega-
geted for degradation in cells infected with poliovirus, which is tion of several Nups, as well as disruption of nucleocytoplasmic
likely facilitated by a viral 2A protease. Poliovirus additionally transport [46]. HD is a member of a group of neurodegenerative
targets two other Nups, Nup153 and Nup62, but cleavage of disorders known as polyQ diseases, which are all characterized
Nup98 appears to occur more rapidly [158]. by repetitive CAG sequences that encode a long polyglutamine
(polyQ) tract within corresponding proteins [46, 175]. Nup
aggregates have been shown to colocalize with the mutant form
Triple A syndrome of the huntingtin protein, Htt. In addition, both the number and
Triple A or Allgrove syndrome, an adrenal insufficiency (acha- size of these aggregates increase with pathological progression
lasia–addisonianism–alacrima), is a rare autosomal recessive of HD [46].
congenital disorder with three common features of achalasia, Disruption of nucleocytoplasmic transport as a contributing
Addison disease, and alacrima [159, 160]. In most cases, there factor of HD has been evidenced by disturbance of the Ran gra-
is no family history of inception. Triple A is caused by muta- dient, which provides power for active transport and maintains
tions in the gene AAAS, encoding the Nup designated as proper transport directionality through the interaction of the
ALADIN [7, 161, 162]. The ALADIN nucleoporin is only protein Ran‐GTP with the transport receptor during nuclear
found within higher eukaryotes. Loss of ALADIN integration import, so cargo can be released. This gradient is maintained by
into the NPC results in the disease state [163]. DNA repair pro- a GTPase‐activating protein located on the cytoplasmic fila-
tein transport may be diminished if ALADIN is absent. Lack of ments of the NPC, RanGAP1 [46, 176]. Interactions between
DNA repair leads to instability and eventual cell death, espe- RanGAP1 and RNAs that contain a hexanucleotide repeat
cially in nerve cells [164]. POM protein NDC1 helps to target expansion (HRE) are linked to disruption of the Ran gradient.
ALADIN to the NPC [165–167]. Mutations in ALADIN affect These HRE mutations are correlated with some forms of HD
interactions and targeting, leading to changes in selective and are also commonly seen in amyotrophic lateral sclerosis
nuclear protein import. Depletion of NDC1 results in mislocali- (ALS) and frontotemporal dementia (FTD) [177–180]. In addi-
zation of ALADIN and may be involved in the pathogenesis of tion, RanGAP1 and Nup88 have been shown to form aggregates
the disease [165–167]. in a mouse model containing HD, and Nup62 has demonstrated
severe mislocalization in HD striatum tissue [46].
Premature and accelerated aging
Function of the NPC has also been shown to have a role in aging
Primary biliary cirrhosis
and age‐related deterioration. Deterioration occurs by means of Primary biliary cirrhosis (PBC) is an autoimmune disease of
molecular damage that builds over time, and Nups have among the liver [8, 10]. Production of autoantibodies signals a loss of
the greatest longevity of proteins within the mammalian brain tolerance and tissue damage [10]. In PBC, the gradual destruc-
[46, 168, 169]. Nup damage is associated with increased nuclear tion of the bile ducts leads to the onset of cirrhosis of the liver.
permeability, which poses the risk of infiltration by toxins and Patients with PBC generate a panel of autoantibodies [10, 181,
cytoplasmic proteins [46, 170]. In some studies, leakage of the 182] against mitochondrial antigens (AMAs), and nuclear pro-
cytoplasmic protein MAP2 has been observed within in vitro teins called antinuclear antibodies (ANAs) [10]. These ANAs
models of Huntington’s disease (HD). Since aging is a signifi- are directed against the NPC [183, 184] and generate a nuclear
cant risk factor for neurodegeneration, this also demonstrates rim staining by immunofluorescence. ANAs against Pom210
the role of the NPC in the development of neurodegenerative and Nup62 appear to be highly specific for PBC [10].
disorders [46]. A molecular study that observed changes in Combined with AMA detection, α‐gp210 and α‐Nup62
104 THE LIVER:  REFERENCES

antibodies may provide additional diagnostic and prognostic 15. Schwartz, T.U. Modularity within the architecture of the nuclear pore com-
tools [182, 185, 186]. Anti‐Pom210 ANAs in PBC specifically plex. Curr Opin Struct Biol, 2005;15(2):221–6.
16. Lim, R.Y., Aebi, U., and Fahrenkrog, B. Towards reconciling structure and
recognize a 15 amino acid domain within the C‐terminus function in the nuclear pore complex. Histochem Cell Biol, 2008;
[187]. It has been observed that increased ANA against 129(2):105–16.
Pom210 is associated with more severe disease and a worse 17. Kabachinski, G. and Schwartz, T.U. The nuclear pore complex – structure
prognosis [186], including greater vulnerability for progres- and function at a glance. J Cell Sci, 2015;128(3):423–9.
18. Hetzer, M.W., Walther, T.C., and Mattaj, I.W. Pushing the envelope: struc-
sion to end‐stage liver disease [181, 188–190]. The persistence
ture, function, and dynamics of the nuclear periphery. Annu Rev Cell Dev
of expression of AMAs and ANAs after liver transplants have Biol, 2005;21:347–80.
been described [10, 181]. In some patients, increased ANA 19. Belgareh, N., Rabut, G., Bai, S.W. et al. An evolutionarily conserved NPC
Pom210 levels persist years after transplantation. The question subcomplex, which redistributes in part to kinetochores in mammalian cells.
remains as to the clinical relevance of anti‐NPC antibodies in J Cell Biol, 2001;154(6):1147–60.
20. Walther, T.C., Alves, A., Pickersgill, H. et al. The conserved Nup107–160
PBC and the likelihood that an autoimmune response may
complex is critical for nuclear pore complex assembly. Cell, 2003;
ascend as a product of molecular mimicry [10, 181]. 113(2):195–206.
Understanding the organization of the membranous structure 21. Glavy, J.S., Krutchinsky, A.N., Cristea, I.M. et  al. Cell‐cycle‐dependent
of the NE and NPC is critical for defining the specificity of phosphorylation of the nuclear pore Nup107–160 subcomplex. Proc Natl
ANAs in PBC. But it is not yet known how mechanistically Acad Sci U S A, 2007;104(10):3811–16.
22. Aaronson, R.P. and Blobel G. On the attachment of the nuclear pore com-
these Nups are involved in the disease state [7, 10, 185]. plex. J Cell Biol, 1974;62(3):746–54.
23. Beck, M., Lucic, V., Forster, F., Baumeister, W., and Medalia O. Snapshots
of nuclear pore complexes in action captured by cryo‐electron tomography.
Nature, 2007;449(7162):611–15.
OUTLOOK 24. Beck, M., Forster, F., Ecke, M. et  al. Nuclear pore complex structure and
dynamics revealed by cryoelectron tomography. Science, 2004;306(5700):
1387–90.
With advances in identifying the structural connections of the 25. Alber, F., Dokudovskaya, S., Veenhoff, L.M. et al. The molecular architecture
NPC, we have gained valuable insights into the transporter. As of the nuclear pore complex. Nature, 2007;450(7170):695–701.
more puzzle pieces are fitted into place, we expect to link the 26. Alber, F., Dokudovskaya, S., Veenhoff, L.M. et al. Determining the architectures
NPC with its evolution and significance in abnormal states of of macromolecular assemblies. Nature, 2007;450(7170):683–94.
nuclear function. 27. Fontoura, B.M., Blobel, G., and Matunis, M.J. A conserved biogenesis path-
way for nucleoporins: proteolytic processing of a 186‐kilodalton precursor
generates Nup98 and the novel nucleoporin, Nup96. J Cell Biol,
1999;144(6):1097–112.
28. Loiodice, I., Alves, A., Rabut, G. et  al. The entire Nup107–160 complex,
REFERENCES including three new members, is targeted as one entity to kinetochores in
mitosis. Mol Biol Cell, 2004;15(7):3333–44.
1. Tran, E.J. and Wente, S.R. Dynamic nuclear pore complexes: life on the 29. Cristea, I.M., Williams, R., Chait, B.T., and Rout, M.P. Fluorescent proteins
edge. Cell, 2006;125(6):1041–53. as proteomic probes. Mol Cell Proteomics, 2005;4(12):1933–41.
2. Lusk, C.P., Blobel, G., and King, M.C. Highway to the inner nuclear mem- 30. Ori, A., Banterle, N., Iskar, M. et al. Cell type‐specific nuclear pores: a case
brane: rules for the road. Nat Rev Mol Cell Biol, 2007;8(5):414–20. in point for context‐dependent stoichiometry of molecular machines. Mol
3. Weis, K. Regulating access to the genome: nucleocytoplasmic transport Syst Biol, 2013;9:648.
throughout the cell cycle. Cell, 2003;112(4):441–51. 31. Boehmer, T., Enninga, J., Dales, S., Blobel, G., and Zhong, H. Depletion of a
4. Rout, M.P., Aitchison, J.D., Suprapto, A. et al. The yeast nuclear pore com- single nucleoporin, Nup107, prevents the assembly of a subset of nucleoporins
plex: composition, architecture, and transport mechanism. J Cell Biol, into the nuclear pore complex. Proc Natl Acad Sci U S A, 2003;100(3):981–5.
2000;148(4):635–51. 32. Boehmer, T., Jeudy, S., Berke, I.C., and Schwartz, T.U. Structural and func-
5. Cronshaw, J.M., Krutchinsky, A.N., Zhang, W., Chait, B.T., and Matunis, tional studies of Nup107/Nup133 interaction and its implications for the
M.J. Proteomic analysis of the mammalian nuclear pore complex. J Cell architecture of the nuclear pore complex. Mol Cell, 2008;30(6):721–31.
Biol, 2002;158(5):915–27. 33. Devos, D., Dokudovskaya, S., Alber, F. et al. Components of coated vesicles
6. DeGrasse, J.A., DuBois, K.N., Devos, D. et al. Evidence for a shared nuclear and nuclear pore complexes share a common molecular architecture. PLoS
pore complex architecture that is conserved from the last common eukary- Biol, 2004;2(12):e380.
otic ancestor. Mol Cell Proteomics, 2009;8(9):2119–30. 34. Drin, G., Casella, J.F., Gautier, R. et al. A general amphipathic alpha‐helical
7. Cronshaw, J.M. and Matunis, M.J. The nuclear pore complex: disease associa- motif for sensing membrane curvature. Nat Struct Mol Biol, 2007;14(2):
tions and functional correlations. Trends Endocrinol Metab, 2004;15(1):34–9. 138–46.
8. Giorgini, A., Selmi, C., Invernizzi, P. et al. Primary biliary cirrhosis: solving 35. Lutzmann, M., Kunze, R., Buerer, A., Aebi, U., and Hurt, E. Modular self‐
the enigma. Ann N Y Acad Sci, 2005;1051:185–93. assembly of a Y‐shaped multiprotein complex from seven nucleoporins.
9. Mor, A., White, M.A., and Fontoura, B.M. Nuclear trafficking in health and EMBO J, 2002;21(3):387–97.
disease. Curr Opin Cell Biol, 2014;28:28–35. 36. Berke, I.C., Boehmer, T., Blobel, G., and Schwartz, T.U. Structural and func-
10. Duarte‐Rey, C., Bogdanos, D., Yang, C.Y. et al. Primary biliary cirrhosis and tional analysis of Nup133 domains reveals modular building blocks of the
the nuclear pore complex. Autoimmun Rev, 2012;11(12):898–902. nuclear pore complex. J Cell Biol, 2004;167(4):591–7.
11. Bui, K.H., von Appen, A., Diguilio, A.L. et al. Integrated structural analysis 37. Hsia, K.C., Stavropoulos, P., Blobel, G., and Hoelz A. Architecture of a coat
of the human nuclear pore complex scaffold. Cell, 2013;155(6):1233–43. for the nuclear pore membrane. Cell, 2007;131(7):1313–26.
12. Kosinski, J., Mosalaganti, S., von Appen, A. et al. Molecular architecture of 38. Brohawn, S.G., Leksa, N.C., Spear, E.D., Rajashankar, K.R., and Schwartz,
the inner ring scaffold of the human nuclear pore complex. Science, T.U. Structural evidence for common ancestry of the nuclear pore complex
2016;352(6283):363–5. and vesicle coats. Science, 2008;322(5906):1369–73.
13. von Appen, A., Kosinski, J., Sparks, L. et al. In situ structural analysis of the 39. Stuwe, T., Correia, A.R., Lin, D.H. et al. Nuclear pores. Architecture of the
human nuclear pore complex. Nature, 2015;526(7571):140–3. nuclear pore complex coat. Science, 2015;347(6226):1148–52.
14. Hoelz, A., Glavy, J.S., and Beck, M. Toward the atomic structure of the 40. Hawryluk‐Gara, L.A., Shibuya, E.K., and Wozniak, R.W. Vertebrate Nup53
nuclear pore complex: when top down meets bottom up. Nat Struct Mol Biol, interacts with the nuclear lamina and is required for the assembly of a
2016;23(7):624–30. Nup93‐containing complex. Mol Biol Cell, 2005;16(5):2382–94.
9:  Nuclear Pore Complex 105

41. Jeudy, S. and Schwartz, T.U. Crystal structure of nucleoporin Nic96 reveals 66. Chen, J., Smoyer, C.J., Slaughter, B.D., Unruh, J.R., and Jaspersen, S.L. The
a novel, intricate helical domain architecture. J Biol Chem, 2007; SUN protein Mps3 controls Ndc1 distribution and function on the nuclear
282(48):34904–12. membrane. J Cell Biol, 2014;204(4):523–39.
42. Antonin, W., Ellenberg, J., and Dultz E. Nuclear pore complex assembly 67. Mitchell, J.M., Mansfeld, J., Capitanio, J., Kutay, U., and Wozniak, R.W.
through the cell cycle: regulation and membrane organization. FEBS Lett, Pom121 links two essential subcomplexes of the nuclear pore complex core
2008;582(14):2004–16. to the membrane. J Cell Biol, 2010;191(3):505–21.
43. Melcak, I., Hoelz, A., and Blobel G. Structure of Nup58/45 suggests flexible 68. Onischenko, E., Stanton, L.H., Madrid, A.S., Kieselbach, T., and Weis K.
nuclear pore diameter by intermolecular sliding. Science, 2007;315(5819): Role of the Ndc1 interaction network in yeast nuclear pore complex assem-
1729–32. bly and maintenance. J Cell Biol, 2009;185(3):475–91.
44. Rajanala, K. and Nandicoori, V.K. Localization of nucleoporin Tpr to the 69. Stavru, F., Hulsmann, B.B., Spang, A. et al. NDC1: a crucial membrane‐inte-
nuclear pore complex is essential for Tpr mediated regulation of the export gral nucleoporin of metazoan nuclear pore complexes. J Cell Biol,
of unspliced RNA. PLoS One, 2012;7(1):e29921. 2006;173(4):509–19.
45. Rajanala, K., Sarkar, A., Jhingan, G.D. et al. Phosphorylation of nucleoporin 70. Franz, C., Askjaer, P., Antonin, W. et al. Nup155 regulates nuclear envelope
Tpr governs its differential localization and is required for its mitotic func- and nuclear pore complex formation in nematodes and vertebrates. EMBO J,
tion. J Cell Sci, 2014;127(Pt 16):3505–20. 2005;24(20):3519–31.
46. Grima, J.C., Daigle, J.G., Arbez N et  al. Mutant Huntingtin disrupts the 71. Doucet, C.M. and Hetzer, M.W. Nuclear pore biogenesis into an intact
nuclear pore complex. Neuron, 2017;94(1):93–107 e6. nuclear envelope. Chromosoma, 2010;119(5):469–77.
47. Napetschnig, J., Blobel, G., and Hoelz A. Crystal structure of the N‐terminal 72. Shaulov, L., Gruber, R., Cohen, I., and Harel A. A dominant‐negative form of
domain of the human protooncogene Nup214/CAN. Proc Natl Acad Sci U S POM121 binds chromatin and disrupts the two separate modes of nuclear
A, 2007;104(6):1783–8. pore assembly. J Cell Sci, 2011;124(Pt 22):3822–34.
48. Tamura, K., Fukao, Y., Iwamoto, M., Haraguchi, T., and Hara‐Nishimura I. 73. Funakoshi, T., Clever, M., Watanabe, A., and Imamoto, N. Localization of
Identification and characterization of nuclear pore complex components in Pom121 to the inner nuclear membrane is required for an early step of inter-
Arabidopsis thaliana. Plant Cell, 2010;22(12):4084–97. phase nuclear pore complex assembly. Mol Biol Cell, 2011;22(7):1058–69.
49. Chakraborty, P., Wang, Y., Wei, J.H. et al. Nucleoporin levels regulate cell 74. Funakoshi, T., Maeshima, K., Yahata, K. et al. Two distinct human POM121
cycle progression and phase‐specific gene expression. Dev Cell, 2008; genes: requirement for the formation of nuclear pore complexes. FEBS Lett,
15(5):657–67. 2007;581(25):4910–16.
50. Chakraborty, P., Seemann, J., Mishra, R.K. et al. Vesicular stomatitis virus 75. Chadrin, A., Hess, B., San Roman, M. et al. Pom33, a novel transmembrane
inhibits mitotic progression and triggers cell death. EMBO Rep, 2009;10(10): nucleoporin required for proper nuclear pore complex distribution. J Cell
1154–60. Biol, 2010;189(5):795–811.
51. Liang, Y., Franks, T.M., Marchetto, M.C., Gage, F.H., and Hetzer, M.W. 76. Floch, A.G., Tareste, D., Fuchs, P.F. et al. Nuclear pore targeting of the yeast
Dynamic association of NUP98 with the human genome. PLoS Genet, Pom33 nucleoporin depends on karyopherin and lipid binding. J Cell Sci,
2013;9(2):e1003308. 2015;128(2):305–16.
52. Franks, T.M., McCloskey, A., Shokirev, M.N. et al. Nup98 recruits the Wdr82‐ 77. Galy, V., Antonin, W., Jaedicke, A. et al. A role for gp210 in mitotic nuclear‐
Set1A/COMPASS complex to promoters to regulate H3K4 trimethylation in envelope breakdown. J Cell Sci, 2008;121(Pt 3):317–28.
hematopoietic progenitor cells. Genes Dev, 2017;31(22):2222–34. 78. Dittmer, T.A. and Misteli T. The lamin protein family. Genome Biol,
53. Rosenblum, J.S. and Blobel G. Autoproteolysis in nucleoporin biogenesis. 2011;12(5):222.
Proc Natl Acad Sci U S A, 1999;96(20):11370–5. 79. Graziano, S., Kreienkamp, R., Coll‐Bonfill, N., and Gonzalo S. Causes and
54. Kalverda, B., Pickersgill, H., Shloma, V.V., and Fornerod, M. Nucleoporins consequences of genomic instability in laminopathies: replication stress and
directly stimulate expression of developmental and cell‐cycle genes inside interferon response. Nucleus, 2018;9(1):258–75.
the nucleoplasm. Cell, 2010;140(3):360–71. 80. Stewart, C.L. and Roux, K.J., Burke B. Blurring the boundary: the nuclear
55. Capelson, M., Liang, Y., Schulte, R. et  al. Chromatin‐bound nuclear pore envelope extends its reach. Science, 2007;318(5855):1408–12.
components regulate gene expression in higher eukaryotes. Cell, 2010;140(3): 81. Broers, J.L., Peeters, E.A., Kuijpers, H.J. et al. Decreased mechanical stiff-
372–83. ness in LMNA‐/‐ cells is caused by defective nucleo‐cytoskeletal integrity:
56. Laurell, E., Beck, K., Krupina, K. et al. Phosphorylation of Nup98 by multi- implications for the development of laminopathies. Hum Mol Genet,
ple kinases is crucial for NPC disassembly during mitotic entry. Cell, 2004;13(21):2567–80.
2011;144(4):539–50. 82. Crisp, M., Liu, Q., Roux, K. et al. Coupling of the nucleus and cytoplasm:
57. Laurell, E. and Kutay U. Dismantling the NPC permeability barrier at the role of the LINC complex. J Cell Biol, 2006;172(1):41–53.
onset of mitosis. Cell Cycle, 2011;10(14). 83. Manju, K., Muralikrishna, B., and Parnaik, V.K. Expression of disease‐caus-
58. Fontoura, B.M., Blobel, G., and Yaseen, N.R. The nucleoporin Nup98 is a ing lamin A mutants impairs the formation of DNA repair foci. J Cell Sci,
site for GDP/GTP exchange on ran and termination of karyopherin beta 2‐ 2006;119(Pt 13):2704–14.
mediated nuclear import. J Biol Chem, 2000;275(40):31289–96. 84. Hoelz, A. and Blobel G. Cell biology: popping out of the nucleus. Nature,
59. Enninga, J., Levy, D.E., Blobel, G., and Fontoura, B.M. Role of nucleoporin 2004;432(7019):815–16.
induction in releasing an mRNA nuclear export block. Science, 2002;295(5559): 85. Weis K. Nucleocytoplasmic transport: cargo trafficking across the border.
1523–5. Curr Opin Cell Biol, 2002;14(3):328–35.
60. Jeganathan, K.B., Malureanu, L., and van Deursen, J.M. The Rae1‐Nup98 86. Ribbeck, K. and Gorlich D. Kinetic analysis of translocation through nuclear
complex prevents aneuploidy by inhibiting securin degradation. Nature, pore complexes. EMBO J, 2001;20(6):1320–30.
2005;438(7070):1036–9. 87. Yang, W., Gelles, J., and Musser, S.M. Imaging of single‐molecule transloca-
61. Singer, S., Zhao, R., Barsotti, A.M. et al. Nuclear pore component Nup98 is tion through nuclear pore complexes. Proc Natl Acad Sci U S A, 2004;
a potential tumor suppressor and regulates posttranscriptional expression of 101(35):12887–92.
select p53 target genes. Mol Cell, 2012;48(5):799–810. 88. Hoelz, A., Debler, E.W., and Blobel G. The structure of the nuclear pore
62. Linder, M.I., Kohler, M., Boersema, P. et al. Mitotic disassembly of nuclear complex. Annu Rev Biochem, 2011;80:613–43.
pore complexes involves CDK1‐ and PLK1‐mediated phosphorylation of 89. Stewart, M. Molecular mechanism of the nuclear protein import cycle. Nat
key interconnecting nucleoporins. Dev Cell, 2017;43(2):141–56 e7. Rev Mol Cell Biol, 2007;8(3):195–208.
63. Kupke, T., Malsam, J., and Schiebel E. A ternary membrane protein complex 90. Lange, A., Mills, R.E., Lange, C.J. et al. Classical nuclear localization sig-
anchors the spindle pole body in the nuclear envelope in budding yeast. J nals: definition, function, and interaction with importin alpha. J Biol Chem,
Biol Chem, 2017;292(20):8447–58. 2007;282(8):5101–5.
64. Mansfeld, J., Guttinger, S., Hawryluk‐Gara, L.A. et al. The conserved trans- 91. Moore, M.S. and Blobel G. The GTP‐binding protein Ran/TC4 is
membrane nucleoporin NDC1 is required for nuclear pore complex assem- required for protein import into the nucleus. Nature, 1993;365(6447):
bly in vertebrate cells. Mol Cell, 2006;22(1):93–103. 661–3.
65. Antonin, W. and Mattaj, I.W. Nuclear pore complexes: round the bend? Nat 92. Gorlich, D. and Kutay U. Transport between the cell nucleus and the
Cell Biol, 2005;7(1):10–12. cytoplasm. Annu Rev Cell Dev Biol, 1999;15:607–60.
106 THE LIVER:  REFERENCES

93. King, M.C., Lusk, C.P., and Blobel G. Karyopherin‐mediated import of 120. Kutay, U. and Hetzer, M.W. Reorganization of the nuclear envelope during
integral inner nuclear membrane proteins. Nature, 2006;442(7106):1003–7. open mitosis. Curr Opin Cell Biol, 2008;20(6):669–77.
94. Meinema, A.C., Poolman, B., and Veenhoff, L.M. The transport of integral 121. Dultz, E., Zanin, E., Wurzenberger, C. et al. Systematic kinetic analysis of
membrane proteins across the nuclear pore complex. Nucleus, 2012;3(4): 322–9. mitotic dis‐ and reassembly of the nuclear pore in living cells. J Cell Biol,
95. Ellenberg, J., Siggia, E.D., Moreira, J.E. et al. Nuclear membrane dynamics 2008;180(5):857–65.
and reassembly in living cells: targeting of an inner nuclear membrane 122. Champion, L., Linder, M.I., and Kutay, U. Cellular reorganization during
protein in interphase and mitosis. J Cell Biol, 1997;138(6):1193–206. mitotic entry. Trends Cell Biol, 2017;27(1):26–41.
96. Worman, H.J. and Courvalin, J.C. The inner nuclear membrane. J Membr 123. De Souza, C.P. and Osmani, S.A. Mitosis, not just open or closed. Eukaryot
Biol, 2000;177(1):1–11. Cell, 2007;6(9):1521–7.
97. Wu, W., Lin, F., and Worman, H.J. Intracellular trafficking of MAN1, an 124. Bodoor, K., Shaikh, S., Salina, D. et  al. Sequential recruitment of NPC
integral protein of the nuclear envelope inner membrane. J Cell Sci, proteins to the nuclear periphery at the end of mitosis. J Cell Sci,
2002;115(Pt 7):1361–71. 1999;112(Pt 13):2253–64.
98. Soullam, B. and Worman, H.J. Signals and structural features involved in 125. Onischenko, E.A., Gubanova, N.V., Kiseleva, E.V., and Hallberg, E. Cdk1
integral membrane protein targeting to the inner nuclear membrane. J Cell and okadaic acid‐sensitive phosphatases control assembly of nuclear pore
Biol, 1995;130(1):15–27. complexes in Drosophila embryos. Mol Biol Cell, 2005;16(11):5152–62.
99. Ohba, T., Schirmer, E.C., Nishimoto, T., and Gerace L. Energy‐ and tempera- 126. Salina, D., Bodoor, K., Eckley, D.M. et al. Cytoplasmic dynein as a facilita-
ture‐dependent transport of integral proteins to the inner nuclear membrane tor of nuclear envelope breakdown. Cell, 2002;108(1):97–107.
via the nuclear pore. J Cell Biol, 2004;167(6):1051–62. 127. Beaudouin, J., Gerlich, D., Daigle, N., Eils, R., and Ellenberg J. Nuclear
100. Yavuz, S., Santarella‐Mellwig, R., Koch, B. et  al. NLS‐mediated NPC envelope breakdown proceeds by microtubule‐induced tearing of the lam-
functions of the nucleoporin Pom121. FEBS Lett. 2010;584(15):3292–8. ina. Cell, 2002;108(1):83–96.
101. Tapley, E.C., Ly, N., and Starr, D.A. Multiple mechanisms actively target 128. Busson, S., Dujardin, D., Moreau, A., Dompierre, J., and De Mey, J.R.
the SUN protein UNC‐84 to the inner nuclear membrane. Mol Biol Cell, Dynein and dynactin are localized to astral microtubules and at cortical
2011;22(10):1739–52. sites in mitotic epithelial cells. Curr Biol, 1998;8(9):541–4.
102. Turgay, Y., Ungricht, R., Rothballer, A. et al. A classical NLS and the SUN 129. Blethrow, J.D., Glavy, J.S., Morgan, D.O., and Shokat, K.M. Covalent cap-
domain contribute to the targeting of SUN2 to the inner nuclear membrane. ture of kinase‐specific phosphopeptides reveals Cdk1‐cyclin B substrates.
EMBO J, 2010;29(14):2262–75. Proc Natl Acad Sci U S A, 2008;105(5):1442–7.
103. Antonin, W., Ungricht, R., and Kutay U. Traversing the NPC along the 130. Romana, S.P., Radford‐Weiss, I., Ben Abdelali, R. et al. NUP98 rearrange-
pore  membrane: targeting of membrane proteins to the INM. Nucleus, ments in hematopoietic malignancies: a study of the Groupe Francophone
2011;2(2):87–91. de Cytogenetique Hematologique. Leukemia, 2006;20(4):696–706.
104. Meinema, A.C., Laba, J.K., Hapsari, R.A. et  al. Long unfolded linkers 131. Reader, J.C., Meekins, J.S., Gojo, I., and Ning Y. A novel NUP98‐PHF23
facilitate membrane protein import through the nuclear pore complex. fusion resulting from a cryptic translocation t(11;17)(p15;p13) in acute
Science, 2011;333(6038):90–3. myeloid leukemia. Leukemia, 2007;21(4):842–4.
105. Liu, D., Wu, X., Summers, M.D., Lee, A., Ryan, K.J., and Braunagel, S.C. 132. Panagopoulos, I., Kerndrup, G., Carlsen, N. et al. Fusion of NUP98 and the
Truncated isoforms of Kap60 facilitate trafficking of Heh2 to the nuclear SET binding protein 1 (SETBP1) gene in a paediatric acute T cell lymphoblas-
envelope. Traffic, 2010;11(12):1506–18. tic leukaemia with t(11;18)(p15;q12). Br J Haematol, 2007;136(2):294–6.
106. Blobel G. Gene gating: a hypothesis. Proc Natl Acad Sci U S A, 133. Suzuki, A., Ito, Y., Sashida, G. et al. t(7;11)(p15;p15) Chronic myeloid leu-
1985;82(24):8527–9. kaemia developed into blastic transformation showing a novel NUP98/
107. Schneider, R. and Grosschedl, R. Dynamics and interplay of nuclear architecture, HOXA11 fusion. Br J Haematol, 2002;116(1):170–2.
genome organization, and gene expression. Genes Dev, 2007;21(23):3027–43. 134. Wong, R.W. and D’Angelo M. Linking nucleoporins, mitosis, and colon
108. Taddei A. Active genes at the nuclear pore complex. Curr Opin Cell Biol, cancer. Cell Chem Biol, 2016;23(5):537–9.
2007;19(3):305–10. 135. Vecchione, L., Gambino, V., Raaijmakers, J. et  al. A vulnerability of a
109. Lanctot, C., Cheutin, T., Cremer, M., Cavalli, G., and Cremer, T. Dynamic ­subset of colon cancers with potential clinical utility. Cell, 2016;165(2):
genome architecture in the nuclear space: regulation of gene expression in 317–30.
three dimensions. Nat Rev Genet, 2007;8(2):104–15. 136. Popovici, V., Budinska, E., Tejpar, S. et al. Identification of a poor‐progno-
110. Menon, B.B., Sarma, N.J., Pasula, S. et al. Reverse recruitment: the Nup84 sis BRAF‐mutant‐like population of patients with colon cancer. J Clin
nuclear pore subcomplex mediates Rap1/Gcr1/Gcr2 transcriptional activa- Oncol, 2012;30(12):1288–95.
tion. Proc Natl Acad Sci U S A, 2005;102(16):5749–54. 137. Hashizume, C., Kobayashi, A., and Wong, R.W. Down‐modulation of
111. Dieppois, G., Iglesias, N., and Stutz F. Cotranscriptional recruitment to the nucleoporin RanBP2/Nup358 impaired chromosomal alignment and
mRNA export receptor Mex67p contributes to nuclear pore anchoring of induced mitotic catastrophe. Cell Death Dis, 2013;4:e854.
activated genes. Mol Cell Biol, 2006;26(21):7858–70. 138. Falini, B., Mecucci, C., Tiacci, E. et  al. Cytoplasmic nucleophosmin in
112. Ishii, K., Arib, G., Lin, C., Van Houwe, G., and Laemmli, U.K. Chromatin acute myelogenous leukemia with a normal karyotype. N Engl J Med,
boundaries in budding yeast: the nuclear pore connection. Cell, 2005;352(3):254–66.
2002;109(5):551–62. 139. Jeyasekharan, A.D., Liu, Y., Hattori, H. et al. A cancer‐associated BRCA2
113. Dilworth, D.J., Tackett, A.J., Rogers, R.S. et  al. The mobile nucleoporin mutation reveals masked nuclear export signals controlling localization.
Nup2p and chromatin‐bound Prp20p function in endogenous NPC‐medi- Nat Struct Mol Biol, 2013;20(10):1191–8.
ated transcriptional control. J Cell Biol, 2005;171(6):955–65. 140. Maciejowski, J. and Imielinski M. Modeling cancer rearrangement land-
114. Marshall, I.C. and Wilson, K.L. Nuclear envelope assembly after mitosis. scapes. Curr Opin Syst Biol, 2017;1:54–61.
Trends Cell Biol, 1997;7(2):69–74. 141. Hatch, E.M. Nuclear envelope rupture: little holes, big openings. Curr Opin
115. Casolari, J.M., Brown, C.R., Drubin, D.A., Rando, O.J., and Silver, P.A. Cell Biol, 2018;52:66–72.
Developmentally induced changes in transcriptional program alter spatial 142. Alexandrov, L.B., Nik‐Zainal, S., Wedge, D.C. et al. Signatures of muta-
organization across chromosomes. Genes Dev, 2005;19(10):1188–98. tional processes in human cancer. Nature, 2013;500(7463):415–21.
116. Taddei, A., Van Houwe, G., Hediger, F. et  al. Nuclear pore association 143. Roberts, S.A. and Gordenin, D.A. Hypermutation in human cancer
­confers optimal expression levels for an inducible yeast gene. Nature, genomes: footprints and mechanisms. Nat Rev Cancer, 2014;14(12):
2006;441(7094):774–8. 786–800.
117. Brickner, J.H. and Walter, P. Gene recruitment of the activated IN01 locus 144. Zhang, C.Z., Spektor, A., Cornils, H. et al. Chromothripsis from DNA dam-
to the nuclear membrane. PLoS Biol, 2004;2(11):e342. age in micronuclei. Nature, 2015;522(7555):179–84.
118. Casolari, J.M., Brown, C.R., Komili, S. et al. Genome‐wide localization of 145. Maciejowski, J., Li, Y., Bosco, N., Campbell, P.J., and de Lange, T.
the nuclear transport machinery couples transcriptional status and nuclear Chromothripsis and kataegis induced by telomere crisis. Cell,
organization. Cell, 2004;117(4):427–39. 2015;163(7):1641–54.
119. Casolari, J.M. and Silver, P.A. Guardian at the gate: preventing unspliced 146. Korbel, J.O. and Campbell, P.J. Criteria for inference of chromothripsis in
pre‐mRNA export. Trends Cell Biol, 2004;14(5):222–5. cancer genomes. Cell, 2013;152(6):1226–36.
9:  Nuclear Pore Complex 107

147. Hatch, E.M., Fischer, A.H., Deerinck, T.J., and Hetzer, M.W. Catastrophic 169. Toyama, B.H., Savas, J.N., Park, S.K. et al. Identification of long‐lived pro-
nuclear envelope collapse in cancer cell micronuclei. Cell, 2013;154(1): teins reveals exceptional stability of essential cellular structures. Cell,
47–60. 2013;154(5):971–82.
148. Crasta, K., Ganem, N.J., Dagher, R. et al. DNA breaks and chromosome 170. D’Angelo, M.A., Raices, M., Panowski, S.H., and Hetzer, M.W. Age‐
pulverization from errors in mitosis. Nature, 2012;482(7383):53–8. dependent deterioration of nuclear pore complexes causes a loss of nuclear
149. Terradas, M., Martin, M., Hernandez, L., Tusell, L., and Genesca A. integrity in postmitotic cells. Cell, 2009;136(2):284–95.
Nuclear envelope defects impede a proper response to micronuclear DNA 171. De Sandre‐Giovannoli, A., Bernard, R., Cau, P. et al. Lamin a truncation in
lesions. Mutat Res, 2012;729(1–2):35–40. Hutchinson‐Gilford progeria. Science, 2003;300(5628):2055.
150. Harding, S.M., Benci, J.L., Irianto, J. et al. Mitotic progression following 172. Eriksson, M., Brown, W.T., Gordon, L.B. et  al. Recurrent de novo point
DNA damage enables pattern recognition within micronuclei. Nature, mutations in lamin A cause Hutchinson‐Gilford progeria syndrome. Nature,
2017;548(7668):466–70. 2003;423(6937):293–8.
151. Mackenzie, K.J., Carroll, P., Martin, C.A. et  al. cGAS surveillance of 173. Gordon, L.B., Shappell, H., Massaro, J. et  al. Association of Lonafarnib
micronuclei links genome instability to innate immunity. Nature, treatment vs no treatment with mortality rate in patients with Hutchinson‐
2017;548(7668):461–5. Gilford progeria syndrome. JAMA, 2018;319(16):1687–95.
152. Dou, Z., Ghosh, K., Vizioli, M.G. et  al. Cytoplasmic chromatin triggers 174. Casasola, A., Scalzo, D., Nandakumar, V. et  al. Prelamin A processing,
inflammation in senescence and cancer. Nature, 2017;550(7676):402–6. accumulation and distribution in normal cells and laminopathy disorders.
153. Faria, A.M., Levay, A., Wang, Y. et al. The nucleoporin Nup96 is required Nucleus, 2016;7(1):84–102.
for proper expression of interferon‐regulated proteins and functions. 175. Finkbeiner, S. Huntington’s disease. Cold Spring Harb Perspect Biol,
Immunity, 2006;24(3):295–304. 2011;3(6).
154. Gough, S.M., Slape, C.I., and Aplan, P.D. NUP98 gene fusions and hemat- 176. Floch, A.G., Palancade, B., and Doye V. Fifty years of nuclear pores and
opoietic malignancies: common themes and new biologic insights. Blood, nucleocytoplasmic transport studies: multiple tools revealing complex
2011;118(24):6247–57. rules. Methods Cell Biol, 2014;122:1–40.
155. Sakuma, S. and D’Angelo, M.A. The roles of the nuclear pore complex in cel- 177. Freibaum, B.D., Lu, Y., Lopez‐Gonzalez, R. et al. GGGGCC repeat expan-
lular dysfunction, aging and disease. Semin Cell Dev Biol, 2017;68 :72–84. sion in C90rf72 compromises nucleocytoplasmic transport. Nature,
156. Takeda, A., Goolsby, C., and Yaseen, N.R. NUP98‐HOXA9 induces long‐term 2015;525(7567):129–33.
proliferation and blocks differentiation of primary human CD34+ hemat- 178. Jovicic, A., Mertens, J., Boeynaems, S. et al. Modifiers of C90rf72 dipep-
opoietic cells. Cancer Res, 2006;66(13):6628–37. tide repeat toxicity connect nucleocytoplasmic transport defects to FTD/
157. Satterly, N., Tsai, P.L., van Deursen, J. et  al. Influenza virus targets the ALS. Nat Neurosci, 2015;18(9):1226–9.
mRNA export machinery and the nuclear pore complex. Proc Natl Acad Sci 179. Hensman Moss, D.J., Poulter, M., Beck, J. et al. C90rf72 expansions are the
U S A, 2007;104(6):1853–8. most common genetic cause of Huntington disease phenocopies. Neurology,
158. Park, N., Katikaneni, P., Skern, T., and Gustin, K.E. Differential targeting 2014;82(4):292–9.
of nuclear pore complex proteins in poliovirus‐infected cells. J Virol, 180. Zhang, K., Donnelly, C.J., Haeusler, A.R. et  al. The C90rf72 repeat
2008;82(4):1647–55. expansion disrupts nucleocytoplasmic transport. Nature, 2015;
159. Allgrove, J., Clayden, G.S., Grant, D.B., and Macaulay, J.C. Familial glu- 525(7567):56–61.
cocorticoid deficiency with achalasia of the cardia and deficient tear pro- 181. Nakamura, M., Shimizu‐Yoshida, Y., Takii, Y. et  al. Antibody titer to
duction. Lancet, 1978;1(8077):1284–6. gp210‐C terminal peptide as a clinical parameter for monitoring primary
160. Sandrini, F., Farmakidis, C., Kirschner, L.S. et al. Spectrum of mutations of biliary cirrhosis. J Hepatol, 2005;42(3):386–92.
the AAAS gene in Allgrove syndrome: lack of mutations in six kindreds 182. Invernizzi, P., Selmi, C., Ranftler, C., Podda, M., and Wesierska‐Gadek J.
with isolated resistance to corticotropin. J Clin Endocrinol Metab, Antinuclear antibodies in primary biliary cirrhosis. Semin Liver Dis,
2001;86(11):5433–7. 2005;25(3):298–310.
161. Krumbholz, M., Koehler, K., and Huebner A. Cellular localization of 17 natural 183. Worman, H.J. and Courvalin, J.C. Antinuclear antibodies specific for pri-
mutant variants of ALADIN protein in triple A syndrome – shedding light on mary biliary cirrhosis. Autoimmun Rev, 2003;2(4):211–17.
an unexpected splice mutation. Biochem Cell Biol, 2006;84(2):243–9. 184. Invernizzi, P., Podda, M., Battezzati, P.M. et  al. Autoantibodies against
162. Mukhopadhya, A., Danda, S., Huebner, A., and Chacko A. Mutations of the nuclear pore complexes are associated with more active and severe liver
AAAS gene in an Indian family with Allgrove’s syndrome. World J disease in primary biliary cirrhosis. J Hepatol, 2001;34(3):366–72.
Gastroenterol, 2006;12(29):4764–6. 185. Worman, H.J. Nuclear envelope protein autoantigens in primary biliary cir-
163. Dusek, T., Korsic, M., Koehler, K. et  al. A novel AAAS gene mutation rhosis. Hepatol Res, 2007;37 Suppl 3:S406–11.
(p.R194X) in a patient with triple A syndrome. Horm Res, 2006;65(4): 186. Sfakianaki, O., Koulentaki, M., Tzardi, M. et  al. Peri‐nuclear antibodies
171–6. correlate with survival in Greek primary biliary cirrhosis patients. World J
164. Zimmer, V., Liebe, R., and Lammert F. Nuclear receptor variants in liver Gastroenterol, 2010;16(39):4938–43.
disease. Dig Dis, 2015;33(3):415–19. 187. Nickowitz, R.E. and Worman, H.J. Autoantibodies from patients with pri-
165. Cho, A.R., Yang, K.J., Bae, Y. et al. Tissue‐specific expression and subcel- mary biliary cirrhosis recognize a restricted region within the cytoplasmic
lular localization of ALADIN, the absence of which causes human triple A tail of nuclear pore membrane glycoprotein Gp210. J Exp Med,
syndrome. Exp Mol Med, 2009;41(6):381–6. 1993;178(6):2237–42.
166. Kind, B., Koehler, K., Lorenz, M., and Huebner A. The nuclear pore complex 188. Itoh, S., Ichida, T., Yoshida, T. et al. Autoantibodies against a 210 kDa glyco-
protein ALADIN is anchored via NDC1 but not via POM121 and GP210 in protein of the nuclear pore complex as a prognostic marker in patients with
the nuclear envelope. Biochem Biophys Res Commun, 2009;390(2):205–10. primary biliary cirrhosis. J Gastroenterol Hepatol, 1998;13(3): 257–65.
167. Yamazumi, Y., Kamiya, A., Nishida, A. et al. The transmembrane nucleop- 189. Bogdanos, D.P., Liaskos, C., Pares, A. et al. Anti‐gp210 antibody mirrors
orin NDC1 is required for targeting of ALADIN to nuclear pore complexes. disease severity in primary biliary cirrhosis. Hepatology, 2007;45(6):1583;
Biochem Biophys Res Commun, 2009;389(1):100–4. author reply ‐4.
168. Savas, J.N., Toyama, B.H., Xu, T., Yates, J.R., 3rd, and Hetzer, M.W. 190. Wesierska‐Gadek, J., Penner, E., Battezzati, P.M. et al. Correlation of initial
Extremely long‐lived nuclear pore proteins in the rat brain. Science, autoantibody profile and clinical outcome in primary biliary cirrhosis.
2012;335(6071):942. Hepatology, 2006;43(5):1135–44.
Protein Maturation
10 and Processing at the
Endoplasmic Reticulum
Ramanujan S. Hegde
MRC Laboratory of Molecular Biology, Cambridge, UK

INTRODUCTION are completely translocated across the ER membrane into the


luminal space, while membrane proteins are woven into the ER
The liver, and hepatocytes isolated from it, have long played a cen- membrane, with specific regions exposed to the cytosol and other
tral role in our understanding of cellular organization, protein traf- portions exposed to the lumen. After achieving their correct
ficking, and secretion. The abundance, accessibility, high secretory topology (i.e. orientation relative to the membrane), these pro-
capacity, and wide range of functions made the liver an ideal choice teins undergo a series of maturation events in the ER to produce
for many of the earliest and seminal studies of cellular function. a functional protein. This involves folding, various modifications,
The first electron microscopic studies revealed a remarkable com- and assembly with other proteins or cofactors. Only upon suc-
partmentalization to hepatocytes, including the highly abundant cessful maturation are the secretory and membrane proteins
membrane‐bound organelles of the endoplasmic reticulum (ER), allowed to exit the ER for their final destinations via vesicular
Golgi, mitochondria, peroxisomes, and others. Upon the develop- trafficking pathways. Thus, the ER can be considered an intracel-
ment of subcellular fractionation methods, each of these organelles lular factory for the biosynthesis and maturation of secretory and
could be isolated and studied by classical biochemistry and membrane proteins. The key steps in the assembly line of this
­enzymology [1]. The isolation of liver and hepatocyte membrane factory are the selective targeting of secretory and membrane
fractions enriched in “rough” ER [2] was instrumental in the dis- proteins to the ER, their translocation across or into the ER mem-
covery that ER‐bound ribosomes synthesize secretory and mem- brane, their folding into the correct conformation, and assembly
brane proteins [3]. These proteins were found to be selectively into a functional product.
imported into the ER or inserted into the ER membrane. Indeed,
the ER eventually proved to be the major site of secretory and
membrane protein biosynthesis and maturation. Twenty years later,
Secretory and membrane proteins are
the ER was discovered to also be a major site of quality control, the recognized via signal sequences
process whereby immature, defective, or otherwise incompletely Cellular protein synthesis occurs on ribosomes located in the
assembled proteins are triaged for selective degradation [4]. This cytosol, but over half of all proteins eventually reside in non‐
chapter will cover the basic principles of secretory and membrane cytosolic locales within the highly compartmentalized eukary-
protein maturation and quality control in the ER. otic cell. Thus, it is obvious that cells must effectively sort and
segregate proteins among numerous intracellular organelles.
One of the major conceptual advances in understanding this
THE SEGREGATION OF SECRETORY process was the articulation of a cogent hypothesis for the basis
AND MEMBRANE PROTEINS TO THE ER of intracellular protein segregation. The “signal hypothesis”
postulated that specific regions of a newly synthesized polypep-
Essentially all secretory and membrane proteins destined for tide (i.e. signal sequences) provide unique codes that specify the
the cell surface, extracellular environment, Golgi, lysosomes, and eventual destination for that protein [5]. This concept, originally
endosomes begin their biosynthesis at the ER. Secretory proteins developed to explain how secretory proteins are selectively

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
10:  Protein Maturation and Processing at the Endoplasmic Reticulum 109

Protein Targeting sequence


Prolactin MNIKGSPWKGSLLLLLVSNLLLCQSVAP
TGF-β2 MHYCVLSAFLILHLVTVAL. ..
Growth hormone MATGSRTSLLLAFGLLCLPWLQEGSA...

N-terminal signals
Osteopontin MRIAVICFCLLGITCA...
Leptin MHWGTLCGFLWLWPYLFYVQA...

Cleaved
Apo-A1 MKAAVLTLAVLFLTGSQA...
EGF-receptor MRPSGTAGAALLALLAALCPRA...
Glucagon MKSIYFVAGLFVMLVQG...
lnhibinβ MPLLWLRGFLLASCWllVRSSPTPGS...
BiP MKLSLVAAMLLLLSAARA...
Choriogonadotropin MEMFQGLLLLLLLSMGGTWA...

Rhodopsin ...WQFSMLAAYMFLLIVLGFPINFLTLYVTVQH...
Transmembrane

CFTR ...FFWRFMFYGIFLYLGEVTKAVQPLLLGRllA...
domains

ASGR I ...KFRLSLLALAFNILLLVVICVVSSQSMQLQK...
Synaptotagmin II ...EINKIPLPPWALIAMAVVAGLLLLTCCFCIC...
Glycophorin C ...TIMDIVVIAGVIAAVAIVLVSLLFVMLRYMYR...
Mannosidase II ... RRRFALVICSGCLLVFLSLYllLNFAAPAATQ...

Figure 10.1  Examples of signal sequences used for protein targeting to the endoplasmic reticulum. In many cases, the signal sequence is located
at the N‐terminus and is cleaved by the enzyme signal peptidase shortly after targeting has been accomplished. In other cases, the first transmem-
brane domain is used for targeting. These are not removed, and are part of the final protein product. In all cases, the defining feature is a stretch
of ~10 or more hydrophobic residues (underlined sequence).

segregated to the ER, eventually proved to be widely generaliz- TMDs, and topology with respect to the ER membrane. This
able. The signals for targeting to the ER, nucleus, mitochondria, diversity means that no single mechanism can recognize, target,
peroxisomes, and other locations have been identified. In each translocate, and insert all of these proteins. Hence, cells
case, the sorting signals are selectively recognized by dedicated have evolved different pathways to deal with different types of
machinery that mediates correct targeting to the intended proteins.
destination. There are two general mechanisms of protein targeting to the
For secretory and membrane proteins, the distinguishing ER. In co‐translational targeting, proteins are recognized as
feature of their signal sequences essential for selective they are being translated (Figure 10.2a). Thus, the entire ribo-
­recognition is hydrophobicity (Figure  10.1). In many mem- some is targeted to the ER, where the nascent protein is trans-
brane proteins, the first transmembrane domain (TMD), which located across or inserted into the membrane during its
necessarily must be hydrophobic in order to eventually reside synthesis. This mechanism has the advantage that an unfolded
stably in the lipid bilayer, serves as the signal sequence for polypeptide can be moved across or inserted into the mem-
selective recognition. However, secretory proteins are soluble brane before it can fold in the cytosol. In post‐translational
in their mature form, and therefore do not typically contain targeting, proteins are synthesized completely in the cytosol
long stretches of hydrophobic residues in their final primary before their delivery to the ER membrane for translocation or
sequence. Instead, they are usually made as precursors contain- insertion. This mechanism generally operates on proteins that
ing a cleavable signal sequence at the N‐terminus. The cleava- do not have large domains that need to be translocated across
ble signal is typically ~15–40 amino acids long and contains a the membrane.
central hydrophobic region essential for its selective recogni- In eukaryotic cells, the majority of proteins use co‐transla-
tion and targeting to the ER. This signal is removed at the ER tional targeting [8]. However, some secretory proteins are very
by an enzyme called signal peptidase, and is therefore not part small (less than ~100 amino acids) and their synthesis occurs
of the final mature protein. Some membrane proteins also use a too quickly to be effectively recognized during translation.
cleavable N‐terminal signal for targeting to the ER. Thus, all These proteins, which include various hormones, cytokines, and
secretory and membrane proteins are recognized via a hydro- antibacterial peptides, are instead post‐translationally targeted
phobic motif that sets them apart from other non‐ER targeted and translocated into the ER. Similarly, some membrane pro-
proteins [6]. teins contain a single TMD located close to the C‐terminus [9].
This TMD serves as the signal for targeting such “tail‐anchored”
Multiple pathways of protein targeting membrane proteins to the ER. Because a C‐terminal TMD
will  not emerge from the ribosome until after the protein has
to the ER membrane been fully synthesized, these proteins are also targeted post‐­
Of the ~20 000 proteins encoded in the human genome, ~7000 translationally. The pathways for co‐translational and post‐
are targeted to the ER [7]. These proteins are highly diverse in translational targeting involve different factors, and will be
their biophysical characteristics, number and location of their considered in successive sections.
110 THE LIVER:  THE SEGREGATION OF SECRETORY AND MEMBRANE PROTEINS TO THE ER

(a)

Secretory
SRP- protein
receptor Translocon Membrane
protein

ER lumen

SRP Cytosol

Signal
sequence
Translocation Membrane
Targeting integration
Ribosome

Signal
recognition
mRNA
Initiation of
synthesis

(b) Translocation
Sec61
complex Signal Nascent
Plug
sequence protein ER membrane

Membrane
insertion

Lateral gate
Ribosome Ribosome

Inactive translocon Opened by a signal Active translocon

Figure 10.2  Segregation of secretory and membrane proteins to the endoplasmic reticulum (ER). (a) The key steps of co‐translational transloca-
tion including: signal sequence recognition by the signal recognition particle (SRP), targeting to the ER membrane via SRP receptor (SR), and
translocation through a translocon. Membrane proteins utilize the same basic steps and machinery, but undergo integration into the membrane rather
than complete translocation. (b) The Sec61 complex forms a protein‐conducting channel in the ER membrane. In the inactive state (left), a seam in
the channel (the lateral gate) is closed, and the pore is occluded by a plug. A signal sequence can open the Sec61 complex by intercalating into the
lateral gate (middle). In the active state (right), the channel can move polypeptides in two directions: across the membrane (translocation) or into
the lipid bilayer (membrane insertion).

Signal recognition particle mediates Selective ER targeting of SRP‐bound ribosomes is medi-


ated by the presence in the ER membrane of a specific recep-
co‐translational protein targeting tor for SRP. The interaction between SRP and SRP receptor
Nearly all N‐terminal signals and TMDs are recognized co‐­ (SR) not only mediates targeting, but also the transfer of the
translationally as they first emerge from the ribosome. A large, nascent polypeptide to a translocon formed by the Sec61
ribosome‐associated factor termed signal recognition particle complex (described in the next section). This unidirectional
(SRP) interacts directly with the hydrophobic signal sequence transfer of the nascent chain from SRP to the translocon is
[10]. Co‐translational SRP‐mediated recognition has three impor- energy dependent. Both SRP and SR are GTPases. SRP binds
tant consequences. First, SRP shields the hydrophobic TMD to a signal sequence in its GTP form, while SR binds to the
or signal from the bulk aqueous environment to prevent inappro- translocon in its GTP‐bound form. When the GTP‐bound
priate interactions or aggregation. In this sense, SRP can be forms of SRP and SR interact with each other, they hydrolyze
viewed as a chaperone of sorts that protects nascent secretory and their GTPs. This induces a series of conformational changes
membrane proteins during the earliest stages of their synthesis. that ultimately results in SRP dissociating from the signal
Second, the SRP–signal interaction modulates ribosome function sequence, SR, and the ribosome. The ribosome and the
to slow the rate of translation. This transient slowing provides ­liberated signal sequence now engage the translocon. In this
extra time for targeting to the ER membrane. And third, SRP is way,  nascent proteins containing a signal sequence or TMD
a  key distinguishing feature that allows some ribosomes to be are selectively delivered to translocons at the ER where they
targeted to the ER, while others remain in the cytosol. complete their synthesis.
10:  Protein Maturation and Processing at the Endoplasmic Reticulum 111

Sec61 complex forms a translocation channel by a signal sequence and a large soluble domain has been
across the ER ­translocated into the lumen, then the topology has already been
committed. When a TMD then emerges from the ribosome in
After targeting to the ER translocon, secretory proteins are this type of protein, it will insert with its N‐terminal‐flanking
translocated through a channel across the membrane, while domain in the lumen and C‐terminal‐flanking domain facing the
membrane proteins are inserted into the lipid bilayer (discussed cytosol.
in the next section). Both of these processes are mediated by a For proteins that are targeted via their first TMD, the determi-
translocation channel that can open perpendicularly across the nation of topology is more complex [17]. Here, the length of the
membrane for translocation and laterally toward the membrane polypeptide preceding the TMD, the charged residues flanking
for insertion [11, 12]. This translocation channel is formed by the TMD, the TMD length, and its overall hydrophobicity all
the three‐protein Sec61 complex. influence its final topology. In general, the N‐terminus is favored
The structure of the Sec61 complex [13, 14] shows it to be a to face the cytosol if it is long and contains basic residues adja-
toroidal channel with a central hourglass‐shaped pore and a cent to the TMD. Otherwise, the TMD will be oriented with the
seam along the side (Figure 10.2b, left). When the Sec61 com- N‐terminus facing the ER lumen.
plex is quiescent, the pore is plugged and the front of the seam, For proteins containing multiple TMDs, each TMD after the
known as the lateral gate, is closed. Thus, ions and other small first takes on the opposite orientation of the TMD that precedes
molecules cannot leak into or out of the ER. it. This means that for multipass membrane proteins, topology
After SRP targets a translating ribosome to the translocon, the of the first TMD essentially fixes the topology of the remaining
Sec61 complex binds tightly to a site that is located adjacent to ones [18]. Each TMD is thought to enter the membrane via the
the ribosomal tunnel through which the newly made protein lateral gate of Sec61 in succession. There may be accessory pro-
emerges. This positions the Sec61 complex such that its channel teins that facilitate TMD insertion, and there may be complex
is roughly co‐linear with the ribosomal tunnel. At this point, the membrane proteins whose insertion is not simply vectorial [16].
signal sequence of a secretory protein still needs to open These aspects of how a multipass membrane protein is inserted
the  Sec61 channel to initiate translocation. This occurs when remain to be investigated in detail.
the  signal sequence becomes embedded into the lateral gate
(Figure 10.2b, middle). This only occurs for bona fide signals, Post‐translational targeting of secretory
thereby providing a proofreading mechanism that prevents pro-
teins lacking a signal sequence from opening the channel. and membrane proteins
Embedding the signal into the lateral gate of Sec61 widens Proteins whose sole targeting signal is close to the C‐terminus
the channel [15], which leads to displacement of the plug that cannot be targeted co‐translationally [9]. The reason is that the
normally occupies the central pore. Thus, recognition of a signal ribosomal tunnel through which nascent proteins emerge
or TMD causes the Sec61 complex to open and makes it active houses ~40 amino acids. This means that at the very minimum,
for translocation and membrane insertion (Figure 10.2b, right). ~40 residues are needed beyond the targeting signal for it to be
In this mode, the Sec61 complex is able to translocate proteins recognized during synthesis. However, it probably takes
across the membrane through the central channel, or insert ~5–10 seconds to carry out the reactions involved in co‐
membrane proteins into the lipid bilayer via the lateral gate. translational targeting. At a translation speed of 6 residues per
In the case of secretory proteins, the segment of polypeptide second, this means that another ~30–60 residues are needed
downstream from the signal sequence enters the channel and after the targeting signal is exposed for co‐translational target-
accesses the lumen of the ER. Continued translation by the ing to occur successfully. Thus, proteins whose targeting sig-
bound ribosome expels the growing polypeptide into the lumen, nal is located within the last ~90–120 residues must be targeted
where molecular chaperones (discussed later) bind to the post‐translationally.
­nascent protein and ensure its translocation. Once translocation In the case of secretory proteins, the hydrophobic N‐terminal
has been initiated, the signal sequence is removed by an enzyme signal sequence must be shielded to prevent its aggregation,
called signal peptidase. Translocation continues until translation while the remainder of the polypeptide must be prevented from
is finished, at which point the last bit of the nascent protein folding. The abundant cytosolic protein calmodulin is thought to
enters the ER lumen, the ribosome dissociates from the Sec61 bind the signal sequence dynamically to prevent aggregation,
complex, and the channel again returns to its quiescent state. while still allowing opportunities for it to engage a translocon at
the ER [19]. Folding of the remainder of the protein is probably
Co‐translational membrane protein insertion prevented by chaperones, typically of the Hsp70 family, that
release the protein prior to its translocation.
by the Sec61 complex The translocon for post‐translational translocation contains
If the protein targeted to the Sec61 translocon contains TMDs, the Sec61 complex together with the accessory proteins Sec62
it needs to be inserted into the ER membrane [16]. The lateral and Sec63 [20]. This translocon can recognize the signal
gate of the Sec61 complex provides the key point of transition sequences of post‐translationally translocated secretory proteins
between the aqueous cytosol and hydrophobic lipid bilayer. As by a mechanism that is probably similar to that described above
a TMD emerges from the ribosome, it exploits this gate to enter for co‐translational proteins. This recognition leads to the initial
the membrane (Figure 10.2b, right). insertion of the small secretory protein into the Sec61 translo-
The orientation (or topology) in which a TMD is inserted is cation channel, exposing a portion of the polypeptide to the
influenced by several parameters. If a protein has been targeted lumen. The ER luminal chaperone BiP, recruited to the site of
112 THE LIVER:  FOLDING AND MATURATION OF SECRETORY AND MEMBRANE PROTEINS

translocation by its interaction with Sec63, probably interacts Although the mechanisms that coordinate the multistep matu-
with this exposed segment of polypeptide to favor its transloca- ration events of complex proteins remain poorly understood,
tion into the ER lumen [21]. many of the individual steps have been studied in detail and
Post‐translationally inserted membrane proteins use an are discussed here.
entirely different set of factors for targeting and insertion [22].
The targeting factor for such proteins is a cytosolic factor called Numerous co‐ and post‐translational
TRC40 (known as Get3 in yeast) that directly recognizes, binds,
modifications accompany protein maturation
and shields TMDs [23]. TRC40 has a specific receptor at the ER
membrane. This receptor dislodges the TMD from TRC40 and Most secretory and membrane proteins are covalently modified
facilitates its insertion into the membrane. TRC40 then returns during their maturation. The most common modification is
to the cytosol for another round of targeting. Some proteins have asparagine‐linked (or N‐linked) glycosylation [30]. This usually
TMDs that are not hydrophobic enough to bind TRC40. These occurs co‐translationally, as a protein is entering the ER lumen
are kept soluble in the cytosol by calmodulin and inserted into through the translocon. The enzyme that mediates glycosyla-
the ER by another complex called EMC (for ER membrane pro- tion, oligosaccharyl transferase (OST), is tightly associated with
tein complex) [24]. EMC probably has additional functions dur- the translocon [31]. As certain sequences (Asn‐Xxx‐Ser or Asn‐
ing co‐translational membrane protein insertion, but these are Xxx‐Thr, where Xxx is any amino acid except proline) pass
poorly understood. through the translocon into the ER lumen, OST transfers a fully
assembled 14‐sugar core glycan to the asparagine residue
(Figure 10.3, left).
Mis‐targeted proteins are degraded by the cell Not all consensus sites are necessarily glycosylated, and
Despite the extensive machinery dedicated to protein targeting, glycosylation is not uniformly efficient in all cell types or under
failures occur. A secretory or membrane protein that does not all conditions. Although the reasons for this heterogeneity are
reach the ER poses a threat to the cytosol. Such mis‐localized poorly understood, it is likely to influence the protein’s function
proteins are prone to aggregation, are unlikely to fold in the and/or trafficking in subtle ways that are important in a tissue‐
cytosolic environment, and can make inappropriate interactions specific manner. These subtleties notwithstanding, glycosyla-
with other proteins. These adverse consequences are ­underscored tion as a whole is critical to the proper folding and maturation of
by the finding that inherited mutations in a signal sequence can a wide range of proteins, and failure or even partial deficiency
cause disease in humans [25] and transgenic mice [26]. Under of glycosylation can be disastrous for a cell. Indeed, impaired
normal circumstances, such consequences are avoided by qual- processing of glycans causes a variety of severe inherited dis-
ity control pathways that recognize mis‐localized proteins and eases in humans [32].
promptly route them for degradation [27]. Another frequent modification is signal sequence cleavage.
The cytosol contains a family of quality‐control factors that As mentioned above, N‐terminal signal sequences are tempo-
have the capability of recognizing exposed signal sequences and rary sequence elements that are used for correct targeting and
TMDs [28, 29]. When protein targeting works properly, these translocation of secretory and membrane proteins, but which
hydrophobic regions are either removed (in the case of signal are not part of the final mature product. They are removed
sequences) or buried in the membrane (in the case of TMDs). after they have completed their function; this removal usually
Thus, their exposure in the cytosol is a cue that targeting or happens co‐translationally at the translocon. A proteolytic
membrane insertion has failed. The quality‐control factors that enzyme called signal peptidase interacts with the translocon in
recognize these exposed targeting elements recruit ubiquitin such a manner that its active site is positioned adjacent to the
ligases that mark the mis‐localized protein with multiple ubiq- opening of the translocation channel in the ER lumen [33]. As
uitins. The ubiquitinated protein is then delivered to the protea- a nascent polypeptide begins translocation, the boundary
some for degradation. In this manner, the cytosol is kept free of between a signal sequence and mature domain becomes
proteins that belong in other compartments of the cell. exposed to signal peptidase, which efficiently cuts the poly-
peptide at a precise position.
The third widely used modification is attachment of a glyco-
sylphosphatidylinositol (GPI) anchor to some proteins at their
FOLDING AND MATURATION C‐terminus [34]. GPI anchors are essentially phospholipid mol-
OF SECRETORY AND MEMBRANE ecules containing a glycan (i.e. a glycolipid). Attachment of the
PROTEINS GPI anchor to an otherwise soluble protein therefore tethers it to
the membrane. This allows GPI‐anchored proteins to reside on the
Correct ER targeting and translocation of a secretory or mem- cell surface within certain microdomains that are favored by the
brane protein are only the first steps of its eventual matura- lipids to which they are tethered. Furthermore, GPI anchors can
tion into a functional product. Many polypeptides are be cleaved by extracellular phospholipases, allowing for regu-
processed, modified, and assembled with cofactors before lated release (i.e. secretion) of certain proteins under specific
they leave the ER. Furthermore, the linear polypeptide must conditions [35].
be folded into a stable three‐dimensional structure, and in GPI anchor addition occurs in the ER and is mediated by a
many cases, assembled with other proteins. Each protein has GPI–transamidase complex. The choice of which proteins will
its own unique requirements; some are highly specialized in be modified by the transamidase is determined by the presence
their maturation, while others use ubiquitous machinery. at the very C‐terminus of a hydrophobic “signal sequence” for
10:  Protein Maturation and Processing at the Endoplasmic Reticulum 113

ER
export

UGGT

Calreticulin
Folded
ERP57 protein

Core N-linked Folding


glycan cycle
EDEM and
Man I
OST gluc. I
Misfolded
protein
gluc. II
Glucosidase II

ER lumen

Cytosol

Degradation

Figure 10.3  Glycan‐dependent protein folding by lectin chaperones in the ER. Many nascent proteins entering the endoplasmic reticulum (ER)
are modified co‐translationally with a 14‐sugar core glycan containing two N‐acetyl glucosamines (squares), nine mannoses (circles), and three
glucoses (triangles). The glucoses are rapidly trimmed by glucosidases (gluc. I and II), allowing the mono‐glucosylated glycan to recruit a lectin
chaperone (calreticulin), along with associated factors (such as the protein disulfide isomerase‐like molecule ERP57). Protein folding is mediated
by repeated cycles of binding and release from calreticulin, regulated by removal and re‐addition of the terminal glucose. The enzyme UGGT only
adds the glucose to non‐native proteins, ensuring that calreticulin only re‐binds if folding has not been achieved. If folding fails despite repeated
cycles of calreticulin binding, the mannoses get further trimmed (by mannosidases such as EDEM), and the glycan now becomes a target for bind-
ing by another lectin that routes the protein for degradation (see Figure 10.4). The same general scheme is used by other chaperones: they are
recruited to substrates early in their biosynthesis, and undergo cycles of binding and release until the polypeptide either achieves its correct folded
state, or is transferred to the degradation pathway. OST, oligosaccharyl transferase.

GPI anchor addition. The transamidase recognizes this signal, Of course, a nascent protein cannot fold when bound to a chap-
and proteolytically removes it simultaneously with addition of erone. Thus, chaperones periodically release the nascent protein
the GPI anchor (in a transamidation reaction). and transiently afford it an opportunity to fold. If folding is
Beyond these relatively universal ER‐specific modifications, unsuccessful, the chaperone quickly re‐binds. By such repeated
numerous substrate‐specific and cell type‐specific m
­ odifications (and usually energy‐dependent) cycles of binding and release,
have been described. For example, proline and lysine residues in chaperones can facilitate protein folding before inappropriate
collagen are hydroxylated, some secreted signaling molecules interactions lead to nonproductive aggregation.
and morphogens are lipid‐modified (e.g. with cholesterol), and Essential to chaperone function is their ability to selectively
there are other proteolytic processing events that have been bind immature, but not properly folded, conformations of a nas-
described in the ER and other compartements of the secretory cent protein. This selectivity is thought to be achieved in several
pathway. In each case, the modifications are typically critical ways, but typically involves the recognition of polypeptide seg-
for either the protein’s maturation, stability, and/or eventual ments that should normally be shielded in the final folded struc-
function. ture. For example, the buried core of a folded soluble protein
usually contains hydrophobic parts of the polypeptide. Exposure
of such hydrophobic patches, which necessarily means the pro-
Chaperones assist in protein maturation tein is not properly folded, represents a common recognition
Molecular chaperones are essential cellular factors that partici- motif for chaperone binding. Similarly, hydrophilic residues
pate in the maturation of nearly every protein [36]. Chaperones within some TMDs become buried when multiple TMDs
assist in the folding and assembly of nascent polypeptides by are assembled correctly. Exposure of such residues within the
direct binding, thereby preventing inappropriate interactions hydrophobic lipid bilayer can therefore signify a non‐native
with other proteins in the highly crowded cellular environment. conformation. In the same way, exposed cysteines that
114 THE LIVER:  FOLDING AND MATURATION OF SECRETORY AND MEMBRANE PROTEINS

eventually should be disulfide bonded with other cysteines, can term which simply means carbohydrate‐binding). The two
recruit certain types of chaperones. By recognizing these and ­lectin‐based chaperones in the ER are called calnexin (CNX,
other sequence elements, chaperones can distinguish between a  membrane protein) and calreticulin (CRT, a soluble
non‐native and native folded states. ­luminal  protein). They function by a very similar mechanism
Chaperones are recruited almost immediately after a nascent (Figure 10.3) and are recruited to substrates via interaction with
secretory or membrane protein begins translocation. This is a specific isoform of the N‐linked glycan. As mentioned above,
because proteins are translocated in an unfolded conformation, the initial glycan added to a protein consists of 14 sugars: two
which presents an ideal substrate for chaperone binding. There N‐acetyl‐glucosamines, nine mannoses, and three terminal
are many chaperone systems in the ER (see next section), and ­glucoses. Shortly after addition of this core glycan, two of the
the choice of which chaperones are recruited depends on the terminal glucoses are removed by the action of glucosidases,
specific sequence elements within the particular translocating resulting in a mono‐glucosylated form of the N‐glycan that is
protein [37]. The chaperones usually remain on the maturing recognized specifically by CNX or CRT. Upon release from
polypeptide well after protein synthesis and translocation are CNX/CRT, the remaining glucose is removed by a glucosidase
completed, and different subsets of chaperones and maturation and the protein has an opportunity to fold.
factors may associate with the same polypeptide at different If it fails to fold properly, a enzyme termed UDP‐
stages of its maturation. This is especially likely if a protein’s glucose:glycoprotein glucosyltransferase (UGGT) adds a single
maturation is complex, although such non‐model proteins are glucose back to the N‐glycan, thereby making it a substrate for
relatively poorly studied. Because these chaperones are ER resi- CNX/CRT binding and initiating another round of attempted
dents, their repeated binding to immature secretory and mem- folding. Thus, as with the general chaperones discussed above,
brane proteins serves the dual function of preventing the repeated cycles of chaperone binding and release permit a sub-
premature exit of nonfunctional products from the ER. strate to attempt folding while being protected from inappropri-
ate interactions [42]. In this case, a reversible glucose tag is used
Multiple chaperone systems operate to modulate substrate–chaperone interactions, with UGGT
­acting as a “folding sensor” to determine when the substrate
in the ER lumen
has folded. If a substrate passes UGGT inspection, it is not re‐­
Among the various ER chaperones, perhaps the most abundant glucosylated and the glucose‐free N‐glycosylated protein is
and well studied is a luminal protein called BiP. BiP is a mem- released from the CNX/CRT folding cycle. These proteins are
ber of a very large and well‐conserved family of chaperones deemed to be folded, and are exported out of the ER to the Golgi
(called the Hsp70 family) with homologs in all known organ- and beyond.
isms and most cellular compartments [38]. These chaperones Glycosylated proteins do not have an unlimited time to fold.
are ATPases that utilize the energy of ATP hydrolysis to bind The ER also contains mannosidases such as the EDEM family
and release from exposed hydrophobic patches on substrates. and mannosidase I. These are typically very slow, but over time,
The ADP‐bound state of the Hsp70s has high affinity for sub- can remove terminal mannose residues from the N‐glycan. Once a
strate; upon ATP binding, substrates are released so they can particular terminal mannose has been removed, the protein is not
attempt to fold. The steps of ATP binding, ATP hydrolysis, and allowed to attempt folding any longer and is routed for d­ egradation.
exchange of ADP for ATP are all tightly regulated by additional The degradation pathway from the ER is discussed later.
factors broadly termed co‐chaperones. Thus, BiP and its various A distinctive feature of the ER lumen is its oxidizing environ-
co‐chaperones form a universal and general chaperone system ment [43]. This means that cysteines in nascent proteins will
in the ER lumen that recognizes and shields exposed hydropho- typically be oxidized into disulfide bonds with other cysteines
bic patches of secretory and membrane proteins in the process (either within the same protein, or in other proteins). Disulfide
of folding and assembly. bonds can greatly stabilize the folded state of a protein or pro-
Although less extensively studied, another very abundant tein complex, thereby facilitating its function in the harsh extra-
ATPase in the ER lumen termed GRP94 also functions as a cellular environment. For example, antibodies (e.g. IgG) are
chaperone. Like BiP, GRP94 has a cytosolic homolog (termed composed of multiple polypeptide chains that are held together
Hsp90), further illustrating the theme that similar conserved by a series of intra‐ and intermolecular disulfide bonds that lend
mechanisms of protein folding are utilized in multiple cellular it remarkable stability and longevity after secretion.
compartments. Hsp90 interactions with its wide range of sub- The correct formation of disulfide bonds in a nascent protein
strates are regulated by its ATPase cycle and co‐chaperones is mediated by a class of chaperone enzymes termed oxidore-
[39]. It is likely that GRP94 functions analogously to Hsp90, ductases, the most well known of which is protein disulfide
and is important for the folding and maturation of a subset of isomerase (PDI) [44]. These chaperones can interact directly
proteins that include integrins (cell surface adhesion molecules), with substrates (often via exposed hydrophobic patches). Upon
toll‐like receptors (cell surface receptors of the innate immune substrate interaction, an internal disulfide bond within PDI can
system), and immunoglobulins [40]. be reduced concomitantly with oxidation of cysteines in the
Because most proteins transiting the ER become glyco- substrate. Thus, a disulfide bond is effectively transferred from
sylated, eukaryotic cells have evolved to exploit this modifica- PDI to a substrate with which it interacts. The PDI then becomes
tion for numerous purposes including protein stability, solubility, re‐oxidized by other enzymes (e.g. the oxidase Ero1) for another
protein folding, quality control, trafficking, and protein–protein round of substrate oxidation. PDI can also function in reverse,
interactions [41]. In the ER, a major use for N‐linked glycans is and has the capacity of reducing disulfide bonds within a sub-
the recruitment of a class of factors known as lectins (a general strate. This means that by reducing and re‐oxidizing different
10:  Protein Maturation and Processing at the Endoplasmic Reticulum 115

pairs of cysteines, PDI can “shuffle” disulfide bonds until the And yet other ER‐specific processes are highly specialized for
correct protein fold is achieved (and hence the name disulfide unique substrates in only certain cell types. All of these path-
isomerase). Some oxidoreductases are recruited to substrates ways operate simultaneously; some are probably purely parallel
indirectly via interaction with other chaperones. For example, processes, while most are partially overlapping and coordinated
the oxidoreductase ERP57 interacts with CRT (Figure  10.3), events. Thus, the loss or reduced capacity of some maturation
thereby allowing it to operate on glycoproteins [45]. Thus, pathways can likely be compensated, at least temporarily for
different members of the family of oxidoreductases in the
­ many substrates, by other maturation pathways. Indeed, the loss
ER  lumen probably operate on different subsets of client of a surprising number of the major chaperones (e.g. CRT,
substrates. CNX, GRP94, UGGT, and many co‐chaperones) is compatible
Although the above chaperone systems are the most abundant with cellular viability. Yet, in each case, the function of a com-
and well‐studied components of the protein maturation machin- plex organism or differentiated tissues is severely compromised
ery in the ER, they are not the only ones. The sheer volume and (usually leading to embryonic lethality), illustrating the limits of
diversity of substrates that transit the ER necessitates numerous redundancy. Although many basic principles have been deline-
other factors to ensure correct protein folding and maturation. ated using model systems, our knowledge of protein maturation
For example, peptidyl‐prolyl‐isomerases catalyze the change in should be considered fragmentary and incomplete at best.
conformation of proline‐containing peptide bonds between the
cis and trans isomers. Other chaperones, such as tapasin, help in
the assembly of peptides onto MHC class I. Yet other factors are
likely to be specialized for the assembly of multimeric protein
complexes or specifically involved in membrane protein folding QUALITY CONTROL AND THE CULLING
and assembly, although both of these complex processes are OF IMMATURE PROTEINS
rather poorly understood. And finally, there are likely to exist
numerous highly specialized chaperones dedicated to specific The relegation of secretory and membrane protein biosynthesis
substrates with unique requirements. For example, collagen bio- to an intracellular compartment (in contrast to the plasma mem-
synthesis utilizes a chaperone called Hsp47 that is selectively brane as occurs in prokaryotes) provides many advantages to the
expressed in fibroblasts. In the liver, apolipoprotein B biosyn- eukaryotic cell. Perhaps the most obvious and biggest benefit is
thesis uniquely requires a factor termed microsomal triglyceride the opportunity for considerably more control over what does
transfer protein (MTTP) that mediates non‐covalent lipid– and does not gain access to the cell surface. This control is
protein interactions needed for lipoprotein particle assembly. ­manifested in many ways. For example, the cell can effectively
regulate the quantity of a secretory or membrane protein
released to the cell surface independently of the protein’s syn-
Many proteins are assembled into thesis. Thus, key regulatory proteins such as hormones, surface
receptors, or ion channels can be made and stored intracellu-
multimeric complexes larly, then rapidly mobilized to the surface at a moment’s notice.
Numerous proteins function as part of multiprotein complexes. This can provide cells with exquisite temporal control and
Secretory and membrane protein complexes are often assem- responsiveness to changing environmental conditions.
bled in the ER by processes that are relatively poorly under- The other main regulatory benefit is quality control [46]. By
stood. In fact, some of the most important and highly expressed making and inspecting all secretory and membrane proteins
proteins transiting the secretory pathway are multiprotein com- intracellularly, the cell can ensure that only properly folded and
plexes, including immunoglobulins, the T‐cell receptor, ion functional products have access to the surface. This is critical
channels, MHC complexes, and numerous others. It is thought because partially functional or misfolded proteins can often be
that non‐assembled subunits of a larger complex are associated far worse than no protein at all. Thus, in highly complex
with chaperone(s) until they find and associate with the appro- ­multicellular organisms where intercellular communication via
priate interacting partner(s). For example, BiP (IgG‐binding secreted proteins and their receptors are essential for overall
protein) was originally discovered by its association with ­fitness, quality control is a key functional role of the ER.
incompletely assembled IgG heavy chains. How assembly is The basic logic of ER quality control consists of five general
coordinated is not known in most cases, and it is unclear whether steps (Figure  10.4). First, the cell must have mechanisms to
it is simply a stochastic process of subunits finding each other ­recognize a misfolded, misassembled, or otherwise damaged
by diffusion, or if there are more regulated mechanisms that protein. This task is thought to be carried out by chaperones,
facilitate assembly. which, as discussed above, can discriminate between mature
From this discussion, it should be obvious that the ER lumen and non‐native protein structures. Second, once a potentially
is a remarkably complex protein‐folding and maturation factory. misfolded protein is identified, it must be targeted to degrada-
Some features of this factory are universal: it is highly rich in tion machinery. Third, misfolded proteins must be retrotranslo-
evolutionarily conserved chaperone systems that protect nascent cated back to the cytosol. Fourth, once the polypeptide is
proteins from inappropriate interactions and aggregation. Other exposed to the cytosol, it must be ubiquitinated, typically by the
aspects of the ER are unique to this organelle: it has N‐linked same machinery that mediates retrotranslocation. And finally,
glycosylation, lectin‐based chaperones, an oxidizing environ- the ubiquitin is used as a handle for extraction of the protein
ment favoring disulfide bond formation, and a very high flux of from the ER and delivery for degradation by the proteasome.
membrane proteins that need proper folding and maturation. This entire process is often referred to as ER‐associated
116 THE LIVER:  QUALITY CONTROL AND THE CULLING OF IMMATURE PROTEINS

1 2 3
Misfolded
protein Recognition Targeting Retrotranslocation

Lectin Poly-
ubiquitin
Chaperone

Deglycosylation
Ubiquitination
E3 ligase
complex
4

Extraction and
degradation

5
Proteasome

Figure 10.4  The major steps in endoplasmic reticulum (ER)‐associated degradation (ERAD). Schematic depiction of the overall logic of ERAD
using a misfolded glycoprotein as an example. The misfolded protein is recognized by a combination of a lectin (such as OS‐9) and chaperone(s)
(step 1). The recognition complex is then targeted to a retrotranslocation site built around an ER‐resident E3 ubiquitin ligase complex (step 2). The
misfolded protein then dissociates from the targeting complex and is retrotranslocated through a channel that is formed, at least partially, by the E3
ligase complex (step 3). On the cytosolic side, the exposed misfolded protein is ubiquitinated (step 4). The poly‐ubiquitin serves as a handle for
extraction from the membrane by the p97 ATP complex (not depicted) and degradation by the proteasome (step 5). Non‐glycoproteins are thought
to use the same general steps, but do not utilize lectins for recognition.

degradation, or ERAD [47]. This core series of events applies to protein and other cellular factors. Second, the chaperone
both secretory and membrane proteins that fail to mature prop- ­maintains the protein in a largely unfolded state (or facilitates
erly in the ER, although the specific factors involved in their unfolding in some cases), which may be important for its retro-
recognition and retrotranslocation may be different. translocation out of the ER (see next section). And third, the
Although less common, there may be situations where mis- chaperone may interact with or deliver proteins to the machin-
folded proteins in the ER can be routed via alternative pathways ery for retrotranslocation.
for degradation in lysosomes. One class of proteins that seem In the case of glycoproteins, sugar trimming and ER lectins
to  use the lysosomal pathway are GPI‐anchored proteins. It play a key role in degradation [49] in addition to their role during
appears that after attachment of the GPI lipid anchor, the protein the folding cycle. After repeated deglucosylation and regluco-
cannot be efficiently retrotranslocated back to the cytosol. sylation cycles through the CNX/CRT system, mannosidases
Instead, these proteins use vesicular trafficking pathways to (including proteins termed EDEMs) remove terminal mannoses,
access the lysosome for degradation [48]. including the one on which glucosylation normally occurs. This
allows the substrate to exit the CNX/CRT cycle, and now makes
the mannose‐trimmed glycan a substrate for binding by another
Misfolded proteins are recognized lectin termed OS‐9. Mannose trimming by EDEMs and binding
by chaperones by OS‐9 routes substrates into the degradation pathway via OS‐9
Nascent proteins entering the ER typically undergo repeated associations with the retrotranslocation machinery [50, 51].
cycles of chaperone binding and release in an attempt to achieve
their final folded and assembled state. If maturation nonetheless A retrotranslocation pathway for misfolded
fails, the cell must eventually triage the misfolded protein for
degradation. In addition to their role in folding, chaperones also
protein degradation in the cytosol
play an important role in degradation. Most mutant or otherwise The export of misfolded secretory or membrane proteins to the
misfolded proteins are typically found associated with chaper- cytosol for degradation is termed retrotranslocation. This pro-
ones. This is likely to be important for three reasons. First, chap- cess involves components in the ER lumen, within the mem-
erones prevent inappropriate interactions between a misfolded brane, and in the cytosol [47]. At the heart of the retrotranslocation
10:  Protein Maturation and Processing at the Endoplasmic Reticulum 117

reaction is a multiprotein complex centered around one of REGULATION OF ER HOMEOSTASIS


several membrane‐embedded E3 ubiquitin ligases [50, 51].
­ AND ABUNDANCE
These ligase complexes are thought to directly bind to misfolded
membrane proteins via interactions within the membrane. In the The ER is not a static organelle. Its abundance, composition,
case of misfolded proteins in the ER lumen, chaperones and/or and functions adapt to changes in functional need. Highly secre-
lectins associated with the misfolded proteins deliver them to the tory cells and tissues (such as hepatocytes, the exocrine pan-
E3 ligase complex. Then, in a step that remains poorly under- creas, and antibody‐secreting B cells) are packed with ER in
stood, the E3 ligase complex provides a channel through which amounts that greatly exceed the amounts found in nonsecretory
the misfolded protein can access the cytosol [52]. cells. In fact, this correlation between ER abundance and secre-
Once misfolded proteins are cytosolically exposed, they are tory capacity first led to the hypothesis that the ER was involved
poly‐ubiquitinated by the E3 ligase complex. Ubiquitin is a in secretion. It is therefore not surprising that cells have mecha-
small protein whose covalent attachment to substrates marks nisms to sense changes in the load of secretory and membrane
them for degradation. In addition to providing a mark, ubiquitin proteins transiting the ER, and adapt accordingly.
may prevent sliding of the misfolded protein back into the ER The discovery of these pathways came about through the
lumen. The ubiquitinated protein then recruits a large ATPase study of how cells respond to excessive misfolded proteins in
complex containing a protein termed p97 (also called VCP or the ER. The “unfolded protein response” (UPR) is activated
Cdc48) [53]. Together with associated cofactors, the p97 com- whenever the load of secretory and membrane proteins exceeds
plex appears to recognize both substrate and ubiquitin, and uses the capacity of the ER to properly mature and/or metabolize
the energy of ATP hydrolysis to mediate extraction. The p97‐ them, a situation generally termed ER stress [56]. The net result
associated poly‐ubiquitinated substrate is then delivered to the of UPR activation is the initiation of signaling pathways that
proteasome, a large multicatalytic enzyme responsible for most simultaneously alleviate the load on the ER (temporarily), while
protein degradation in the cytosol. upregulating the biosynthetic and maturation machinery to
Although the general scheme of recognition, retrotransloca- expand ER processing capacity. Failure to effectively adapt to
tion, ubiquitination, and degradation is likely to apply for essen- ER stress leads to chronic UPR activation, cellular dysfunction,
tially all misfolded secretory and membrane proteins, the details and, at some point, apoptotic cell death [57]. Chronic ER stress
may differ in a substrate‐specific manner. For example, it is and the consequences arising from it are increasingly appreci-
already clear that glycoproteins and non‐glycoproteins utilize ated to play central roles in various diseases, including many
different (but overlapping) pathways. Similarly, soluble luminal afflicting the liver.
proteins have different requirements for their degradation than
integral membrane proteins. Furthermore, the specific location
of a misfolded domain (whether in the lumen, in the TMD, or The unfolded protein response
cytosol) may influence which retrotranslocation machinery is
communicates ER stress
utilized. Thus, just as multiple parallel and partially overlapping
pathways operate during protein maturation of different types of Non‐native proteins, whether in the process of folding or in
substrates, quality control and retrotranslocation pathways are preparation for degradation, are typically chaperone bound.
similarly complex. The diversity of substrates is simply too vast Hence, exceeding the protein‐processing capacity of the ER
for a single unifying machinery to handle all potential cases. has two consequences: the availability of unoccupied chaper-
Nevertheless, the steps of ubiquitination, p97‐dependent extrac- ones decreases and at least some nascent polypeptides will not
tion, and proteasomal degradation appear to be universal to all have chaperones bound to them. Both of these consequences
substrates. seem to be utilized for sensing ER stress and activating
the UPR.
There are three known ER transmembrane proteins that act as
Physiologic uses of quality control UPR sensors: Ire1, PERK, and ATF6 [56]. The luminal domains
Although quality control and degradation are typically considered of Ire1 and PERK associate with BiP, which maintains these
pathways of aberrant protein disposal, they are physiologically proteins in a monomeric state that is inactive for signaling.
exploited for regulatory control in some cases. Two examples of When the levels of unfolded proteins increase, BiP molecules
central importance to liver physiology are the regulation of apoli- dissociate from Ire1 and PERK and associate instead with
poprotein B and HMG‐CoA reductase. Both proteins are exten- unfolded proteins. Dissociation of BiP from the UPR sensors
sively degraded at the ER by the same quality‐control and ERAD frees them to be activated. In the case of Ire1 (and possibly
machinery used for misfolded proteins [54, 55]. Yet, changes in PERK), misfolded proteins then bind directly to the luminal
the environment or physiologic state of a cell can rapidly and domain, acting as a “ligand” for its activation. ATF6 is also inac-
selectively alter their fate from degradation to maturation. For tive under normal conditions and activated during ER stress;
apolipoprotein B, triglyceride abundance negatively regulates its however, the mechanism of activation is not known. Upon acti-
ER‐associated degradation, leading to greater secretion of lipo- vation, each UPR sensor initiates a different set of downstream
protein particles. Analogously, reduced cholesterol levels prevent reactions that ultimately result in transcriptional activation of
HMG‐CoA reductase degradation, allowing its increased func- genes that increase the protein‐folding capacity of the ER
tional expression in the cholesterol biosynthetic pathway. Thus, (Figure 10.5). After homeostasis in the ER is restored and the
general quality‐control pathways have been exploited for highly levels of unfolded proteins are reduced, the UPR sensors are
selective physiologic regulation in these and other situations. returned to their inactive states.
118 THE LIVER:  REGULATION OF ER HOMEOSTASIS AND ABUNDANCE

Normal conditions ER stress


Abortive
translocation
Endoplasmic reticulum Golgi

BiP

??

Ire1 PERK ATF6

Trafficking and
XbP1 splicing elF2α phosp. proteolysis

Translation
[general]

ATF4 Translation

Transcription
[genes for chaperones, ERAD, metabolism, redox, trafficking]

Figure 10.5  Endoplasmic reticulum (ER) stress initiates a multi‐tiered unfolded protein response. During unstressed normal conditions (left), the
signal transduction factors Ire1, PERK, and ATF6 are inactive. Ire1 and PERK are thought to be held in an inactive state by their binding to the ER
luminal chaperone BiP, while the mechanism of ATF6 repression is unknown. During ER stress (right), the chaperone BiP is engaged with the
excessive quantity of misfolded proteins. This relieves repression of Ire1, PERK, and ATF6, leading to their activation by different mechanisms.
Ire1 catalyzes the splicing of mRNA coding for Xbp1, resulting in production of Xbp1 protein. PERK phosphorylates eIF2α, which causes reduced
global translation and selectively increased translation of ATF4. Relieving the repression of ATF6 leads to its trafficking to the Golgi, where it is
proteolytically cleaved to release a cytosolic domain. Xbp1, ATF4, and ATF6 are all transcription factors that collectively modulate the expression
of hundreds of genes that act to improve the protein‐processing capacity of the ER.

Ire1 activation (via autophosphorylation) causes its cytosolic In addition to these transcriptional changes, other adaptations
domain to act as an endonuclease to mediate the splicing of are also induced in the more acute time frame. As already noted
mRNA coding for Xbp1. Upon Ire1‐mediated removal of an above, the most prominent acute adaptation is a global reduction
intron from Xbp1 mRNA, a functional transcription factor can in translation via phosphorylation of eIF2α by PERK. In addi-
be translated. tion, certain ER‐associated mRNAs are rapidly degraded during
PERK activation leads to its cytosolic kinase domain phospho- ER stress due to the endonuclease activity of Ire1 [60].
rylating the translation initiation factor eIF2α. Phosphorylation of Furthermore, the import of some secretory or membrane pro-
eIF2α leads to a general reduction in overall translation (thereby teins into the ER may be attenuated as a consequence of reduced
reducing the burden of new proteins requiring ER function), while availability of chaperones needed for translocation [61]. This
allowing increased translation of a select few messages, including translocational regulation is substrate‐selective, and may serve
the transcription factor ATF4. The UPR sensor ATF6 trafficks to to reduce the burden during acute stress of non‐essential pro-
the Golgi, where it is proteolytically processed within its trans- teins that are particularly prone to misfolding. Thus, a combina-
membrane domain. This cleavage releases the cytosolic domain tion of general and specific responses can temporarily reduce
of ATF6, which functions as a transcription factor. substrate flux into the ER until the transcriptional response has
Thus, UPR activation leads to the production of at least three an opportunity to improve ER homeostasis.
transcription factors (Xbp1, ATF4, and ATF6), which together
initiate complex changes in gene expression involving hundreds Excessive ER stress leads to programmed
of factors [58, 59]. The most obvious adaptations are the upreg-
cell death
ulation of chaperones (including BiP) needed for protein matu-
ration in the ER. In addition, ER components involved in protein If cells experience ER stress for a prolonged time, they eventu-
translocation, protein degradation, lipid biosynthesis, redox ally undergo apoptosis (also termed programmed cell death)
homeostasis, and others are also increased. The net result is an [57]. Remarkably, the signals that initiate apoptosis derive from
increase in protein‐processing capacity of the ER. Depending Ire1 and PERK, the same molecules charged with alleviating ER
on the extent of this increase, the ER itself may be expanded to stress. In the case of Ire1, two mechanisms have been proposed.
accommodate the higher demand on its function. First, Ire1 may associate with the signal transduction factor
10:  Protein Maturation and Processing at the Endoplasmic Reticulum 119

TRAF2, whose activation can initiate a signaling pathway cul- decisions means that the UPR can have a major impact on
minating in apoptosis. Second, the endonuclease activity of Ire1 organism physiology. Hence, ER stress and UPR signaling have
has been proposed to degrade the mRNAs coding for various been implicated in the pathogenesis of a variety of diseases.
pro‐survival factors, thereby leading to cell death. Evidence for There are two general ways the UPR impacts pathophysiology.
these Ire‐mediated signaling pathways come from experiments First, the UPR may cause the death of stressed cells that would
in cell culture, but their role in vivo remains to be established. have been useful to the organism had they not died. Second, the
Cell death downstream of PERK activation is unambiguous. adaptive features of the UPR might allow dysfunctional cells to
The transcription factor downstream of PERK, ATF4, includes live and grow despite their detrimental effects on the organism.
CHOP as one of its transcriptional targets. CHOP is another tran- Examples of undesirable UPR‐induced cell death include
scription factor which can induce numerous genes, including inherited diseases of protein misfolding and lifestyle‐induced
those that mediate apoptosis. The longer CHOP is activated, the tissue damage. For example, Charcot–Marie–Tooth 1B neurop-
higher its levels and the greater the probability of initiating cell athy can be caused by a mutation in the P0 glycoprotein of
death pathways. Conversely, cells and animals lacking CHOP Schwann cells. Expression of this mutant protein causes ER
are refractory to cell death in response to prolonged ER stress. stress, cell death, and demyelination. Strikingly, deletion of
With both Ire1 and PERK, the signaling responses leading to CHOP in a mouse model of this disease alleviates both cell
cell death are slower than those leading to an improvement of death and neuropathy [65]. In another example, type 2 diabetes
ER homeostasis. This difference in timing means that with mod- imposes a high demand for insulin production from pancreatic
erate or short‐term stresses, ER homeostasis is restored and beta cells. This causes chronic ER stress and eventual beta cell
signaling from Ire1 and PERK is attenuated before they trigger death, the latter of which can be partially reversed by deletion of
cell death. Only with prolonged UPR activation is cell death CHOP [66]. Thus, excessive UPR signaling leading to CHOP
favored. It appears that in multicellular organisms, it is prefera- expression may eliminate cells that would otherwise retain par-
ble to lose a chronically stressed cell than risk the adverse con- tial functionality.
sequences of its prolonged dysfunction or death by necrotic Examples of undesirable UPR‐mediated protection include
mechanisms that may induce inflammation. certain types of cancer [67] and infectious diseases [68]. For
example, in multiple myeloma, an activated UPR facilitates the
high levels of antibody production and rapid cell growth by con-
The UPR is used to expand the ER during stantly improving ER function. In virally infected cells, the pro-
development duction of viral glycoproteins is also facilitated by UPR‐mediated
upregulation of ER functionality. In both of these cases, cell
The transcriptional response downstream of the UPR sensors
death is a more desirable outcome of the UPR than adaptation.
increases expression of factors that build the ER. This includes
The recent development of small molecule modulators of differ-
lipid biosynthesis enzymes, the various proteins in the ER mem-
ent aspects of the UPR [69, 70] may hold promise for interven-
brane and lumen, and the trafficking factors that mediate protein
ing in at least some of the diseases where slightly more or less
export from the ER. These factors improve the protein‐process-
UPR signaling is desirable.
ing capacity of the ER in response to stress as discussed above.
Remarkably, the same signaling pathway is also used to expand
the ER in cell types that are specialized for the high level pro-
duction of secretory or membrane proteins. CONCLUSIONS AND UNKNOWNS
The most dramatic examples of a vastly expanded ER are
antibody‐secreting plasma cells and the exocrine pancreas that It has been over 50 years since the ER was discovered as the site
secretes vast quantities of digestive enzymes. Experiments in of secretory and membrane protein biosynthesis. In that time, a
mice have shown that Xbp1 (the downstream target of Ire1) is cohesive molecular framework has been developed to explain
essential for development of plasma cells [62] and exocrine tis- the mechanisms of selective segregation of these proteins to the
sues [63]. It was initially thought that the high expression of ER, their translocation across or insertion into the ER mem-
secretory proteins induced during development of these cell brane, and their maturation by various chaperones and process-
types caused ER stress, leading to Ire1 signaling and ER expan- ing enzymes. The major pathways for ER‐based quality control
sion. However, later experiments showed that ER expansion and maintenance of ER homeostasis have been elucidated.
does not depend on first inducing secretory protein load; instead, While the broad concepts and core machinery for most of these
the ER expands in preparation for the secretory proteins to come pathways are now in hand, many gaps exist in our knowledge.
[64]. Thus, in addition to its role as a stress sensor, Ire1 signaling How are the processes of targeting and translocation to the ER
is utilized as an intracellular signaling pathway that is develop- regulated, especially for complex proteins or under different
mentally regulated to control ER abundance. Whether the other conditions? How are multi‐spanning membrane proteins relia-
UPR sensors are also developmentally regulated is not known. bly inserted in the precisely desired topology and subsequently
folded and assembled into a functional product? What kind of
specialized maturation and quality‐control pathways operate in
Role of the UPR in various diseases different cell types, and how do they differ from the universal
As we have discussed, the UPR has the capacity to protect cells ones that have been studied so far? How are the various steps in
from the effects of stress, but execute a cell death program if the a protein’s folding and maturation coordinated? How does the
stress is prolonged. This key role in making life‐or‐death cell distinguish between folding proteins en route to a mature
120 THE LIVER:  REFERENCES

product, and a misfolded protein that should be degraded? How 25. Karaplis, A.C., Lim, S.K., Baba, H., Arnold, A., and Kronenberg, H.M.
is the response to ER stress and the choice between adaptation Inefficient membrane targeting, translocation, and proteolytic processing by
signal peptidase of a mutant preproparathyroid hormone protein. J Biol
and death tuned differently among different cell types? And Chem, 1995;270:1629–35.
what is the role of ER stress in the various diseases where it is 26. Rane, N.S., Kang, S.‐W., Chakrabarti, O., Feigenbaum, L., and Hegde, R.S.
observed to be activated? The answers to these and numerous Reduced translocation of nascent prion protein during ER stress contributes
other questions await further study and should occupy cell biol- to neurodegeneration. Dev Cell, 2008;15:359–70.
ogists for the foreseeable future. 27. Juszkiewicz, S. and Hegde, R.S. Quality control of orphaned proteins. Mol
Cell, 2018;71:443–57.
28. Hessa, T., Sharma, A., Mariappan, M. et al. Protein targeting and degradation
are coupled for elimination of mislocalized proteins. Nature, 2011;475:394–7.
29. Itakura, E., Zavodszky, E., Shao, S. et  al. Ubiquilins chaperone and triage
REFERENCES mitochondrial membrane proteins for degradation. Mol Cell, 2016;63:21–33.
30. Aebi, M. N‐linked protein glycosylation in the ER. Biochim Biophys Acta
1. Claude, A. Fractionation of mammalian liver cells by differential centrifuga- Mol Cell Res, 2013;1833:2430–7.
tion: II. Experimental procedures and results. J Exp Med, 1946;84:61–89. 31. Braunger, K., Pfeffer, S., Shrimal, S. et al. Structural basis for coupling pro-
2. Palade, G.E. and Siekevitz, P. Liver microsomes; an integrated morphologi- tein transport and N‐glycosylation at the mammalian endoplasmic reticulum.
cal and biochemical study. J Biophys Biochem Cytol, 1956;2:171–200. Science, 2018;360:215–19.
3. Redman, C.M. and Sabatini, D.D. Vectorial discharge of peptides released by 32. Ng, B.G. and Freeze, H.H. Perspectives on glycosylation and its congenital
puromycin from attached ribosomes. Proc Natl Acad Sci U S A, disorders. Trends Genet, 2018;34:466–76.
1966;56:608–15. 33. Auclair, S.M., Bhanu, M.K., and Kendall, D.A. Signal peptidase I: cleaving
4. Lippincott‐Schwartz, J., Bonifacino, J.S., Yuan, L.C., and Klausner, R.D. the way to mature proteins. Protein Sci, 2012;21:13–25.
Degradation from the endoplasmic reticulum: disposing of newly synthe- 34. Orlean, P. and Menon, A.K. Thematic review series: lipid posttranslational
sized proteins. Cell, 1988;54:209–20. modifications. GPI anchoring of protein in yeast and mammalian cells, or:
5. Blobel, G. Intracellular protein topogenesis. Proc Natl Acad Sci U S A, how we learned to stop worrying and love glycophospholipids. J Lipid Res,
1980;77:1496–500. 2007;48:993–1011.
6. von Heijne, G. Signal sequences. The limits of variation. J Mol Biol, 35. Fujihara, Y. and Ikawa, M. GPI‐AP release in cellular, developmental, and
1985;184:99–105. reproductive biology. J Lipid Res, 2016;57:538–45.
7. UniProt Consortium. UniProt: the universal protein knowledgebase. Nucleic 36. Hartl, F.U., Bracher, A., and Hayer‐Hartl, M. Molecular chaperones in pro-
Acids Res, 2018;46:2699. tein folding and proteostasis. Nature, 2011;475:324–32.
8. Jan, C.H., Williams, C.C., and Weissman, J.S. Principles of ER cotransla- 37. Molinari, M. and Helenius, A. Chaperone selection during glycoprotein
tional translocation revealed by proximity‐specific ribosome profiling. translocation into the endoplasmic reticulum. Science, 2000;288:331–3.
Science, 2014;80(346):716. 38. Mayer, M.P. and Bukau, B. Hsp70 chaperones: Cellular functions and
9. Kutay, U., Hartmann, E., and Rapoport, T.A. A class of membrane proteins molecular mechanism. Cell Mol Life Sci, 2005;62:670–84.
with a C‐terminal anchor. Trends Cell Biol, 1993;3:72–5. 39. Schopf, F.H., Biebl, M.M., and Buchner, J. The HSP90 chaperone machin-
10. Keenan, R.J., Freymann, D.M., Stroud, R.M., and Walter, P. The signal rec- ery. Nat Rev Mol Cell Biol, 2017;18:345–60.
ognition particle. Annu Rev Biochem, 2001;70:755–75. 40. Ansa‐Addo, E.A., Thaxton, J., Hong, F., et al. Clients and oncogenic roles of
11. Rapoport, T.A., Li, L., and Park, E. Structural and mechanistic insights into molecular chaperone gp96/grp94. Curr Top Med Chem, 2016;16:2765–78.
protein translocation. Annu Rev Cell Dev Biol, 2017;33:369–90. 41. Helenius, A. and Aebi, M. Roles of N‐linked glycans in the endoplasmic
12. Voorhees, R.M. and Hegde, R.S. Toward a structural understanding of co‐ reticulum. Annu Rev Biochem, 2004;73:1019–49.
translational protein translocation. Curr Opin Cell Biol, 2016;41:91–9. 42. Tannous, A., Pisoni, G.B., Hebert, D.N., and Molinari, M. N‐linked sugar‐
13. Van den Berg, B., Clemons, W.M., Collinson, I. et al. X‐ray structure of a regulated protein folding and quality control in the ER. Semin Cell Dev Biol,
protein‐conducting channel. Nature, 2004;427:36–44. 2015;41:79–89.
14. Voorhees, R.M., Fernández, I.S., Scheres, S.H.W., and Hegde, R.S. Structure 43. Bulleid, N.J. and van Lith, M. Redox regulation in the endoplasmic reticu-
of the mammalian ribosome‐Sec61 complex to 3.4 Å resolution. Cell, lum. Biochem Soc Trans, 2014;42:905–8.
2014;157:1632–43. 44. Gruber, C.W., Cemazar, M., Heras, B., Martin, J.L., and Craik, D.J. Protein
15. Voorhees, R.M. and Hegde, R.S. Structure of the Sec61 channel opened by a disulfide isomerase: the structure of oxidative folding. Trends Biochem Sci,
signal sequence. Science, 2016;351:88–91. 2006;31:455–64.
16. Shao, S. and Hegde, R.S. Membrane protein insertion at the endoplasmic 45. Oliver, J.D., van der Wal, F.J., Bulleid, N.J., and High, S. Interaction of the
reticulum. Annu Rev Cell Dev Biol, 2011;27:25–56. thiol‐dependent reductase ERp57 with nascent glycoproteins. Science,
17. Higy, M., Junne, T., and Spiess, M. Topogenesis of membrane proteins at the 1997;275:86–8.
endoplasmic reticulum. Biochemistry, 2004;43:12716–22. 46. Sitia, R. and Braakman, I. Quality control in the endoplasmic reticulum pro-
18. Wessels, H.P. and Spiess, M. Insertion of a multispanning membrane pro- tein factory. Nature, 2003;426:891–4.
tein  occurs sequentially and requires only one signal sequence. Cell, 47. Vembar, S.S. and Brodsky, J.L. One step at a time: endoplasmic reticulum‐
1988;55:61–70. associated degradation. Nat Rev Mol Cell Biol, 2008;9:944–57.
19. Shao, S. and Hegde, R.S. A calmodulin‐dependent translocation pathway for 48. Satpute‐Krishnan, P., Ajinkya, M., Bhat, S. et al. ER stress‐induced clear-
small secretory proteins. Cell, 2011;147:1576–88. ance of misfolded GPI‐anchored proteins via the secretory pathway. Cell
20. Panzner, S., Dreier, L., Hartmann, E., Kostka, S., and Rapoport, T.A. 2014;158:522–33.
Posttranslational protein transport in yeast reconstituted with a purified com- 49. Ruddock, L.W. and Molinari, M. N‐glycan processing in ER quality control.
plex of Sec proteins and Kar2p. Cell, 1995;61:561–70. J Cell Sci, 2006;119:4373–80.
21. Matlack, K.E., Misselwitz, B., Plath, K., and Rapoport, T.A. BiP acts as a 50. Denic, V., Quan, E.M., and Weissman, J.S. A luminal surveillance complex
molecular ratchet during posttranslational transport of prepro‐alpha factor that selects misfolded glycoproteins for ER‐associated degradation. Cell,
across the ER membrane. Cell, 1999;97:553–64. 2006;126:349–59.
22. Hegde, R.S. and Keenan, R.J. Tail‐anchored membrane protein insertion into 51. Carvalho, P., Goder, V., and Rapoport, T.A. Distinct ubiquitin‐ligase com-
the endoplasmic reticulum. Nat Rev Mol Cell Biol, 2011;12:787–98. plexes define convergent pathways for the degradation of ER proteins. Cell
23. Stefanovic, S. and Hegde, R.S. Identification of a targeting factor 2006;126:361–73.
for  ­
posttranslational membrane protein insertion into the ER. Cell, 52. Wu, X. and Rapoport, T.A. Mechanistic insights into ER‐associated protein
2007;128:1147–59. degradation. Curr Opin Cell Biol, 2018;53:22–8.
24. Guna, A., Volkmar, N., Christianson, J.C., and Hegde, R.S. The ER mem- 53. Ye, Y., Meyer, H.H., and Rapoport, T.A. The AAA ATPase Cdc48/p97 and
brane protein complex is a transmembrane domain insertase. Science, its  partners transport proteins from the ER into the cytosol. Nature,
2018;359:470–3. 2001;414:652–6.
10:  Protein Maturation and Processing at the Endoplasmic Reticulum 121

54. Wangeline, M.A., Vashistha, N., and Hampton, R.Y. Proteostatic tactics in 63. Lee, A.‐H., Chu, G.C., Iwakoshi, N.N., and Glimcher, L.H. XBP‐1 is
the strategy of sterol regulation. Annu Rev Cell Dev Biol, 2017;33:467–89. required for biogenesis of cellular secretory machinery of exocrine glands.
55. Fisher, E., Lake, E., and McLeod, R.S.. Apolipoprotein B100 quality control EMBO J, 2005;24:4368–80.
and the regulation of hepatic very low density lipoprotein secretion. J Biomed 64. Hu, C.‐C.A., Dougan, S.K., McGehee, A.M., Love, J.C., and Ploegh, H.L.
Res, 2014;28:178–93. XBP‐1 regulates signal transduction, transcription factors and bone marrow
56. Walter, P. and Ron, D. The unfolded protein response: from stress pathway to colonization in B cells. EMBO J, 2009;28:1624–36.
homeostatic regulation. Science, 2011;80(334):1081–6. 65. Pennuto, M., Tinelli, E., Malaguti, M. et al. Ablation of the UPR‐mediator
57. Tabas, I. and Ron, D. Integrating the mechanisms of apoptosis induced by CHOP restores motor function and reduces demyelination in Charcot‐Marie‐
endoplasmic reticulum stress. Nat Cell Biol, 2011;13:184–90. Tooth 1B mice. Neuron, 2008;57:393–405.
58. Harding, H.P., Zhang, Y., Zeng, H. et al. An integrated stress response regu- 66. Song, B., Scheuner, D., Ron, D., Pennathur, S., and Kaufman, R.J. Chop ­deletion
lates amino acid metabolism and resistance to oxidative stress. Mol Cell, reduces oxidative stress, improves β cell function, and promotes cell survival in
2003;11:619–33. multiple mouse models of diabetes. J Clin Invest, 2008;118:3378–89.
59. Travers, K.J., Patil, C.K., Wodicka, L. et al. Functional and genomic analy- 67. Oakes, S.A. Endoplasmic reticulum proteostasis: a key checkpoint in cancer.
ses reveal an essential coordination between the unfolded protein response Am J Physiol Physiol, 2017;312:C93–C102.
and ER‐associated degradation. Cell, 2000;101:249–58. 68. Li, S., Kong, L., and Yu, X. The expanding roles of endoplasmic reticulum
60. Hollien, J., and Weissman, J.S. Decay of endoplasmic reticulum‐localized stress in virus replication and pathogenesis. Crit Rev Microbiol,
mRNAs during the unfolded protein response. Science, 2006;313:104–7. 2015;41:150–64.
61. Kang, S.‐W., Rane, N.S., Kim, S.J. et al. Substrate‐specific translocational 69. Das, I., Krzyzosiak, A., Schneider, K. et al. Preventing proteostasis diseases
attenuation during ER stress defines a pre‐emptive quality control pathway. by selective inhibition of a phosphatase regulatory subunit. Science,
Cell, 2006;127:999–1013. 2015;348:239–42.
62. Iwakoshi, N.N., Lee, A.‐H., and Glimcher, L.H. The X‐box binding pro- 70. Sidrauski, C., Acosta‐Alvear, D., Khoutorsky, A. et  al. Pharmacological
tein‐1 transcription factor is required for plasma cell differentiation and the brake‐release of mRNA translation enhances cognitive memory. Elife,
unfolded protein response. Immunol Rev, 2003;194:29–38. 2013;2:e00498.
Protein Degradation
11 and the Lysosomal System
Susmita Kaushik and Ana Maria Cuervo
Department of Developmental and Molecular Biology, Institute for Aging Research, Albert Einstein College
of Medicine, Bronx, NY, USA

INTRODUCTION TO INTRACELLULAR continuous recycling of the essential building blocks makes


PROTEOLYSIS protein degradation a very conservative and economic process
with a positive net energetic outcome. Lastly, protein degradation
Intracellular proteins undergo continuous synthesis and degra- is also necessary during major cellular remodeling (i.e. embryo-
dation [1, 2]. This constant renewal of the proteome assures its genesis, morphogenesis, cell differentiation), and as a defensive
stability and contributes to regulate its function. Altered or dam- mechanism against harmful agents and pathogens [5–7].
aged proteins are eliminated by the proteolytic systems before Protein degradation is an essential cellular process active and
their accumulation inside cells interferes with normal cell functional in all types of cells and conserved throughout evolu-
function [1, 3]. Damaged proteins are first recognized by tion. Due to the very active metabolic nature of liver and the
molecular chaperones, which facilitate protein refolding/repairing unique properties of this organ as an experimental tool (acces-
(Figure 11.1). However, if the damage is too extensive, or under sibility, relative cellular uniformity, and ease for morphological
conditions unfavorable for protein repair, damaged proteins are analysis), most of the early studies in protein degradation were
targeted for degradation. The proteolytic systems thus consti- performed in liver. In fact, the major proteolytic systems, their
tute, along with intracellular chaperones, essential components essential components, and many of the regulatory mechanisms
of the surveillance mechanisms responsible for cellular quality of protein degradation were first identified in liver [8–12]. The
control. Furthermore, the coordinated balance between protein initial interest in liver proteolysis revolved around the role of
synthesis and degradation also allows cells to rapidly modify this process as a source of energy, in particular during starva-
intracellular levels of subsets of proteins to accommodate to a tion. However, the importance of the proteolytic systems in
changing extracellular environment or to particular cellular con- quality control in this organ has become increasingly appreciated
ditions. Increased protein degradation can enhance the effect of in recent years both in normal liver physiology and in certain
reduced protein synthesis to decrease the levels of particular pathological conditions.
proteins inside cells. Conversely, shutdown of protein degrada-
tion is often used by cells to augment the cellular protein pool.
In addition to these roles in cellular quality control and pro-
tein homeostasis, proteolysis is also often used by cells as an INTRACELLULAR PROTEOLYTIC
extra source of energy when nutrients are scarce [2, 4]. Thus, the SYSTEMS
free amino acids resulting from the breakdown of cellular pro-
teins can be utilized for the synthesis of essential proteins dur- Hepatocytes, like most cells, contain a large variety of proteases
ing starvation, but also as an additional source of energy through both in the cytosol and confined in intracellular compartments.
their breakdown in the urea and creatinine cycles, or by their Some of these proteases, such as caspases, calpains, or different
transformation into glucose through gluconeogenesis in the case secretases, contribute to partial cleavage of cellular proteins, for
of glucogenic amino acids. Despite the associated expenditure the most part with regulatory purposes (reviewed in [13]). For
of energy required to break down intracellular products, this example, many enzymes are synthesized as inactive prozymogens

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
11:  Protein Degradation and the Lysosomal System 123

Life-cycle of proteins
(c)
(a) Folding/ CHAPERONES
assembling
Cytosol (b) REPAIR
Protein ENZYMES
Aggregation
Chaperone
Ribosomes
Refolding/
repair
TRANSLATION
MACHINERY CHAPERONES

(d)
Degradation Proteasome

Disaggregation

ER Lysosomes

PROTEOLYTIC SYSTEMS

Figure 11.1  Quality‐control mechanisms in liver. Two major systems contribute to quality control in liver: chaperones and the proteolytic
systems. Chaperones guarantee proper folding of all proteins after synthesis (a) and act as their companions throughout their biological cycle,
assisting in the unfolding/refolding required for translocation into different subcellular compartments or for assembly into protein complex.
Alterations in protein folding (b, c) are first detected by chaperones, which often facilitate their refolding back into a functional protein. However,
if proper refolding is not possible, the altered proteins are eliminated from the cell by degradation through the proteolytic systems, namely the
ubiquitin–proteasome system and the lysosomes (d).

that need to undergo partial cleavage in order to become active. because some of the subunits of this core are exchangeable, pro-
Similarly, cleavage of particular regions in certain proteins teasomes with different catalytic activities coexist in all cells.
results in changes in their intracellular location while still The activity of the catalytic core is modulated through the
­preserving full functionality. However, the term “proteolytic assembly of different regulatory subunits (19S or 11S) that dock
system,” the main focus of this chapter, has been reserved to on one or both sides of the 20S proteasome to form several
define the group of intracellular proteases, assisting components, proteasome species, likely to participate in different cellular
and intracellular compartments that participate in complete deg- processes (Figure 11.2) [3, 15, 16]. The components of the regu-
radation of proteins into their constitutive amino acids. Two latory subunits are primarily chaperones, ATPases, and enzymes
major proteolytic systems are responsible for most of the intra- able to remove degradation tags from the substrate proteins
cellular protein degradation: the ubiquitin–proteasome system delivered to the proteasome [17, 18]. Although degradation of
(UPS) and the lysosomes [2]. Although the main focus of this substrate proteins directly by the nude 20S proteasome has been
chapter is on the lysosomal system, we briefly review here the described, degradation of a vast number of substrates involves
characteristics of the UPS and refer interested readers to recent the participation of the regulatory complexes [19–21]. The subunits
reviews on this topic for details. of this complex mediate substrate recognition and unfolding,
and often act as the driving force that opens the proteasome bar-
rel and “pushes” substrate proteins into the catalytic core [22].
The proteasome can degrade untagged proteins, but most
THE UBIQUITIN–PROTEASOME SYSTEM substrates of this proteolytic system are selectively targeted for
degradation through the covalent linkage of ubiquitin, a small
The proteasome is a multicatalytic complex with a proteolytic (8 kDa) heat‐stable protein that also undergoes self‐conjugation,
core, known as the 20S proteasome, which in high eukaryotes resulting in the formation of poly‐ubiquitin chains bound to a
results from the association of 28 subunits in four rings stacked lysine in the candidate substrate [14, 20, 21, 23]. Linkage of
as a cylinder‐like structure [3, 14–16]. Three major types of pro- ubiquitin to the cargo proteins is mediated by a series of
teolytic activity have been described in 20S proteasomes, but enzymes – generically known as E ligases – that act sequentially
124 THE LIVER:  THE LYSOSOMAL SYSTEM

E4 Polyub
26S
19S

E3 β

20S β
α
Ubiquitination
Substrate
19S
Ubiquitin

E1
E2

Figure 11.2  The ubiquitin–proteasome system. The proteasome is a multicatalytic complex formed by a catalytic core (20S) and regulatory subunits
(19S and 11S). The catalytic core contains four rings of alpha and beta subunits assembled as a cylinder. Ubiquitin is attached to proteasome
substrates through a process mediated by regulatory enzymes known as ubiquitin ligases (E). E1 activates ubiquitin for assembly to the substrates,
while E2 transfers the activated ubiquitin to the substrate presented by E3. Multiple rounds through this cycle lead to the formation of poly‐ubiquitin
chains covalently linked to the targeted protein. Removal of the ubiquitin tag by the de‐ubiquitinating subunits located in the regulatory 19S particle
is required for internalization of the substrate into the catalytic core. Subunits of the regulatory 19S also modulate the aperture of the cylinder‐like
structure of the 20S proteasome to facilitate substrate access.

to activate ubiquitin, present it to the substrate, and catalyze the Added to the participation of the UPS system in cellular
conjugation [24]. Repeated cycles of ubiquitination result in the homeostasis and protein quality control by the selective removal
formation of the poly‐ubiquitin chain recognized by the chaper- of altered proteins, this proteolytic complex also contributes to
ones and the ubiquitin‐binding subunits of the regulatory com- the regulatory degradation of essential intracellular proteins,
plex of the proteasome. In many instances, ubiquitination is usually in a very rapid manner. This fast degradation of factors
preceded by phosphorylation of the substrate protein, which involved in cell cycle progression, cell division, transcription,
often favors the exposure of the lysine residue that will be used and cell signaling confers a critical regulatory role to the UPS in
for conjugation of the ubiquitin by specific E‐recognizing multiple essential cellular processes [14, 30].
enzymes [20, 21, 23]. Ubiquitination is a universal form of pro- Malfunctioning of the UPS severely impairs cell viability.
tein tagging used not only as a marker for protein degradation Primary defects of components of this system have been identified
through the proteasome but also for degradation through other in different types of aggregopathies, establishing a link between
proteolytic systems, intracellular targeting of proteins to par- the UPS and cellular degeneration [23, 25, 31–35]. The toxic con-
ticular cellular compartments, signaling, enzyme activation, and sequences of proteasome blockage have been extensively explored
regulation of membrane dynamics, among other things [24, 25]. as anticancer treatment since functional proteasome is required for
How the same tagging mechanism could be involved in so dif- cell cycle progression and cell division [36, 37].
ferent cellular processes has been, for a long time, one of the
burning questions in the field. The way in which ubiquitin is
conjugated to the proteins is decisive for the function of the tag
[26, 27]. Ubiquitin contains seven different lysine residues that THE LYSOSOMAL SYSTEM
can potentially be used for conjugation to the substrates.
Linkage through particular residues has been shown now to Lysosomes are single‐membrane organelles containing a wide
determine their recognition by different auxiliary proteins variety of hydrolases, including proteases, lipases, glycosidases,
involved in different cellular pathways, and hence to mediate and nucleotidases, which make them able to degrade all kinds of
the different fates of the conjugated proteins [23]. macromolecules. This high concentration of enzymes was actu-
The UPS plays a critical role as part of the cellular protein ally what determined their original discovery as a cellular frac-
quality‐control system for both cytosolic and secretory proteins tion with high acid phosphatase activity [38]. These initial
[3, 14, 23, 25]. Unfolded proteins, largely newly synthesized biochemical observations by the group of de Duve were shortly
cytosolic proteins that cannot reach their proper folded confor- followed by the ultrastructural studies of Novikoff and col-
mation, and a large subset of post‐translationally damaged pro- leagues, who performed the first electron microscopy study in
teins (oxidized, glycated, etc.) are removed from the cytosol via the isolated fraction and confirmed the presence of single‐membrane
the UPS system [28, 29]. vesicles of 0.1–0.5 μm diameter and spherical shape [39]. Most
11:  Protein Degradation and the Lysosomal System 125

of the main criteria that defined a lysosome have persisted have considerable residual activity at neutral pH. The particular
despite advances in the understanding of the molecular compo- redox conditions of the lysosomal lumen and the low pH facili-
nents and functional relevance of this lytic compartment. The tate unfolding of internalized cytosolic proteins, which allows
analysis of the different routes followed by the cargo (compo- proteases to gain access to internal residues. Changes in the
nents to be degraded by the lysosomes) in order to reach the redox potential and in the luminal pH modulate the proteolytic
lysosomal lumen has given rise to the description of a series of activity inside lysosomes [49, 50].
lysosome‐related compartments, such as endosomes, phago- Defective lysosomal hydrolysis has been associated with
somes, and autophagosomes, all of which will be described in severe human disorders known generically as lysosomal storage
the following sections. disorders (LSDs). These disorders differ in the particular hydro-
Cytosolic vesicles are catalogued as lysosomes, also referred lase affected, but all of them share the presence of engorged
to as secondary lysosomes once they have received cargo, and lysosome‐related compartments in the cytosol of the cell, with
distinguished from the other lysosome‐related vesicles when the consequent cell and organ expansion (i.e. hepatomegaly and
they fulfill the following criteria: single membrane, pH in the splenomegaly are features common to many LSDs). Pathology
range of 4.5–5.5, active (cleaved) hydrolases, presence of lyso- arises both as a result of the defective processing of the substrate
some membrane proteins, and absence of typical endosomal for the hydrolase affected in each LSD, which will limit recy-
markers such as mannose‐6‐phosphate receptor or particular cling of particular essential components, and as a consequence
endosome‐associated Rab proteins [13, 38, 40, 41]. of an abnormally expanded lysosomal system filled with unde-
graded products that alter the lysosomal lumenal conditions.
Thus, accumulation of these products inside lysosomes reduces
The enzymatic machinery
their, pH, changes their redox status, and alters the dynamics of
Lysosomal hydrolases are synthesized as pre‐proenzymes that proteins between lysosomal lumen and membrane, as well as
get activated through proteolytic cleavage as the pH decreases. the ability of lysosomes to fuse with other lysosomes (homo-
After synthesis in the endoplasmic reticulum (ER), all lysoso- typic fusion) or with other vesicular compartments in the cell
mal enzymes traffic through the Golgi, where they become gly- (heterotypic fusion). Readers are referred to reviews for detailed
cosylated and covalently tagged with a mannose‐6‐phosphate descriptions of the different LSDs and current therapeutic
residue. This tag is selectively recognized by the mannose‐6‐ advances for these disorders [51–54].
phosphate receptor at the trans‐Golgi network (TGN), which
helps concentrate lysosomal enzymes inside small vesicles bud-
ding from the Golgi and targeted to endosomes. Sorting at the
Proteins at the lysosomal membrane
Golgi requires the participation of adaptor proteins such as AP1 The lysosomal membrane is far from a mere static barrier to
and of the GGAs or Golgi‐localized gamma‐adaptin ear homol- contain the highly active luminal hydrolases away from the
ogy domain proteins [42, 43]. As the acidification of this cytosol. On the contrary, proteins at the lysosomal membrane
compartment increases with maturation, the hydrolases dis- mediate the essential functions of this organelle. Thus, trans-
sociate from the receptor and are delivered to lysosomes, porters at the membrane allow trafficking of cytosolic compo-
whereas the receptor is recycled back to the Golgi. A percentage nents in and out of the lysosomal compartment, a proton pump
of lysosomal enzymes escape this sorting step, despite the tag- maintains the low luminal pH, and integral membrane proteins
ging, and follow the secretory pathway to be released in the and the membrane’s associated partners facilitate the fusion of
extracellular media. Even though there is growing evidence that lysosomes with other vesicular compartments [42, 49, 55, 56].
lysosomal enzymes can play important functions in extracellu- The most abundant of the integral proteins at the lysosomal
lar matrix remodeling, cell defense, and maintenance of the membrane are the lysosome‐associated membrane proteins
extracellular environment [44], some of the released enzymes (LAMPs) [55–58]. Two different but highly homologous pro-
are internalized back into the cell via endocytosis after interact- teins, LAMP‐1 and LAMP‐2, are the most prominent members
ing with a variant of the mannose‐6‐phosphate receptor present of this family. They are single‐span membrane proteins with a
at the plasma membrane. Double knockdown of the two very heavily glycosylated luminal region and a very short cyto-
mannose‐6‐phosphate receptors has revealed the presence of solic tail. Glycosylation constitutes about 60% of the mass of
an as yet poorly understood mannose‐6‐phosphate‐independent these proteins and is required to preserve their stability, likely
­targeting [45]. by preventing luminal hydrolases from gaining access to their
Although more than 50 different lysosomal hydrolases have peptide core. Despite their high homology, studies in animal
been described in the lysosomal lumen, lysosomal pro- models knocked out for LAMP‐1 or LAMP‐2 have revealed
teases – also known as cathepsins – are of particular interest for marked functional differences between these two proteins.
this chapter. Cathepsins can act both as endo‐ and exopeptidases Knockout of LAMP‐1 has a very mild phenotype and upregula-
(cleaving directly the internal residues of the amino acid tion of LAMP‐2 [59], supporting the supposition that deficiency
sequence of the cargo proteins, or only N‐terminal or C‐terminal of LAMP‐2 can be compensated for by LAMP‐2. In contrast,
residues, respectively) and belong to the families of serine, loss of LAMP‐2 expression results in a dramatic phenotype with
cysteine, and aspartic proteases [46–48]. It is generally accepted major alterations in lysosomal biogenesis, problems in clear-
that degradation of substrate proteins is initiated by endoproteo- ance of autophagic compartments, and altered vesicular traffick-
lytic cleavage that generates peptides amenable to exoprotease ing [60]. This diversity of effects can be attributed to the
degradation. Lysosomal proteases reach maximal activity at the existence in most cells of three isoforms of LAMP‐2 resulting
very acidic pH of the lysosomal lumen, but many of them still from the alternative splicing of the Lamp2 gene [61]. These
126 THE LIVER:  THE LYSOSOMAL SYSTEM

isoforms  –  LAMP‐2A, LAMP‐2B, and LAMP‐2C  –  have of the vacuoles on the cardiac muscle. LAMP‐2 deficiency is the
identical luminal regions but distinctive transmembrane and primary defect in this disease, which is associated with altered
cytosolic tails. As described below, LAMP‐2A is essential for a lysosomal biogenesis and malfunctioning of the autophagic
selective type of autophagy of cytosolic proteins [62], whereas system.
LAMP‐2B seems to play a role in vesicular fusion between lys-
osomes and autophagic vesicles [60, 63]. Lastly, LAMP‐2C has
been described as a receptor for DNA and RNA uptake and deg- Lysosomal pathways for proteolysis
radation in lysosomes [64].
Lysosomes are the common final compartment for degradation
The other abundant type of lysosomal membrane protein is
of components that originate from both the inside of the cell
the lysosomal integral membrane protein (LIMP). The two
(autophagy) and from the plasma membrane or extracellular
members of this family described to date, LIMP1 and LIMP2,
media (heterophagy). Because of the focus of this chapter on
insert into the lysosomal membrane in a hairpin fashion, with
intracellular protein degradation, only a brief description of the
N‐ and C‐termini exposed in the cytosol. LIMP1 has been
heterophagic pathways is presented in this section. Interested
shown to contribute to fusion events in secretory organelles,
readers are referred to recent reviews on endocytosis for more
whereas LIMP2 participates in lysosomal biogenesis through its
details about this fundamental process [69–75].
interactions with vesicle fusion and fission components [65].
Although less abundant than LAMPs and LIMPs, the con-
Heterophagic pathways
served vacuolar proton pump (V‐type H+ ATPase) has been
extensively studied for its essential role in maintaining the acid Hepatocytes, like any other cell type, require continuous sam-
lysosomal lumen [41]. This pump uses the energy derived from pling of their surroundings in order to accommodate changes in
ATP hydrolysis to promote translocation of protons through the the extracellular environment. This constant exchange of infor-
coordinated action of a cytosolic region with ATPase activity mation is attained through endocytosis, which often leads to
and a transmembrane region that acts as a translocon. activation of complex signaling mechanisms originating from
Different transporters are present at the lysosomal membrane the plasma membrane or from the membrane of the internalized
to mediate translocation of hydrolysis byproducts from the lyso- vesicular compartments (endosomes) [76, 77]. The extracellular
somal lumen toward the cytosol for recycling [66]. The only components, or cargo, can be internalized through different
amino acid transporter cloned to date is the cysteine transporter endocytic mechanisms, which mainly differ in the manner in
or cystinosin, a seven‐span transmembrane protein that, when which this cargo is recognized (Figure  11.3). Through fluid‐
mutated, gives rise to the LSD cystinosis [67]. A monosaccha- phase endocytosis or pinocytosis, small portions of the extracel-
ride transporter has also been identified at the lysosomal mem- lular media, including mainly soluble macromolecules and
brane, and mutations leading to functional loss of this transporter micronutrients as well as small particles, are continuously inter-
have been found in the LSD Salla disease. Lysosomal efflux of nalized into small vesicles. The capability of this process is rela-
other products, such as oligosaccharides, small peptides, and tively low but because of its constant nature it can result in the
free fatty acids, has been shown experimentally, but little is internalization of as much as 30% of the plasma membrane per
known about the transporters that modulate their exit from the minute. This forces a coordinate balance between the endocytic
lysosome. The role of lysosomal membrane transporters in process and the hepatocyte secretory mechanisms that will
hepatic homeostasis of metals, such as copper, is also receiving return most of this membrane back to the cell surface. Fluid‐
growing attention [68]. phase endocytosis is a nonsaturable process with relatively low
Lastly, some lysosomal enzymes are located at the lysosomal capacity, used often for internalization of molecules such as
membrane rather than the lumen. Typical examples of these sucrose [78, 79]. One further step in the selectivity and effi-
membrane enzymes are the 70 kDa lysosomal apyrase‐like ciency of cargo internalization is achieved through a second
protein, which contributes to the metabolism of tri‐ and diphos- type of endocytosis known as absorptive endocytosis [80]. In
phate nucleotides, and the acetyl‐CoA:α‐glucosaminide N‐ this case, the internalized molecules interact weakly with the
acetyltransferase, which transfers acetyl groups to heparin plasma membrane before being endocytosed, resulting in some
sulfate. Part of the lysosomal acid phosphatase originally used degree of cargo concentration in the particular region of the
to identify this compartment in the early studies by de Duve is plasma membrane that undergoes invagination. In contrast to
also localized in the lysosomal membrane as an inactive single‐ fluid‐phase endocytosis, absorptive endocytosis is saturable as it
span protein that is released into the lumen by proteolytic depends on the amount of cellular surface available for cargo
processing. interaction, but the concentration step makes it about 100 times
As better understanding is gained on the function of different more efficient than fluid‐phase endocytosis [81]. Many proteins
lysosomal membrane proteins, the number of diseases related to in the blood are internalized in hepatocytes through this mecha-
alterations in these proteins keeps growing. In addition to the nism. The maximal example of selectivity and efficiency in
above‐mentioned LSD due to mutations in transporters, genetic cargo internalization occurs through receptor‐mediated endocy-
alterations in the lamp2 gene have also been connected to a tosis, where cargo molecules first bind directly to particular
muscle vacuolopathy known as Danon disease [63]. Danon dis- receptor proteins at the plasma membrane, which triggers a
ease has been classified as a lysosomal glycogen storage disor- series of events leading to the concentration of receptor cargo
der with normal acid maltase activity in which autophagic in particular regions of the plasma membrane, internalization
vacuoles (AVs) accumulate in all tissues, including liver. in vesicles that form through a complex process of adaptor protein
Patients suffer fatal cardiomyopathy due to the weakening effect assembly, and their delivery to lysosomes through a regulated
11:  Protein Degradation and the Lysosomal System 127

Fluid phase Absorption Receptor-mediated

Cargo interacts
Cargo binds to
Cargo in solution weakly with the
specific receptors
plasma membrane

Nonsaturable Saturable Saturable

Low capacity 2-100 times higher 106-109 higher

Hepatocytes Hepatocytes Hepatocytes

Sucrose Aldolase Insulin-like growth factor

Figure 11.3  Types of endocytosis. Hepatocytes, like most cells in the organism, continuously sample and receive contents and information from
the extracellular environment via endocytosis. The three types of endocytosis simultaneously active in hepatocytes are displayed here. Left:
Fractions of the extracellular media are internalized in a nonselective manner through fluid‐phase endocytosis, a very low‐capability endocytic type.
Middle: The interaction of extracellular components with the membrane prior to internalization explains the higher capability of absorptive endo-
cytosis. This type of endocytosis can be saturated once all the surface of the cell membrane is occupied by interacting molecules. Right: Receptor‐
mediated endocytosis is the most selective type of endocytosis, in which extracellular components interact with particular receptors at the cell
surface. The high capability of this pathway is attained through the concentration of the cargo‐loaded receptor proteins in particular regions of the
plasma membrane where internalization occurs. Examples of extracellular molecules internalized by each pathway are depicted at the bottom.

series of vesicular fusion and fission events [75, 77, 82–85]. and other chapters in this book addressing the importance of the
Like absorptive endocytosis, receptor‐mediated endocytosis is a phagocytic activity of Kupffer cells in chronic liver inflamma-
saturable process, but its capability is 106–109 times higher than tion and in the innate immune response [85–87].
the former.
Independent of the mechanisms that mediate internalization,
Autophagic pathways
lysosomes constitute the final destination of the cargo‐contain-
ing vesicles. These endocytic vesicles, or endosomes, serve also The term autophagy is commonly reserved for the group of dif-
as sorting compartments for molecules such as the receptors that ferent pathways that mediates the lysosomal degradation of
are spared from degradation and are instead delivered back to components present inside cells, including all soluble proteins,
the plasma membrane to continue their function [75]. Endosomes organelles, cellular subcompartments, and particulate protein
mature into acidic lysosomes through fusion and fission events deposits. However, autophagic cargo does not always belong to
regulated by pairs of proteins located in the membranes of both the cell, as extracellular components internalized by phagocyto-
acceptor and donor vesicles [83]. Once a certain degree of acidi- sis can also be the target of degradation by the autophagic
fication is attained, recycling is no longer possible and the cargo ­system through a process described as xenophagy (to point out
is completely degraded by the lysosomal hydrolases. Readers the exogenous nature of the cargo) [88, 89].
are referred to these comprehensive reviews for a detailed Although autophagic degradation has been known since the
description of the molecular components that participate in early discovery of lysosomes and its morphological features
endocytosis [69, 70, 77, 83, 84]. and hormonal regulation have been extensively analyzed using
Lastly, macrophagic cells in the liver such as Kupffer cells liver as a study model, a complete molecular dissection of some
are capable of internalization of extracellular pathogens and of the autophagic pathways was obtained only in the last 15
other large particulate components by phagocytosis, a special- years. The main drive for this “rediscovery” of autophagy has
ized form of heterophagy [83, 84]. In contrast to the discrete been the three different mutagenesis screenings in yeast, which
invaginations of the plasma membrane characteristic of endocy- led to the identification of more than 35 genes generically
tosis, during phagocytosis these professional defense cells known as autophagy‐related genes (ATG) [90–92]. Genetic
undergo major deformity of their plasma membrane and manipulations of these genes  –  knockouts, knockdowns, and
cytoskeletal rearrangements to assure engulfment and internali- overexpressions – have helped in identifying novel physiologi-
zation of the oversized cargo. Phagocytic vesicles or phago- cal roles for autophagy and have linked dysfunction of this
somes fuse with secondary lysosomes through similar lysosomal pathway with prominent human diseases such as
mechanisms to the ones followed by endocytic vesicles for com- cancers, neurodegenerations, myopathies, and different metabolic
plete degradation of their cargo. Readers are referred to reviews disorders [93–96].
128 THE LIVER:  THE LYSOSOMAL SYSTEM

Three major types of autophagy have been described in in mammals) to a lipid (phosphatidylethanolamine) [100]. The
liver and in almost all types of mammalian cells: macroau- series of events required for conjugation and the similarity of
tophagy, microautophagy, and chaperone‐mediated autophagy the enzymes that modulate this process to E ligases make
(CMA) [1, 93, 97] (Figure  11.4). Since the molecular autophagic conjugation resemble, to some extent, the tagging of
mechanisms and physiological relevance of each autophagic proteasome substrates by ubiquitination. Activation of conjuga-
pathway are different, the following sections provide sep- tion, required for the formation of the autophagosome, occurs
arate reviews of the major advances in the current under- through the activation complex. The main components of this
standing of each. complex are Atg6/beclin‐1, a member of the phosphatidylinosi-
tol‐3‐kinase type III (PI3K‐III), and several other exchangeable
proteins that allow this complex to regulate both autophagic and
Macroautophagy endocytic pathways [101].
The original description of a type of nonselective degradation in Although levels of the components of the initiation complex
which whole regions of the cytosol sequestered into double‐ were initially proposed as good markers for autophagic activity,
membrane vesicles were delivered to lysosomes upon fusion of changes in the stoichiometry of their interactions, as well as in
the two vesicular compartments gave rise to the term “macroau- the intracellular location of these complexes, actually determine
tophagy,” to differentiate it from a smaller scale degradation macroautophagy activity [102]. Once the autophagosome is
resulting from small invaginations in the lysosomal membrane, formed, sequestering inside the cytosolic cargo, it is delivered to
or “microautophagy” [8, 9] (Figure  11.4). Macroautophagy is lysosomes in a microtubule‐dependent manner. Both vesicles
quantitatively the most important form of autophagy and also fuse, allowing lysosomal enzymes to gain access to the cargo
the best characterized as a result of the recent identification of (Figure  11.5). Although most of the autophagosomes fuse
the ATG genes. The protein products of these genes organize in directly to secondary lysosomes, forming autophagolysosomes,
four major groups: two conjugation cascades, an initiation com- in almost all cells direct fusion of autophagosomes to late
plex, and a negative regulatory complex [1, 98]. The conjuga- endosomes, forming amphisomes, has also been described
tion events are required for the formation of the limiting [103–105]. This interaction of the autophagic and endocytic
membrane that elongates and seals itself, forming a double‐ pathways does not usually contribute much to the autophagic
membrane vesicle or autophagosome [99] (Figure  11.5). degradation but, under certain conditions when autophago-
Elongation occurs through a complex conjugation of a protein some–lysosome fusion is impaired, it becomes the main route
(Atg5) to another protein (Atg12) and of a protein (Atg8 or LC3 for autophagosome clearance.

Chaperone-mediated autophagy
Microautophagy

LAMP-2A

Macroautophagy

CMA substrate-
chaperone
Lysosome

Autophagosome

Limiting membrane

Figure 11.4  Autophagic pathways in liver. Three major types of autophagy coexist in hepatocytes and almost all mammalian cells: macroau-
tophagy, microautophagy, and chaperone‐mediated autophagy (CMA). In macroautophagy, whole regions of the cytosol, including organelles and
soluble proteins, are sequestered by a limiting membrane that seals to form an autophagosome. This double‐membrane vesicle delivers cargo to
lysosomes for degradation through vesicle‐to‐vesicle fusion. In microautophagy, whole regions of the cytosol are internalized into the lysosomal
lumen in invaginations or projections of the lysosomal membrane that seal to form single‐membrane tubules and vesicles. CMA allows selective
degradation of soluble cytosolic proteins. Substrates are identified by a chaperone/co‐chaperone complex that delivers them to the membrane of the
lysosome where they interact with a receptor protein, the lysosome‐associated membrane protein type 2A (LAMP‐2A). Substrate proteins unfold
and are translocated into the lumen, assisted by a luminal chaperone.
11:  Protein Degradation and the Lysosomal System 129

(a)

(b) (c)

Figure 11.5  Ultrastructure of the autophagic system. (a) Electron micrographs of sections of mouse livers after 6 hours of starvation. Autophagic
vacuoles – autophagosomes and autophagolysosomes – are indicated with arrows. Bottom panels show details of different stages of maturation of
autophagic vacuoles, from early or immature vesicles (left) when cargo is still recognizable, to late or mature vesicles (right) where only degraded
material and membranous structures are observed in the lumen. (b, c) Electron micrographs of lysosomal fractions isolated from the same livers.
Autophagic vacuoles (b) and secondary lysosomes (c) are shown. Insets show representative examples of the type of autophagic/lysosome structure
present in each fraction.

The fourth subset of Atg proteins acts as a negative regulator autophagic pathway are well characterized in yeast, but their
of macroautophagy and has as its central component mTOR, mammalian homologs are yet to be identified.
one of the major nutrient‐sensing intracellular kinases [106– The hormonal regulation of macroautophagy in liver was
108]. mTOR integrates the information received through the well characterized even before the molecular components of
insulin‐signaling pathway and the amino acid receptors, which this pathway were identified. The insulin/glucagon balance in
informs this kinase of the energetic status of the cell. In the pres- blood has a direct effect on macroautophagic activity in liver [9, 11,
ence of nutrients or energy excess, mTOR is activated to exert a 109]. In the postprandial period, the high levels of insulin circu-
negative effect on macroautophagy. However, when nutrients lating in blood repress activation of macroautophagy, whereas
are sparse, mTOR is inactivated, resulting in activation of mac- the gradual increase in circulating levels of glucagon, as the
roautophagy. The downstream effectors of mTOR on the time after nutrient consumption progresses, results in activation
130 THE LIVER:  THE LYSOSOMAL SYSTEM

of macroautophagy to provide the energy and the essential com- Microautophagy


ponents required for the synthesis of proteins necessary during
the nutritional stress. Described initially as a constitutive mechanism for cytosolic
Besides an alternative source for cellular energy, macroau- content turnover in lysosomes in liver, microautophagy is
tophagy is also a major component of the systems responsible still the less understood form of autophagy. The classic defi-
for intracellular quality control [1]. Although selective removal nition of this autophagic pathway resulted from the morpho-
of soluble proteins cannot be attained through this pathway, logical observation of secondary lysosomes containing
macroautophagy contributes to the continuous turnover of orga- multiple vesicular structures in their lumen filled with cyto-
nelles, proteins adopting irreversible insoluble conformations solic content [8, 120] (Figure 11.4). Further studies, mainly
(such as protein inclusions and aggregates), and it is at the fore- in yeast, have demonstrated that the vacuole, equivalent to
front of the cellular response to organelle stress, from ER stress the lysosome in yeast, can trap whole regions of cytosol,
to mitochondrial or Golgi stress [110–112]. In contrast to the including both soluble proteins and organelles, through
lack of selectivity of this pathway for soluble cytosolic proteins, invaginations, tubulations, and projections from this com-
removal of organelles or particulate structures from the cytosol partment’s membrane. This uptake of cytosolic content has
via macroautophagy is a discriminatory process. Thus, only been reproduced recently in vitro using isolated yeast vacu-
those structures identified as nonfunctional or damaged are oles [121, 122]. These studies have revealed that deforma-
sequestered inside autophagosomes for their lysosomal deliv- tion of the membrane requires cytoskeletal components and
ery. Cargo recognition is mediated by a family of soluble is regulated by a subset of gene products specific to this
autophagy receptors (i.e. p62, optineurin, NBR2, Nix) that rec- process. In addition, blockage of particular ATG genes also
ognize specific “marks” in the substrate (i.e. ubiquitination, inhibits microautophagy in yeast, suggesting that macro‐
exposure of specific proteins in the surface of the organelle, and microautophagy share some common components.
etc.) [111, 113]. Based on the mechanism of cargo sequestration, microau-
Although the mechanisms mediating this selective recogni- tophagy has been considered a nonselective type of
tion are under intensive investigation, the importance of macro- autophagy, but, as in the case of macroautophagy, some level
autophagy in quality control has been underscored as a result of of selectivity has been described in relation to the removal
studies with conditional mouse model knockout for macroau- of organelles. Thus, after exposure to conditions associated
tophagy genes in different tissues. Conditional knockout of the with peroxisome proliferation in yeast, microautophagy
essential macroautophagy gene ATG7 in liver [114], the first contributes to the restoration of cellular homeostasis by
established model of this type, revealed, as expected, a lower selectively removing the excess of peroxisomes [123].
ability of the liver of these animals to accommodate proteolysis Lysosomal elimination of peroxisomes has also been
rates to the cellular energetic needs during nutrient deprivation. described in liver after clofibrate‐induced proliferation of
Unexpectedly, damaged organelles and protein inclusions accu- these organelles [124]. Whether or not this process takes
mulate in the livers of these animals even when maintained under place via microautophagy requires future elucidation.
normal nutritional conditions, supporting a critical role for mac- Microautophagy of complete nuclear regions by the vacuole
roautophagy in the maintenance of cellular homeostasis through has recently been described under the name “piecemeal micro-
the continuous removal of intracellular components [114]. autophagy of the nucleus” [125]. In this case, targeted interac-
Similar consequences have been observed upon conditional tions of proteins in the nuclear envelope and in the vacuole
blockage of macroautophagy in other tissues and organs [115– membrane induce the formation of a nuclear bleb that is cap-
117]. This subset of studies underscores a prevalent role for the tured, pinched off, and degraded by the vacuole. Interestingly,
basal activity of macroautophagy, previously classified as only a genetic material is consistently excluded from the protruding
stress‐induced pathway. Whether or not basal and inducible mac- regions of the nucleus.
roautophagy have similar molecular requirements and regulatory Most of the microautophagy‐specific genes do not seem
mechanisms is currently under intensive investigation. to be conserved in higher species, which initially limited the
The high degrading capability of macroautophagy is the study of this process in mammals. However, more recently a
basis for its participation in processes requiring major cellular microautophagy‐like process has been described first in
remodeling, such as embryogenesis, development, cell differ- mammalian cells [126] and lately its presence has also been
entiation, wound healing, and tissue regeneration [93, 96, confirmed in Drosophila [127]. An important difference
118]. In addition, as explained above, macroautophagic with yeast autophagy is that this process takes place in
sequestration of pathogens, free in the cytosol after escaping endosomes and uses the ESCRT machinery, responsible for
the endocytic system or still contained in compartments of the the formation of multivesicular bodies in this compartment,
endocytic and phagocytic pathway (xenophagy), confers mac- for internalization of substrates, hence the name of endoso-
roautophagy an important role in cellular defense against mal microautophagy attributed to this pathway. Endosomal
external aggressors [119]. microautophagy can occur in bulk, by trapping of small por-
The wide variety of functions attributable to macroau- tions of the cytosol, or through selective targeting of pro-
tophagy explains the negative consequences of the malfunc- teins previously identified by the constitutive cytosolic
tioning of this pathway and its association with different chaperone hsc70 [126, 128]. The physiological relevance of
human pathologies. Recent connections established between endosomal microautophagy remains still unclear, although a
macroautophagy failure and liver diseases will be reviewed in recent report proposes a role for this pathway in the rapid
later sections. response to starvation [129].
11:  Protein Degradation and the Lysosomal System 131

Chaperone‐mediated autophagy observed in mice with hepatic CMA blockage [144]. These
results support the theory that the selectivity of CMA might be
In contrast to the autophagic pathways described in previous important under these conditions.
sections, in CMA, soluble cytosolic proteins are delivered to the
lysosomal lumen in a selective manner after crossing the lysoso-
mal membrane [97, 130] (Figure 11.4). The selectivity of this
pathway is conferred by a cytosolic chaperone, hsc70, the con- PROTEIN DEGRADATION IN LIVER
stitutive member of the hsp70 family of chaperones, which DISEASE AND AGING
identifies in the substrate proteins a pentapeptide motif (bio-
chemically related to the pentapeptide KFERQ) and targets The intrinsic high metabolic activity of the liver makes fast
them to the lysosomal surface [131]. The substrate proteins bind adaptation to changes in nutritional conditions and tight quality
there to monomeric forms of LAMP‐2A [62], thus driving the control indispensable for its proper functioning. Consequently,
multimerization of this receptor into the high molecular weight deregulation of the major proteolytic systems described in this
protein complex required for translocation [132]. After unfold- chapter results inevitably in altered liver function and has been
ing, and assisted by a chaperone located in the lysosomal lumen, shown to contribute to the pathogenesis of common liver dis-
substrates cross the lysosomal membrane and are rapidly eases (Figure 11.6) [95, 96].
degraded [133, 134]. Membrane levels of the lysosomal recep- Altered proteasome function contributes, at least in part, to
tor are limiting for this pathway and determine overall the capa- the formation of Mallory bodies, one of the best‐studied types of
bility of particular lysosomes to perform CMA [135]. hepatic protein inclusion, observed in diverse chronic liver dis-
Transcriptional changes, regulated cleavage, and membrane eases such as alcoholic and non‐alcoholic steatohepatitis,
subcompartmentalization of the lysosomal receptor contribute chronic cholestasis, metabolic disorders, and hepatocellular
to regulate the lysosomal levels of LAMP‐2A and consequently neoplasms [145]. Direct inhibition of the proteasome by differ-
the cellular CMA activity [136]. Regulation of this process ent agents, such as ethanol in the case of alcoholic steatohepati-
occurs in part at the lysosomal membrane through a coordinated tis, favors the accumulation of cytokeratins, different
balance between phosphorylation and dephosphorylation events proapoptotic factors, and negative regulators of cytokine signal-
mediated by the TORC2/Akt/PHLPP axis [137]. Other intracel- ing, which contribute to the chronic disease [146].
lular signaling pathways modulating CMA activity include reti- Dysfunction of both the proteasome and the autophagic sys-
noic acid receptor alpha, NRF2, NFAT and humanin, a small tem has been described in α‐1‐antitrypsin deficiency, a protein
mitochondrial peptide [138–141]. conformational disorder resulting from point mutations in a
Basal levels of CMA can be detected in almost all mamma- secretory liver protein that accumulates in hepatocytes, lead-
lian cells, but this pathway is maximally activated under condi- ing to chronic liver inflammation and carcinogenesis [147].
tions of stress [97, 130]. Activation of CMA fulfills two types of The soluble form of the mutant protein is usually degraded by
cellular needs: maintenance of a supply of amino acids neces- the proteasome, whereas the semi‐aggregate and aggregate
sary for protein synthesis under conditions of prolonged starva- forms that accumulate in the ER depend on macroautophagy
tion and selective removal of damaged proteins from the cytosol. for their removal [148]. Furthermore, macroautophagy also
Although macroautophagy is activated in liver during the first contributes to the removal of the altered mitochondrion abun-
hours of starvation to provide the amino acids and other macro- dant in the α‐1‐antitrypsin‐deficient liver [149]. A direct con-
molecules required for different anabolic processes under this nection between the carcinogenesis associated with this
condition, as starvation persists, macroautophagic activity disorder and autophagy has not been established yet. However,
decreases concomitant with a progressive increase in CMA there is extensive evidence for an involvement of autophagy in
activity. hepatocarcinogenesis [150]. Hepatic macroautophagy has
Activation of CMA late in starvation likely allows selective been proposed to have anti‐oncogenic functions, because
degradation of non‐essential proteins to provide amino acids for reduced levels of essential autophagy genes triggers tumori-
the synthesis of proteins essential under this stress condition. genesis in liver and a decrease in autophagy‐related proteins is
CMA is also upregulated during mild oxidative stress and after also common in hepatocarcinoma patients and is associated
exposure of cells to toxic compounds known to induce irrevers- with poor prognosis [151].
ible conformational changes in proteins [142]. Altered proteins Because reduced protein degradation and upregulation of
can thus be eliminated selectively via CMA without affecting protein synthesis are favorable conditions for rapid cell division,
nearby functional proteins. induction of macroautophagy in cancer cells often slows down
Blockage of CMA in cultured cells does not result in major their replication. However, macroautophagy is not completely
changes in cellular homeostasis under basal conditions, as pro- shut down in these cells because it still constitutes a major
teolysis rates are maintained by a compensatory activation of defense against adverse conditions, such as the poor nutritional
macroautophagy [143]. However, although these pathways can environment of the center of hypovascularized tumors or
compensate for one another, they are not redundant, and the lack the damage induced by anti‐oncogenic treatments [119].
of CMA becomes noticeable during stress conditions. Exposure Furthermore, growing evidence supports that macroautophagy
of cells with impaired CMA to different types of stressors can also be utilized by hepatic cancer cells for tumor pro-
reveals their higher sensitivity to stress and a marked increase in gression and an increase in autophagy markers has also been
apoptotic cell death, despite macroautophagy being perfectly associated with poor prognosis and higher rates of recurrence
functional [143]. Similar compensation of macroautophagy was after surgery [152]. Modulators of macroautophagy are
132 THE LIVER:  PROTEIN DEGRADATION IN LIVER DISEASE AND AGING

Alpha-1-antitrypsin deficiency HCV Hepatitis C virus

• Soluble protein degraded by • Promotes ER stress


proteasome ER • Blocks autophagosome
• Aggregates in the ER lysosome fusion
degraded by macroautophagy
• Altered mitochondria

Fatty liver LYS

• Lipid droplet coat proteins, MIT


Cancer
enzymes of lipogenesis are
degraded by CMA Decreased macroautophagy
• Lipid droplets degraded by in many cancers
macroautophagy LD AV
• Chronic lipid stimulus
inhibits macroautophagy
and CMA leading to fatty
liver Aging

• Decreased macroautophagy
• Decreased CMA

Figure 11.6  Autophagy and liver disease. Recent studies have revealed numerous connections between autophagy malfunctioning and liver
pathology. Examples of some of these liver disorders and the changes in the autophagic system under these pathological conditions are shown here.
In certain pathologies, for example α‐1‐antitrypsin deficiency, the soluble protein is degraded by the proteasome, while once the protein forms
aggregates – in this case in the ER – macroautophagy is more adept at degrading these. Sustained lipid challenge leads to blockage of macroau-
tophagy and CMA, which normally degrade the lipid stores, resulting in fatty liver disease. Decreased activity of the proteolytic systems is respon-
sible, in part, for the adverse abnormalities observed in aging liver, while it is beneficial for the cancer cells. Bacteria and viruses subvert the
autophagy machinery for their own replication. AV, autophagic vacuole; ER, endoplasmic reticulum; HCV, hepatitis C virus; LD, lipid droplet;
LYS, lysosomes; MIT, mitochondria.

currently being considered as possible enhancers of anticancer important in liver, as this is the organ that receives the highest
therapies [153, 154]. content of circulating lipids from lipolysis in the adipose tis-
As in other cells, macroautophagy also plays a defensive role sue. However, a maintained lipogenic challenge, such as that
against liver pathogens. Bacteria and viruses are internalized induced by high‐fat diet, has a negative effect on liver macro-
and degraded in lysosomes. However, some pathogens can autophagy [156]. The inability of the liver under these con-
develop ways to overcome the lysosomal system. For example, ditions to reduce the size of the lipid stores through
infection by the hepatitis C virus has been shown to block mac- macroautophagic degradation results in abnormal growth of
roautophagy, but not at the level of autophagosome formation lipid droplets inside hepatocytes, initiating the progression
[155]. Instead, this virus induces ER stress, which promotes toward non‐alcoholic fatty liver. It is also likely that defective
­formation of a large number of autophagosomes but at the same macroautophagy as a result of insulin resistance in conditions
time blocks their fusion with lysosomes. Blockage of autophago- such as aging may be the basis of fatty liver being associated
some formation prevents virus replication, supporting the t­ heory with the metabolic syndrome of aging. For lipophagy to occur,
that the virus might use these double‐membrane vesicles for the lipid droplet coat proteins need to be removed first from
assembly, as previously described for other viruses in other the surface of the lipid droplets. This selective removal is per-
cell types. formed by CMA [159]. Consequently, liver‐specific CMA‐
Connections between liver disease and alterations in null mice present marked steatosis, and derangements in lipid
autophagy are not limited to the role of this pathway in quality metabolism. Several lipogenesis enzymes and lipid carrier
control, but instead extend to the function of this catabolic pro- proteins are also CMA substrates, explaining the severe phe-
cess in the maintenance of the cellular energetic balance. notype of the CMA‐deficient mice [144]. Similar to the effect
Studies support the idea that a defect in macroautophagy could on macroautophagy, acute lipid challenge activates CMA
contribute to the pathogenesis of fatty liver [156–158]. Thus, whereas excess dietary fat inhibits this pathway [160], high-
basal macroautophagy activity has been demonstrated to con- lighting the reciprocal modulatory effect of lipids on
tribute to the regulation of lipid stores in all cells. autophagy, but excess lipids overwhelming and inhibiting the
Macroautophagy of lipid droplets or lipophagy is particularly pathways, mimicking scenarios observed with aging.
11:  Protein Degradation and the Lysosomal System 133

A decrease with age in total rates of protein degradation composition of the lysosomal membrane seem to be behind the
has been described in liver of almost all organisms analyzed, reduced stability of LAMP‐2A. This decrease is initially com-
supporting a negative effect of aging on the intracellular pro- pensated for by an increase in the amount of lysosomal chaper-
teolytic systems [161]. The degradation of short‐lived pro- one that promotes translocation across the membrane but at late
teins, the main substrate of the UPS, is better preserved in old stages this compensation is not enough and CMA activity is
livers, suggesting that the impairment of this system is less severely impaired. In CMA‐null mice, the absence of CMA is
pronounced than that of the lysosomal system. In fact, early initially compensated by upregulation of macroautophagy;
studies in livers from young and old rodents showed no net however, with age, it is not sufficient for maintaining the proteo-
differences in proteasome‐dependent degradation [162, 163]. stasis in response to different stressors [173]. Studies in a trans-
Moreover, the analysis of the three different proteolytic activ- genic mouse model in which the decrease in levels of LAMP‐2A
ities of the 20S proteasome revealed a decrease in some of its in liver was prevented demonstrated that old mice with pre-
activities but an increase in others, suggesting that rather than served CMA activity contain lower levels of oxidized and
dramatic quantitative changes, qualitative changes account aggregate proteins than the livers of their wild‐type littermates
for most of the altered degradation of proteasome substrates [174]. They also show improved ability to respond to stress and
with age [162–164]. Changes in the 20S proteolytic activities with preserved liver function until late in life. These studies support
age may result in part from changes in their subunit composi- the theory that a decline in CMA activity with age could be
tion [165]. Proteomic analyses have revealed the coexistence responsible, at least in part, for the accumulation of damaged
of different subpopulations of proteasomes as cells age. proteins with age, lower resistance to stress, and functional
Defects in ubiquitination steps or in the complex processes decline in this organ [174]. On a positive note, restoration of
that mediate recognition of ubiquitinated proteins by the pro- only one of the different proteolytic systems seems enough to
teasome might also occur with age, but there is still poor induce a major improvement, probably because of the existing
understanding of these changes. cross‐talk among these systems. As more insights are gained
Age‐related changes in the liver lysosomal system at the into the reasons for the functional decrease in other proteolytic
morphological level have been extensively reported in the lit- pathways and as further attempts to correct these defects are
erature. Expansion of the lysosomal vacuolar compartment undertaken, it may be possible to prevent other age‐related
and accumulation of undigested products inside lysosomes in changes in liver and utilize them in the fight against diseases of
the form of an autofluorescent pigment known as lipofuscin the aging liver.
are often used as typical biomarkers of aging [166]. Accumulation
of lipofuscin originates from changes in the proteolytic
susceptibility of the intracellular components and from defects
in autophagy [167]. A reduction in the induction of macroau- CONCLUSION
tophagy in the period in between meals, as well as problems
with the clearance of the already‐formed AVs by lysosomes, The proteolytic systems play a pivotal role in the maintenance
seem to be behind the failure of this autophagic pathway with of liver homeostasis and its energetic balance, and in the fight
age. Because of the important role of this autophagic pathway against intracellular and extracellular aggressors. The differ-
in the maintenance of organelle homeostasis, the decline with ent mechanisms for protein degradation can no longer be con-
age in macroautophagy has been proposed to be behind the sidered as independent units. Increasing numbers of reports
increased number of dysfunctional mitochondria with age; this support the existence of continuous cross‐talk among the dif-
organelle is tightly linked to aging [167]. Defective macroau- ferent autophagic pathways and with the proteasome system.
tophagy induction results from the negative effect on this path- Compensation of one system for another is not complete but
way of the constitutive signaling through the insulin receptor, it is often enough to preserve cellular homeostasis at least
even in the absence of insulin, characteristic of the metabolic under basal nonstressed conditions. Failure of the proteolytic
syndrome of age [168, 169]. Although the molecular basis of systems has been identified as the basis of the pathogenesis of
this defect remains unknown, which prevents any targeted different liver diseases. The recent advances in our under-
restorative effort over macroautophagy, certain interventions standing of the molecular mechanisms behind the different
have proven beneficial in the prevention of the age‐dependent intracellular proteolytic systems, their regulation, and the cel-
macroautophagic decline. Thus, caloric restriction, the only lular consequences of their blockage or altered function could
intervention known to slow down aging, anti‐lipolytic drugs, set the basis for novel therapeutic approaches for the treat-
and intermittent fasting preserve macroautophagy activity in ment of liver diseases.
livers of old rodents at levels comparable to those in young
livers [169–171].
A decrease in CMA activity has been well characterized in
livers of old rodents [135]. The main reason for the CMA ACKNOWLEDGMENTS
decline is a reduction in the lysosomal levels of the LAMP‐2A
receptor with age. Reduced levels of LAMP‐2A are not caused We thank the numerous colleagues in the field of autophagy
by a transcriptional decrease, defects in translation, or altered who through their animated discussions have helped shape this
trafficking to lysosome. Instead, low lysosomal levels of chapter. Work in our laboratory is supported by NIH/NIA
LAMP‐2A originate from increased instability of this protein (AG021904, AG031782), NIH/NIDDK (DK098408), and the
once it reaches the lysosomes [172]. Changes in the lipid generous support of Robert and Renée Belfer.
134 THE LIVER: REFERENCES

REFERENCES 31. Chen, Y., Zhang, Y., and Guo, X. Proteasome dysregulation in human cancer:
implications for clinical therapies. Cancer Metastasis Rev, 2017;36(4):
703–16.
1. Mizushima, N., Levine, B., Cuervo, A., and Klionsky D. Autophagy fights
32. Barac, Y.D., Emrich, F., Krutzwakd‐Josefson, E. et al. The ubiquitin‐proteasome
disease through cellular self‐digestion. Nature, 2008;451:1069–75.
system: a potential therapeutic target for heart failure. J Heart Lung
2. Ciechanover, A. Intracellular protein degradation: From a vague idea thru the
Transplant, 2017;36(7):708–14.
lysosome and the ubiquitin‐proteasome system and onto human diseases and
33. Gilda, J.E. and Gomes, A.V. Proteasome dysfunction in cardiomyopa-
drug targeting. Best Pract Res Clin Haematol, 2017;30(4):341–55.
thies. J Physiol, 2017;595(12):4051–71.
3. Goldberg, A.L. Protein degradation and protection against misfolded or
34. Bonet‐Costa, V., Pomatto, L.C., and Davies, K.J. The proteasome and oxida-
damaged proteins. Nature, 2003;18:895–9.
tive stress in Alzheimer’s disease. Antioxid Redox Signal, 2016;25(16):
4. Singh, R. and Cuervo, A.M. Autophagy in the cellular energetic balance.
886–901.
Cell Metab, 2011;13:495–504.
35. Kammerl, I.E. and Meiners, S. Proteasome function shapes innate and adap-
5. Di Bartolomeo, S., Nazio, F., and Cecconi, F. The role of autophagy during
tive immune responses. Am J Physiol Lung Cell Mol Physiol, 2016;311(2):
development in higher eukaryotes. Traffic, 2010;11(10):1280–9.
L328–36.
6. Cuervo, A.M. and Macian, F. Autophagy and the immune function in aging.
36. Vigneron, N., Abi Habib, J., and Van den Eynde, B.J. Learning from the
Curr Opin Immunol, 2014;29C:97–104.
proteasome how to fine‐tune cancer immunotherapy. Trends Cancer,
7. Mizushima, N. and Levine, B. Autophagy in mammalian development and
2017;3(10):726–41.
differentiation. Nat Cell Biol, 2010;12(9):823–30.
37. Manasanch, E.E. and Orlowski, R.Z. Proteasome inhibitors in cancer ther-
8. Ahlberg, J. and Glaumann, H. Uptake – microautophagy – and degradation
apy. Nat Rev Clin Oncol, 2017;14(7):417–33.
of exogenous proteins by isolated rat liver lysosomes. Effects of pH, ATP,
38. Novikoff, A., Essner, E., and Quintana, N. Golgi apparatus and lysosomes.
and inhibitors of proteolysis. Exp Mol Pathol, 1985;42:78–88.
Fed Proc, 1964;23:1010–22.
9. Deter, R.L., Baudhuin, P., and De Duve, C. Participation of lysosomes in
39. Maggi, V. and Hart, K. Lysosomes and lysosomal enzymes in hearts of ham-
cellular autophagy induced in rat liver by glucagon. J Cell Biol, 1967;35(2):
sters (BIO 14.6 and BBL x7) with congenital cardiomyopathy. Recent Adv
C11–16.
Stud Cardiac Struct Metab, 1973;3:489–95.
10. Marzella, L., Ahlberg, J., and Glaumann, H. Isolation of autophagic vacuoles
40. Jefferies, K.C., Cipriano, D.J., and Forgac, M. Function, structure and regu-
from rat liver: morphological and biochemical characterization. J Cell Biol,
lation of the vacuolar (H+)‐ATPases. Arch Biochem Biophys, 2008;476(1):
1982;93(1):144–54.
33–42.
11. Mortimore, G.E. and Poso, A.R. Intracellular protein catabolism and its con-
41. Braulke, T. and Bonifacino, J.S. Sorting of lysosomal proteins. Biochim
trol during nutrient deprivation and supply. Annu Rev Nutr, 1987;7:539–64.
Biophys Acta, 2009;1793(4):605–14.
12. Mortimore, G.E. and Ward, W.F. Behavior of the lysosomal system during
42. Ghosh, P., Dahms, N., and Kornfeld, S. Mannose 6‐phosphate receptors: new
organ perfusion. An inquiry into the mechanism of hepatic proteolysis. Front
twists in the tale. Nat Rev Mol Cell Biol, 2003;4:202–12.
Biol, 1976;45:157–84.
43. Obermajer, N., Jevnikar, Z., Doljak, B., and Kos, J. Role of cysteine cath-
13. Dice, J. Lysosomal Pathways of Protein Degradation. Landes Bioscience,
epsins in matrix degradation and cell signalling. Connect Tissue Res,
Austin, TX, 2000.
2008;49:193–6.
14. Ciechanover, A. and Stanhill, A. The complexity of recognition of ubiquit-
44. Reczek, D., Schwake, M., Schröder, J. et al. LIMP‐2 is a receptor for lysoso-
inated substrates by the 26S proteasome. Biochim Biophys Acta,
mal mannose‐6‐phosphate‐independent targeting of beta‐glucocerebrosi-
2014;1843(1):86–96.
dase. Cell, 2007;131(4):770–83.
15. Kato, K. and Satoh, T. Structural insights on the dynamics of proteasome
45. Pillay, C.S., Elliott, E., and Dennison, C. Endolysosomal proteolysis and its
formation. Biophys Rev, 2018;10(2):597–604.
regulation. Biochem J, 2002;363(Pt 3):417–29.
16. Pickart, C. and Cohen, R. Proteasomes and their kin: proteases in the
46. Zavasnik‐Bergant, T. and Turk, B. Cysteine proteases: destruction ability
machine age. Nat Rev Mol Cell Biol, 2004;5:177–87.
versus immunomodulation capacity in immune cells. Biol Chem. 2007;388:
17. Rosenzweig, R. and Glickman, M.H. Chaperone‐driven proteasome assem-
1141–9.
bly. Biochem Soc Trans, 2008;36(Pt 5):807–12.
47. Yorimitsu, T., Nair, U., Yang, Z., and Klionsky, D.J. Endoplasmic reticulum
18. Schmidt, M., Hanna, J., Elsasser, S., and Finley, D. Proteasome‐associated
stress triggers autophagy. J Biol Chem, 2006;281(40):30299–304.
proteins: regulation of a proteolytic machine. Biol Chem, 2005;386(8):
48. Dell’Angelica, E.C., Mullins, C., Caplan, S., and Bonifacino, J.S. Lysosome‐
725–37.
related organelles. FASEB J, 2000;14(10):1265–78.
19. Rousseau, A. and Bertolotti, A. Regulation of proteasome assembly and
49. Arunachalam, B., Phan, U.T., Geuze, H.J., and Cresswell, P. Enzymatic
activity in health and disease. Nat Rev Mol Cell Biol, 2018;19:697–712.
reduction of disulfide bonds in lysosomes: characterization of a gamma‐
20. Zhou, P. Targeted protein degradation. Curr Opin Chem Biol,
interferon‐inducible lysosomal thiol reductase (GILT). Proc Natl Acad Sci U
2005;9(1):51–5.
S A, 2000;97(2):745–50.
21. Huibregtse, J.M. UPS shipping and handling. Cell, 2005;120(1):2–4.
50. Settembre, C., Fraldi, A., Jahreiss, L. et al. A block of autophagy in lysoso-
22. Schmidt, M., Lupas, A.N., and Finley, D. Structure and mechanism of ATP‐
mal storage disorders. Hum Mol Genet, 2008;17(1):119–29.
dependent proteases. Curr Opin Chem Biol, 1999;3(5):584–91.
51. Platt, F.M. Emptying the stores: lysosomal diseases and therapeutic strate-
23. Myeku, N. and Duff, K.E. Targeting the 26S proteasome to protect against
gies. Nat Rev Drug Discov, 2018;17(2):133–50.
proteotoxic diseases. Trends Mol Med, 2018;24(1):18–29.
52. Ferreira, C.R. and Gahl, W.A. Lysosomal storage diseases. Transl Sci Rare
24. Ravid, T. and Hochstrasser, M. Diversity of degradation signals in the ubiquitin‐
Dis, 2017;2(1–2):1–71.
proteasome system. Nat Rev Mol Cell Biol, 2008;9(9):679–90.
53. Platt, F.M., d’Azzo, A., Davidson, B.L., Neufeld, E.F., and Tifft, C.J.
25. Mishra, R., Upadhyay, A., Prajapati, V.K., and Mishra, A. Proteasome‐
Lysosomal storage diseases. Nat Rev Dis Primers, 2018;4(1):27.
mediated proteostasis: novel medicinal and pharmacological strategies for
54. Ozen, H. Glycogen storage diseases: new perspectives. World J Gastroenterol,
diseases. Med Res Rev, 2018;38(6):1916–73.
2007;13:2541–53.
26. Li, W. and Ye, Y. Polyubiquitin chains: functions, structures, and mecha-
55. Andrejewski, N., Punnonen, E.L., Guhde, G. et al. Normal lysosomal mor-
nisms. Cell Mol Life Sci, 2008;65(15):2397–406.
phology and function in LAMP‐1‐deficient mice. J Biol Chem,
27. Ikeda, F. and Dikic, I. Atypical ubiquitin chains: new molecular signals.
1999;274(18):12692–701.
‘Protein Modifications: Beyond the Usual Suspects’ review series. EMBO
56. Winchester, B. Lysosomal membrane proteins. Eur J Paediatr Neurol,
Rep, 2008;9(6):536–42.
2001;5(Suppl A):11–19.
28. Goldberg, A.L. Functions of the proteasome: from protein degradation and immune
57. Tanaka, Y., Guhde, G., Suter, A. et al. Accumulation of autophagic vacuoles
surveillance to cancer therapy. Biochem Soc Trans, 2007;35(Pt 1):12–17.
and cardiomyopathy in Lamp‐2‐deficient mice. Nature, 2000;406:902–6.
29. Hanna, J. and Finley, D. A proteasome for all occasions. FEBS Lett,
58. Konecki, D., Foetisch, K., Zimmer, K., Schlotter, M., and Lichter‐Konecki,
2007;581(15):2854–61.
U. An alternatively spliced form of the human lysosome‐associated mem-
30. Wolf, D.H. and Hilt, W. The proteasome: a proteolytic nanomachine of cell
brane protein‐2 gene is expressed in a tissue‐specific manner. Biochem
regulation and waste disposal. Biochim Biophys Acta, 2004;1695(1–3):
Biophys Res Commun, 1995;215:757–67.
19–31.
11:  Protein Degradation and the Lysosomal System 135

59. Cuervo, A. and Dice, J. A receptor for the selective uptake and degradation 90. Levine, B. and Klionsky, D.J. Autophagy wins the 2016 Nobel Prize in
of proteins by lysosomes. Science, 1996;273:501–3. Physiology or Medicine: breakthroughs in baker’s yeast fuel advances in
60. Nishino, I., Fu, J., Tanji, K. et  al. Primary LAMP‐2 deficiency causes X‐ biomedical research. Proc Natl Acad Sci U S A, 2017;114(2):201–5.
linked vacuolar cardiomyopathy and myopathy (Danon disease). Nature, 91. Galluzzi, L., Baehrecke, E.H., Ballabio, A. et al. Molecular definitions of
2000;406:906–10. autophagy and related processes. EMBO J, 2017;36(13):1811–36.
61. Tabuchi, N., Akasaki, K., Sasaki, T., Kanda, N., and Tsuji, H. Identification 92. Dikic, I. and Elazar, Z. Mechanism and medical implications of mamma-
and characterization of a major lysosomal membrane glycoprotein, LGP85/ lian autophagy. Nat Rev Mol Cell Biol, 2018;19(6):349–64.
LIMP II in mouse liver. J Biochem, 1997;122(4):756–63. 93. Mizushima, N. A brief history of autophagy from cell biology to physiol-
62. Chou, H., Vadgama, J., and Jonas, A. Lysosomal transport of small mole- ogy and disease. Nat Cell Biol, 2018;20(5):521–7.
cules. Biochem Med Metab Biol, 1992;48:179–93. 94. Thorburn, A. Autophagy and disease. J Biol Chem, 2018;293(15):
63. Kalatzis, V. and Antignac, C. New aspects of the pathogenesis of cystinosis. 5425–30.
Pediatr Nephrol, 2003;18:207–15. 95. Madrigal‐Matute, J. and Cuervo, A.M. Regulation of liver metabolism by
64. Fujiwara, Y., Hase, K., Wada, K., and Kabuta, T. An RNautophagy/ autophagy. Gastroenterology, 2016;150(2):328–39.
DNautophagy receptor, LAMP2C, possesses an arginine‐rich motif that 96. Schneider, J.L. and Cuervo, A.M. Liver autophagy: much more than just
mediates RNA/DNA‐binding. Biochem Biophys Res Commun, 2015;460(2): taking out the trash. Nat Rev Gastroenterol Hepatol, 2014;11(3):187–200.
281–6. 97. Kaushik, S. and Cuervo, A.M. The coming of age of chaperone‐mediated
65. Geldner, N. and Jurgens, G. Endocytosis in signalling and development. autophagy. Nat Rev Mol Cell Biol, 2018;19(6):365–81.
Curr Opin Plant Biol, 2006;9:589–94. 98. Bento, C.F., Renna, M., Ghislat, G. et al. Mammalian autophagy: how does
66. Fukuda, T., Ewan, L., Bauer, M. et  al. Dysfunction of endocytic and it work? Annu Rev Biochem, 2016;85:685–713.
autophagic pathways in a lysosomal storage disease. Ann Neurol, 2006;59(4): 99. Ohsumi, Y. and Mizushima, N. Two ubiquitin‐like conjugation systems
700–8. essential for autophagy. Semin Cell Dev Biol, 2004;15(2):231–6.
67. Kramer, H. Sorting out signals in fly endosomes. Traffic, 2002;3(2):87–91. 100. Mizushima, N., Noda, T., Yoshimori, T. et al. A protein conjugation system
68. Polishchuk, E.V. and Polishchuk, R.S. The emerging role of lysosomes in essential for autophagy. Nature, 1998;395(6700):395–8.
copper homeostasis. Metallomics, 2016;8(9):853–62. 101. Liang, X., Jackson, S., Seaman, M. et al. Induction of autophagy and inhi-
69. Kaksonen, M. and Roux, A. Mechanisms of clathrin‐mediated endocytosis. bition of tumorigenesis by beclin 1. Nature, 1999;402:672–6.
Nat Rev Mol Cell Biol, 2018;19(5):313–26. 102. Itakura, E., Kishi, C., Inoue, K., and Mizushima, N. Beclin 1 forms two
70. Merrifield, C.J. and Kaksonen, M. Endocytic accessory factors and regula- distinct phosphatidylinositol 3‐kinase complexes with mammalian Atg14
tion of clathrin‐mediated endocytosis. Cold Spring Harb Perspect Biol, and UVRAG. Mol Biol Cell, 2008;19(12):5360–72.
2014;6(11):a016733. 103. Eskelinen, E.L., Illert, A.L., Tanaka, Y. et al. Role of LAMP‐2 in lysosome
71. Traub, L.M. and Bonifacino, J.S. Cargo recognition in clathrin‐mediated biogenesis and autophagy. Mol Biol Cell, 2002;13(9):3355–68.
endocytosis. Cold Spring Harb Perspect Biol, 2013;5(11):a016790. 104. Seglen, P.O., Berg, T.O., Blankson, H. et al. Structural aspects of autophagy.
72. Fares, H. and Grant, B. Deciphering endocytosis in Caenorhabditis elegans. Adv Exp Med Biol, 1996;389:103–11.
Traffic, 2002;3(1):11–19. 105. Seglen, P.O., Gordon, P.B., and Holen, I. Non‐selective autophagy. Semin
73. Saksena, S., Sun, J., Chu, T., and Emr, S.D. ESCRTing proteins in the endo- Cell Biol, 1990;1(6):441–8.
cytic pathway. Trends Biochem Sci, 2007;32(12):561–73. 106. Beugnet, A., Tee, A., Taylor, P., and Proud, C. Regulation of targets of
74. Cullen, P.J. Endosomal sorting and signalling: an emerging role for sorting mTOR signalling by intracellular amino acid availability. Biochem J,
nexins. Nat Rev Mol Cell Biol, 2008;9(7):574–82. 2003;372(Pt 2):555–66.
75. Rappoport, J.Z. Focusing on clathrin‐mediated endocytosis. Biochem J, 107. Dann, S., Selvaraj, A., and Thomas, G. mTOR Complex1‐S6K1 signaling: at the
2008;412(3):415–23. crossroads of obesity, diabetes and cancer. Trends Mol Med, 2007;13:252–9.
76. Predescu, S.A., Predescu, D.N., and Malik, A.B. Molecular determinants of 108. Dennis, P.B., Jaeschke, A., Saitoh, M. et al. Mammalian TOR: a homeo-
endothelial transcytosis and their role in endothelial permeability. Am J static ATP sensor. Science, 2001;294(5544):1102–5.
Physiol Lung Cell Mol Physiol, 2007;293(4):L823–42. 109. Meijer, A.J. Amino acid regulation of autophagosome formation. Methods
77. Meijer, D.K., Scholtens, H.B., and Hardonk, M.J. The role of the liver in Mol Biol, 2008;445:89–109.
clearance of glycoproteins from the general circulation, with special refer- 110. Chu, C.T. Mechanisms of selective autophagy and mitophagy: implications
ence to intestinal alkaline phosphatase. Pharm Weekbl Sci, for neurodegenerative diseases. Neurobiol Dis, 2019;122:23–34.
1982;4(3):57–70. 111. Gatica, D., Lahiri, V., and Klionsky, D.J. Cargo recognition and degradation
78. Mukherjee, S., Ghosh, R.N., and Maxfield, F.R. Endocytosis. Physiol Rev, by selective autophagy. Nat Cell Biol, 2018;20(3):233–42.
1997;77:759–803. 112. Anding, A.L. and Baehrecke, E.H. Cleaning house: selective autophagy of
79. Koenig, J. and Edwardson, J.M. Endocytosis and recycling of G protein‐cou- organelles. Dev Cell, 2017;41(1):10–22.
pled receptors. Trends Pharmacol Sci, 1997;18:276–87. 113. Khaminets, A., Behl, C., and Dikic, I. Ubiquitin‐dependent and independ-
80. Seaman, M.N. Endosome protein sorting: motifs and machinery. Cell Mol ent signals in selective autophagy. Trends Cell Biol, 2016;26(1):6–16.
Life Sci, 2008;65(18):2842–58. 114. Komatsu, M., Waguri, S., Ueno, T. et al. Impairment of starvation‐induced
81. Komada, M. Controlling receptor downregulation by ubiquitination and deu- and constitutive autophagy in Atg7‐deficient mice. J Cell Biol,
biquitination. Curr Drug Discov Technol, 2008;5(1):78–84. 2005;169(3):425–34.
82. Casey, C.A., Lee, S.M., Aziz‐Seible, R., and McVicker, B.L. Impaired recep- 115. Hara, T., Nakamura, K., Matsui, M. et al. Suppression of basal autophagy
tor‐mediated endocytosis: its role in alcohol‐induced apoptosis. J in neural cells causes neurodegenerative disease in mice. Nature,
Gastroenterol Hepatol. 2008;23(Suppl 1):S46–9. 2006;441:885–9.
83. Huynh, K.K., Kay, J.G., Stow, J.L., and Grinstein, S. Fusion, fission, and 116. Komatsu, M., Waguri, S., Chiba, T. et  al. Loss of autophagy in the central
secretion during phagocytosis. Physiology (Bethesda), 2007;22:366–72. nervous system causes neurodegeneration in mice. Nature, 2006;441:880–4.
84. Gao, B., Jeong, W.I., and Tian, Z. Liver: an organ with predominant innate 117. Nakai, A., Yamaguchi, O., Takeda, T. et al. The role of autophagy in cario-
immunity. Hepatology, 2008;47(2):729–36. myocytes in the basal state and in response to hemodynamic stress. Nat
85. Bilzer, M., Roggel, F., and Gerbes, A.L. Role of Kupffer cells in host defense Med, 2007;13:619–24.
and liver disease. Liver Int, 2006;26(10):1175–86. 118. Mizushima, N. and Komatsu, M. Autophagy: renovation of cells and
86. Krenkel, O. and Tacke, F. Liver macrophages in tissue homeostasis and dis- ­tissues. Cell, 2011;147(4):728–41.
ease. Nat Rev Immunol, 2017;17(5):306–21. 119. Levine, B. Cell biology: autophagy and cancer. Nature, 2007;446(7137):745–7.
87. Ferre, N. and Claria, J. New insights into the regulation of liver inflammation 120. Marzella, L., Ahlberg, J., and Glaumann, H. Autophagy, heterophagy,
and oxidative stress. Mini Rev Med Chem, 2006;6(12):1321–30. microautophagy and crinophagy as the means for intracellular degradation.
88. Levine, B. and Klionsky, D. Development by self‐digestion: molecular mecha- Virchows Archiv B Cell Pathol, 1981;36(2–3):219–34.
nisms and biological functions of autophagy. Dev Cell, 2004;6:463–77. 121. Uttenweiler, A., Schwarz, H., and Mayer, A. Microautophagic vacuole
89. Deretic, V. Autophagosome and phagosome. Methods Mol Biol, invagination requires calmodulin in a Ca2+‐independent function. J Biol
2008;445:1–10. Chem, 2005;280:33289–97.
136 THE LIVER: REFERENCES

122. Kunz, J., Schwarz, H., and Mayer, A. Determination of four sequential 150. Galluzzi, L., Pietrocola, F., Bravo‐San Pedro, J.M. et  al. Autophagy in
stages during microautophagy in vitro. J Biol Chem, 2004;279:9987–96. malignant transformation and cancer progression. EMBO J, 2015;34(7):
123. Sakai, Y., Koller, A., Rangell, L., Keller, G., and Subramani, S. Peroxisome 856–80.
degradation by microautophagy in Pichia pastoris. Identification of specific 151. Ding, Z.B., Shi, Y.H., Zhou, J. et al. Association of autophagy defect with a
steps and morphological intermediates. J Cell Biol, 1998;141:625–36. malignant phenotype and poor prognosis of hepatocellular carcinoma.
124. Yokota, S. Degradation of normal and proliferated peroxisomes in rat Cancer Res, 2008;68(22):9167–75.
hepatocytes: regulation of peroxisomes quantity in cells. Microsc Res Tech, 152. Yang, J.D., Seol, S.Y., Leem, S.H. et al. Genes associated with recurrence
2003;61(2):151–60. of hepatocellular carcinoma: integrated analysis by gene expression and
125. Roberts, P., Moshitch‐Moshkovitz, S., Kvam, E. et al. Piecemeal microautophagy methylation profiling. J Korean Med Sci, 2011;26(11):1428–38.
of nucleus in Saccharomyces cerevisiae. Mol Biol Cell, 2003;14:129–41. 153. Song, J., Guo, X., Xie, X. et al. Autophagy in hypoxia protects cancer cells
126. Sahu, R., Kaushik, S., Clement, C.C. et  al. Microautophagy of cytosolic against apoptosis induced by nutrient deprivation through a Beclin1‐
proteins by late endosomes. Dev Cell, 2011;20(1):131–9. dependent way in hepatocellular carcinoma. J Cell Biochem, 2011;112(11):
127. Mukherjee, A., Patel, B., Koga, H., Cuervo, A.M., and Jenny, A. Selective 3406–20.
endosomal microautophagy is starvation‐inducible in Drosophila. 154. Liu, Y.L., Yang, P.M., Shun, C.T. et al. Autophagy potentiates the anti‐can-
Autophagy, 2016;12(11):1984–99. cer effects of the histone deacetylase inhibitors in hepatocellular carci-
128. Tekirdag, K. and Cuervo, A.M. Chaperone‐mediated autophagy and endosomal noma. Autophagy, 2010;6(8):1057–65.
microautophagy: joint by a chaperone. J Biol Chem, 2018;293(15):5414–24. 155. Sir, D., Chen, W., Choi, J. et  al. Induction of incomplete autophagic
129. Mejlvang, J., Olsvik, H., Svenning, S. et al. Starvation induces rapid degra- response by hepatitis C virus via the unfolded protein response. Hepatology,
dation of selective autophagy receptors by endosomal microautophagy. J 2008;48:1054–61.
Cell Biol, 2018;217(10):3640–55. 156. Singh, R., Kaushik, S., Wang, Y. et al. Autophagy regulates lipid metabo-
130. Dice, J. Chaperone‐mediated autophagy. Autophagy, 2007;3:295–9. lism. Nature, 2009;458(7242):1131–5.
131. Chiang, H., Terlecky, S., Plant, C., and Dice, J. A role for a 70 kDa heat 157. Inami, Y., Yamashina, S., Izumi, K. et  al. Hepatic steatosis inhibits
shock protein in lysosomal degradation of intracellular protein. Science, autophagic proteolysis via impairment of autophagosomal acidification
1989;246:382–5. and cathepsin expression. Biochem Biophys Res Commun, 2011;412(4):
132. Bandhyopadhyay, U., Kaushik, S., Vartikovsky, L., and Cuervo, A.M. 618–25.
Dynamic organization of the receptor for chaperone‐mediated autophagy at 158. Fukuo, Y., Yamashina, S., Sonoue, H. et  al. Abnormality of autophagic
the lysosomal membrane. Mol Cell Biol, 2008;28:5747–63. function and cathepsin expression in the liver from patients with non‐alco-
133. Agarraberes, F., Terlecky, S., and Dice, J. An intralysosomal hsp70 is holic fatty liver disease. Hepatol Res, 2014;44(9):1026–36.
required for a selective pathway of lysosomal protein degradation. J Cell 159. Kaushik, S. and Cuervo, A.M. Degradation of lipid droplet‐associated pro-
Biol, 1997;137:825–34. teins by chaperone‐mediated autophagy facilitates lipolysis. Nat Cell Biol,
134. Cuervo, A.M., Dice, J.F., and Knecht, E. A population of rat liver lys- 2015;17(6):759–70.
osomes responsible for the selective uptake and degradation of cytosolic 160. Rodriguez‐Navarro, J.A., Kaushik, S., Koga, H. et al. Inhibitory effect of
proteins. J Biol Chem, 1997;272(9):5606–15. dietary lipids on chaperone‐mediated autophagy. Proc Natl Acad Sci U S A,
135. Cuervo, A.M. and Dice, J.F. Age‐related decline in chaperone‐mediated 2012;109(12):E705–14.
autophagy. J Biol Chem, 2000;275:31505–13. 161. Ward, W. Protein degradation in the aging organism. Prog Mol Subcell Biol,
136. Cuervo, A. and Dice, J. Regulation of lamp2a levels in the lysosomal 2002;29:35–42.
­membrane. Traffic, 2000;1:570–83. 162. Shibatani, T., Nazir, M., and Ward, W. Alteration of rat liver 20S proteas-
137. Arias, E., Koga, H., Diaz, A. et  al. Lysosomal mTORC2/PHLPP1/Akt ome activities by age and food restriction. J Gerontol A Biol Sci Med Sci,
­regulate chaperone‐mediated autophagy. Mol Cell, 2015;59(2):270–84. 1996;51:B316–22.
138. Anguiano, J., Garner, T.P., Mahalingam, M., Das, B.C., Gavathiotis, E., and 163. Shibatani, T. and Ward, W. Effect of age and food restriction on alkaline
Cuervo, A.M. Chemical modulation of chaperone‐mediated autophagy by protease activity in rat liver. J Gerontol, 1996;51:B316–22.
retinoic acid derivatives. Nat Chem Biol, 2013;9(6):374–82. 164. Gray, D., Tsirigotis, M., and Woulfe, J. Ubiquitin, proteasomes, and the
139. Gong, Z., Tasset, I., Diaz, A. et al. Humanin is an endogenous activator of aging brain. Sci Aging Knowledge Environ, 2003;2003:RE6.
chaperone‐mediated autophagy. J Cell Biol, 2018;217(2):635–47. 165. Keller, J., Gee, J., and Ding, Q. The proteasome in brain aging. Ageing Res
140. Pajares, M., Rojo, A.I., Arias, E. et al. Transcription factor NRF2 modu- Rev, 2002;1:279–93.
lates chaperone‐ mediated autophagy through the regulation of LAMP2A. 166. Hochshild, R. Lysosomes, membranes and aging. Exp Gerontol,
Autophagy, 2018;14(8):1310–22. 1970;6:153–66.
141. Valdor, R., Mocholi, E., Botbol, Y. et al. Chaperone‐mediated autophagy 167. Brunk, U. and Terman, A. Lipofuscin: mechanisms of age‐related accu-
regulates T cell responses through targeted degradation of negative regula- mulation and influence on cell function. Free Rad Biol Med, 2002;33:
tors of T cell activation. Nat Immunol, 2014;15(11):1046–54. 611–19.
142. Kiffin, R., Christian, C., Knecht, E., and Cuervo, A. Activation of chaperone‐ 168. Bergamini, E. and Kovacs, J. Exploring the age‐related changes in hor-
mediated autophagy during oxidative stress. Mol Biol Cell, 2004;15:4829–40. mone‐regulated protein breakdown by the use of a physiologic model of
143. Massey, A.C., Kaushik, S., Sovak, G., Kiffin, R., and Cuervo, A.M. stimulation of liver autophagy, in Protein Metabolism in Aging Modern
Consequences of the selective blockage of chaperone‐mediated autophagy. Aging Research, vol. 9 (eds. H. Segal et al.), Wiley‐Liss, New York, 1990,
Proc Natl Acad Sci U S A, 2006;103:5905–10. pp. 361–70.
144. Schneider, J.L., Suh, Y., and Cuervo, A.M. Deficient chaperone‐mediated 169. Donati, A., Cavallini, G., Paradiso, C. et  al. Age‐related changes in the
autophagy in liver leads to metabolic dysregulation. Cell Metab, 2014;20(3): autophagic proteolysis of rat isolated liver cells: effects of antiaging dietary
417–32. restrictions. J Gerontol, 2001;56:B375–83.
145. Strnad, P., Zatloukal, K., Stumptner, C., Kulaksiz, H., and Denk, H. 170. Donati, A., Cavallini, G., Carresi, C. et al. Anti‐aging effects of anti‐lipol-
Mallory‐Denk‐bodies: lessons from keratin‐containing hepatic inclusion ytic drugs. Exp Gerontol, 2004;39:061–7.
bodies. Biochim Biophys Acta, 2008;1782:764–74. 171. Martinez‐Lopez, N., Tarabra, E., Toledo, M. et al. System‐wide benefits of
146. Osna, N. and Donohue, T.J. Implication of altered proteasome function in intermeal fasting by autophagy. Cell Metab, 2017;26(6):856–71 e5.
alcoholic liver injury. World J Gastroenterol, 2007;13:4931–7. 172. Kiffin, R., Kaushik, S., Zeng, M. et al. Altered dynamics of the lysosomal
147. Perlmutter, D.H. The cellular response to aggregated proteins associated receptor for chaperone‐mediated autophagy with age. J Cell Sci,
with human disease. J Clin Invest, 2002;110:1219–20. 2007;120:782–91.
148. Kamimoto, T., Shoji, S., Hidvegi, T. et al. Intracellular inclusions contain- 173. Schneider, J.L., Villarroya, J., Diaz‐Carretero, A. et  al. Loss of hepatic
ing mutant alpha1‐antitrypsin Z are propagated in the absence of autophagic chaperone‐mediated autophagy accelerates proteostasis failure in aging.
activity. J Biol Chem, 2006;281:4467–76. Aging Cell, 2015;14(2):249–64.
149. Teckman, J., An, J.K., Blomenkamp, K., Schmidt, B., and Perlmutter, D. 174. Zhang, C. and Cuervo, A.M. Restoration of chaperone‐mediated autophagy
Mitochondrial autophagy and injury in the liver in alpha 1‐antitrypsin defi- in aging liver improves cellular maintenance and hepatic function. Nat
ciency. Am J Physiol Gastrointest Liver Physiol, 2004;286(5):G851–G62. Med, 2008;14:959–65.
Peroxisome Assembly,
12 Degradation, and Disease
Rong Hua1 and Peter K. Kim1,2
1
Cell Biology Program, Hospital for Sick Children, Toronto, ON, Canada
2
Department of Biochemistry, University of Toronto, Toronto, ON, Canada

INTRODUCTION clinical phenotype compared to PBD‐ZSD [3]. PBDs result


from ­mutations in any of the 14 peroxins, encoded by PEX
Peroxisomes are single membrane‐bound organelles that are genes, which are involved in peroxisome assembly and forma-
ubiquitous to most eukaryotic cells (Figure  12.1). Since their tion. The characterization of peroxins has been greatly assisted
initial description by J. Rhodin in 1954 [1] and biochemical by both genetic and proteomic studies in yeast and mutated
characterization by C. de Duve and P. Baudhuin in 1966 [2], Chinese hamster ovary (CHO) cells with defects in peroxisome
much more has become known about their physiological roles number [4, 5]. In mammalian cells, 14 peroxins have been iden-
in various organisms. Enclosed by a single membrane, their tified (Table 12.1), whereas there are at least 32 and 22 peroxins
matrix contains a diverse array of enzymes involved in multiple in yeast and plants respectively [6]. RCDP1 is due to mutations
metabolic reactions. In mammalian cells, the main functions of in PEX7, whereas mutations in the other 13 peroxins lead to
peroxisomes include β‐oxidation of very‐long‐chain fatty acids ZSD (Table 12.1).
(VLCFAs), detoxification of hydrogen peroxide, as well as the Like other organelles, peroxisomes need to interact and com-
biosynthesis of a number of lipid‐related molecules, such as bile municate with their surroundings, including other intracellular
acids and plasmalogens. compartments, for their proper formation and maintenance, as
The importance of peroxisomes in these catabolic and ana- well as for their metabolic functions. A common method of
bolic reactions is best illustrated in a group of genetically intracellular communication between organelles is via the mem-
­inherited diseases known collectively as the peroxisomal disor- brane contact site (MCS), where the opposing membranes of the
ders. These peroxisomal metabolic diseases are categorized into two organelles are physically tethered. The close proximity
two general groups: single metabolic enzyme disorders, which between organelles allows for the exchange of small molecules
are caused by a defect in single enzymes required for peroxiso- and signals [7]. Peroxisomes have long been seen to be juxta-
mal functions, and peroxisomal biogenesis disorders (PBDs), posed to other organelles in electron micrographs, especially the
which result from a defect in peroxisome formation [3]. PBDs endoplasmic reticulum (ER). However, the molecular mecha-
represent a spectrum of disorders which include the Zellweger nisms underlying these peroxisome contact sites and their
syndrome spectrum (PBD‐ZSD) (Zellweger syndrome, neona- implications for peroxisome biogenesis and function are only
tal adrenoleukodystrophy, and infantile Refsum disease) and recently starting to be explored [8].
rhizomelic chondrodysplasia punctate type 1 (RCDP1). These The estimated half‐life for peroxisomes is approximately
diseases were named based solely on the clinical description, 2–3  days, suggesting that a balance between peroxisome
but not on biochemical or genetic analysis, with Zellweger ­biogenesis and their degradation is crucial for maintaining per-
­syndrome being the most severe clinical manifestation of PBD oxisome homeostasis. Compared to what is known about the bio-
and infantile Refsum disease showing the mildest symptoms genesis of peroxisomes, less is known about the mechanism of
[3]. RCDP1, however, presents a different biochemical and their degradation. It is thought that peroxisomes are mainly

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
138 THE LIVER:  PEROXISOME BIOGENESIS

(a) (b)

Figure 12.1  Peroxisomes in mammalian cells. (a) Live cell confocal image of COS7 cell expressing the peroxisome marker UB‐RFP‐SKL (red).
DAPI (blue), nucleus. (b) Live cell confocal image of COS7 cell expressing the ER marker ssGFP‐KDEL (green) and the peroxisome marker
UB‐RFP‐SKL (red). The white box indicates the magnified area shown in the right panels. Scale bars, 10 μm.

Table 12.1  Identified mammalian PEX genes involved in peroxisome biogenesis


Gene Function in biogenesis Molecular function Disease
PEX1 Matrix protein import ATP‐dependent dislocation of PEX5 ZS, NALD, IRD
PEX2 Matrix protein import E3 ubiquitin ligase for pexophagy ZS, IRD
PEX3 PMP targeting; de novo formation Membrane anchor of PEX19 ZS
PEX5 Matrix protein import PTS1 receptor ZS, NALD
PEX6 Matrix protein import ATP‐dependent dislocation of PEX5 ZS, NALD
PEX7 Matrix protein import PTS2 receptor RCDP type 1
PEX10 Matrix protein import PEX5 recycling ZS, NALD
PEX11 Proliferation Elongation of peroxisome ZS
PEX12 Matrix protein import PEX5 recycling ZS, NALD, IRD
PEX13 Matrix protein import Docking complex ZS, NALD
PEX14 Matrix protein import Docking complex ZS
PEX16 PMP targeting; de novo formation PMP receptor at ER and peroxisome ZS
PEX19 PMP targeting; de novo formation Group II PMP receptor and chaperone ZS
PEX26 Matrix protein import Membrane anchor of PEX1/PEX6 ZS, NALD, IRD
IRD, infantile Refsum disease; PMP, peroxisome membrane protein; PTS, peroxisomal targeting signal; NALD, neonatal adrenoleukodystrophy; RCDP, rhizomelic chondrodys-
plasia punctata; ZS, Zellweger syndrome.

degraded through an autophagy pathway called pexophagy, in PEROXISOME BIOGENESIS


which damaged or superfluous peroxisomes are wrapped around
by a double‐membrane vesicle called a autophagosome. This then Growth and division model
fuses with a lysosome and the engulfed content is degraded [9].
Thus far, several key players that induce and regulate pexophagy Since their first discovery, the origin of peroxisomes has been a
have been identified, and a peroxisome‐specific autophagy subject of debate. Based on electron micrographs of hepatocytes
­adaptor protein (NBR1) has recently been characterized [10]. in the early 1960s, peroxisomes were reported to be in juxtapo-
This chapter will focus largely on our understanding of per- sition with the ER, suggesting the involvement of the ER in the
oxisomes in the mammalian system. In particular, the chapter origin of peroxisomes. This led C. de Duve and P. Baudhuin to
will cover the current models for peroxisome biogenesis, the propose in their early review that peroxisomes were formed by
import mechanisms for both peroxisomal membrane and matrix budding from the ER [2]. This view was later challenged
proteins, their communication with other organelles, especially by Paul B. Lazarrow and Yukio Fujiki in the mid 1980s, when
in the context of lipid transport, and the degradation of peroxi- they introduced the “growth and division” model in their author-
somes. In order to appreciate the importance of peroxisome itative review [11]. This postulation was based on the seminal
regulation (formation and degradation), their metabolic func- finding that peroxisomal proteins, both the matrix and mem-
tions and related diseases will also be discussed. brane proteins, were synthesized on free polyribosomes in the
12:  Peroxisome Assembly, Degradation, and Disease 139

cytosol and then imported post‐translationally into peroxisomes. biogenesis of peroxisome starts with the recruitment of mem-
Peroxisomes were therefore considered to be autonomous brane modulators that induce membrane curvature and the
­organelles, like mitochondria and chloroplasts, that proliferate assembly of pre‐peroxisomal structures/vesicles at the ER.
by the fission of pre‐existing ones. PEX16 is one of the most likely candidates to be involved in this
Today, we know that peroxisomes are formed via two mecha- process, since it is capable of recruiting a variety of PMPs to the
nisms: de novo biogenesis and growth and division. De novo ER, including the early biogenesis factor PEX3 and the PEX11
biogenesis involves the emergence of pre‐peroxisomal vesicles family proteins, which are known membrane modulators that
from the ER, which then fuse together to form mature peroxi- induce peroxisomal membrane elongation [23]. These PMPs
somes. The growth and division mechanism is mediated by together with PEX16 concentrate at specialized subdomains on
elongation and fission machinery, consisting of the PEX11 the ER and thereby lead to the formation of the pre‐peroxisomal
­family, dynamin‐like protein 1 (DLP1/Drp1), MFF, and fission structures. These pre‐peroxisomal structures are then released
1 (FIS1). The PEX11 family proteins induce the elongation of from the ER via an unknown mechanism. One aspect worth
the peroxisomal membrane and also help to recruit the fission mentioning is that the release of these pre‐peroxisomal struc-
machinery to their site of action, leading to the final scission of tures is unlikely via the secretory pathway, as neither COPI or
the membrane [12, 13] (Figure 12.2). COPII is required for the de novo peroxisome formation in
Pex3‐deficient cells upon reintroduction of wild‐type PEX3
[24]. These ER‐derived structures will continuously import
Model for de novo peroxisome biogenesis
PMPs via the action of PEX3 and PEX19, leading to the assem-
This view of peroxisomes as an autonomous organelle came bly of the matrix protein import machinery and eventually give
under question due to a number of key observations. First, rather rise to the formation of new mature peroxisomes.
than multiplying solely by growth and division from pre‐­
existing ones, peroxisomes were found to be formed de novo in
peroxisome‐deficient cells. In cells with mutation or deletion in Contribution of de novo biogenesis versus
any of the three peroxisome biogenesis factors – PEX3, PEX16,
fission in peroxisome formation
and PEX19 – no peroxisomal membrane ghosts were detected.
Upon reintroduction of the wild‐type peroxin, peroxisomal It is now generally accepted that the ER contributes to the
membrane structure could be detected within a short time frame, de novo formation of peroxisomes, especially in cells without
followed by the restoration of matrix protein import [14–16]. pre‐existing ones. However, the extent that the two mecha-
This raises the possibility that peroxisomes could also be formed nisms – ER‐derived de novo formation versus growth and divi-
from yet‐to‐be‐identified pre‐peroxisomal structures or derived sion  –  are utilized in normal mammalian cells is still under
de novo from other organelles. In fact, such small pre‐peroxisomal debate. One hypothesis is that both mechanisms exist, but they
structures/vesicles have been identified in multiple yeast spe- are activated by different metabolic signals. The constitutive
cies, including Yarrowia lipolytica, Saccharomyces cerevisiae, formation of peroxisomes may be mediated by the de novo path-
and Hansenula polymorpha [17–19]. These pre‐peroxisomal way, whereas the rapid peroxisome proliferation is achieved by
vesicles were shown to contain distinct pools of membrane pro- growth and division. Rodent liver cells respond rapidly to clofi-
teins and undergo heterotypic fusion to form new peroxisomes. bric acid treatment by doubling and in some cases, quadrupling
Interestingly, these structures were rapidly degraded via the number of peroxisomes within days. This rapid proliferation
autophagy in pex mutant cells if they did not mature into fully of peroxisomes is mediated by the activation of peroxisome
functional peroxisomes [19]. ­proliferator‐activated receptor α (PPARα), which is a nuclear
Although similar pre‐peroxisomal structures have not yet hormone receptor that regulates the expression of genes associ-
been identified in mammalian cells, recent studies have pointed ated with lipid metabolism [25]. One of the genes that was
to the ER as a contributor for the de novo formation of mam- shown to be upregulated by PPARα activation is PEX11α, which
malian peroxisomes. Using electron microscopy, peroxisomes is a tissue‐selective gene expressed most prominently in the
in mouse dendritic cells were found to be surrounded by a dou- liver [26]. Since the PEX11 family proteins are involved in per-
ble‐membrane structure that is continuous with the rough ER oxisome elongation and fission, it is likely that under conditions
[20]. Similar structures were also observed in CHO cells lack- of rapid peroxisome proliferation, peroxisomes multiply mainly
ing PEX16 [21] and hepatocytes of PEX5 knockout mice [22]. via growth and division. At the same time, the de novo biogen-
Moreover, as discussed in detail below, most mammalian esis pathway is the constitutive mechanism for maintaining
­peroxisomal membrane proteins (PMPs) can be sorted to per- ­peroxisome number during normal cell growth.
oxisomes via the ER. The initial ER recruitment and the subse- On the other hand, it is also possible that the two mechanisms
quent ER‐to‐peroxisome trafficking of these PMPs appear to work in concert with each other. As discussed in detail below,
depend on PEX16 [23]. Therefore, the ER is likely the mem- mammalian PMPs can be targeted to peroxisomes either directly
brane source for these pre‐peroxisomal structures in both yeast from the cytosol or via the ER, suggesting a role of the ER in
and mammals, although the precise mechanism involved constitutively providing PMPs to pre‐existing peroxisomes for
remains unclear. their maintenance. Although the exact mechanism remains
Conceptually, de novo formation of peroxisomes requires the unclear, it is likely that the ER‐targeted PMPs are sorted to per-
action of three groups of peroxins involved in (i) membrane oxisomes via vesicular transport between the two organelles.
modulation, (ii) peroxisomal membrane assembly/PMP import, Therefore, the ER‐derived pre‐peroxisomal vesicles could either
and (iii) matrix protein import (Figure  12.2). The de novo mature into newly formed peroxisomes (i.e. de novo formation),
140 THE LIVER:  PEROXISOMAL PROTEIN TARGETING

Endoplasmic
reticulum
i. Membrane
remodeling

ii. PMP import

de novo biogenesis
Membrane modulators

Peroxins for PMP import Pre-peroxisome

Peroxins for matrix protein import

Other PMPs iii. Matrix protein


import
Matrix proteins
PEX5/PEX7

Mature
peroxisome

Elongation
(PEX11β)

Fission
Constriction
(FIS1/MFF/DLP1)
(FIS1/MFF/DLP1)

Growth and division

Figure 12.2  Peroxisome biogenesis. Mammalian peroxisomes can propagate via growth and division from pre‐existing ones, or be formed de
novo from other organelles, such as the ER. The de novo formation of peroxisomes starts when membrane modulators are recruited to specialized
subdomains on the ER (i). The peroxisome biogenesis factor PEX16 is likely involved in this process, in which it recruits other peroxisome mem-
brane proteins (PMPs) to the, ER, including the early biogenesis factor PEX3 and the known membrane modulator, the PEX11 family. The accu-
mulation of these PMPs together with PEX16 on the ER leads to the formation of the pre‐peroxisomal structures, which are then released via an
unknown mechanism. These ER‐derived structures will continuously import PMPs via the action of PEX3 and PEX19 (ii), leading to the assembly
of the matrix protein import machinery (iii), and eventually mature into fully functional peroxisomes. On the other hand, peroxisomes can also
propagate via growth and division from existing ones. Peroxisome proliferation is initiated by membrane elongation and followed by constriction
and final scission of the membrane. The PEX11 family induce membrane elongation and recruit the fission machinery proteins (DLP1/Drp1, FIS1,
and MFF) to their site of action, leading to the fission of peroxisomes.

or fuse with pre‐existing ones to provide proteins and/or lipid PEROXISOMAL PROTEIN TARGETING
content for their growth and division. One possible reason for
the existence of this specialized ER‐to‐peroxisome trafficking Targeting of peroxisomal membrane proteins
pathway in the mammalian system, as well as perhaps in plants,
is that these organisms need to maintain a much larger popula- The targeting of most if not all PMPs to peroxisomal mem-
tion of peroxisomes than most yeasts. Mammalian and plant branes requires the action of three peroxins named PEX3,
cells typically contain 100–1000 peroxisomes, whereas only PEX16, and PEX19 [23]. Mutation in any of these three
2–10 peroxisomes are found in most yeasts. Hence, the rela- ­peroxins results in the absence of any detectable peroxisomal
tively large numbers of peroxisomes imply that there is more membrane structures; whereas other PMPs are either rapidly
demand on the ER in terms of providing both protein and lipid degraded in the cytosol or mislocalized to other organelles, such
content required for maintaining the steady‐state number of as the mitochondria [14, 15, 27]. Currently, PMPs are thought to
­peroxisomes in both mammals and plants. be targeted to peroxisomes via two distinct but not mutually
12:  Peroxisome Assembly, Degradation, and Disease 141

exclusive pathways, the Group I pathway (i.e. indirectly via proteins is PEX7, which contains six WD40 repeats involved in
the ER) and the Group II pathway (i.e. directly from the cytosol) binding to PTS2 sequences [34]. Unlike PEX5, PEX7 requires
(Figure 12.3a). species‐specific accessory proteins to carry out its function. For
In the direct Group II pathway, most PMPs are thought to be example, the accessory protein for PEX7 is the orthologous
synthesized on free ribosomes in the cytosol and targeted post‐ redundant Pex18p/Pex21p in yeast S. cerevisiae, whereas in
translationally to peroxisomes. PEX19 acts as a soluble receptor mammalian and plant cells it is the longer splicing isoform of
that binds to these newly synthesized PMPs and stabilizes them PEX5 (i.e. PEX5L) [35]. In addition, a small number of peroxi-
in a soluble state in the cytosol [28]. The cargo–PEX19 complex somal matrix proteins do not contain a PTS1 or PTS2 sequence.
is then guided to peroxisomal membranes via the interaction These proteins are thought to be imported into peroxisomal
between PEX19 and the docking factor PEX3. The targeting of matrix either by binding to PEX5 in a PTS1‐independent
PEX3 itself also depends on PEX19 as well as PEX16 on the ­manner or via a piggy‐back mechanism by binding to other
membranes. However, how PMPs are inserted into the peroxiso- PTS‐containing cargos [29, 36].
mal lipid bilayers remains unclear.
The indirect Group I pathway is thought to depend on PEX16, Docking
which is a PMP with dual localization in both peroxisomes and
the ER [23]. PEX16 is shown to target to peroxisomes exclu- The docking complex at the peroxisomal membranes consists of
sively via the ER, as the domains required for its co‐translational two peroxins, PEX13 and PEX14. These proteins bind with
insertion to the ER and subsequent ER‐to‐peroxisome targeting each other and have binding affinity for both PTS receptors.
have been recently identified [23]. At the ER membranes, Interestingly, they contain several binding sites for PEX5,
PEX16 is able to recruit a wide range of PMPs that differ ­suggesting a dynamic stepwise binding of the cargo–receptor
in membrane topology and function to the ER. Interestingly, complex. Moreover, PEX14 is believed to present with the
both PEX3 and PEX19 are dispensable for PEX16‐mediated ­initial binding site for the cargo–receptor complex, as it has a
recruitment of PMPs to the ER, implying mechanistic differ- higher binding affinity than PEX13 [31, 37].
ences between PMP targeting to the ER and to the peroxisomes
[23]. However, how the ER‐targeted PMPs are subsequently Translocation and cargo release
transported to the existing peroxisomes remains largely unclear. The precise mechanism of cargo protein translocation across the
One possibility is that this protein‐targeting pathway is medi- peroxisomal membrane is not fully understood. The current
ated by vesicular transport between the two organelles, in which results favor a model by which peroxisomal matrix proteins
pre‐peroxisomal vesicles budding from the ER fuse with translocate across the membranes through a protein‐conducting
­existing peroxisomes to provide them with essential protein channel or a translocon. In this model, a transient pore/
and/or lipid content. protein‐conducting channel is assembled upon the docking of
the cargo–receptor complex at the peroxisomal membrane, and
Import of peroxisomal matrix proteins disassembled after cargo translocation across the membrane
In contrast to the protein import pathways of the ER and mito- [38]. One candidate protein that may be involved in the forma-
chondria, peroxisomal matrix proteins translocate across the tion of such transient pores is PEX5, as an aqueous pore up to
lipid bilayer as fully folded proteins, and in some cases as oligo- 9 nm in diameter can be formed when Pex5p‐containing com-
meric complex [29, 30]. Therefore, the mechanism of matrix plex purified from yeast peroxisomal membranes is reconsti-
protein import to peroxisomes is more closely related to the tuted into liposomes [39]. It is unclear whether a distinct pore
nucleus, except that no stable translocon or channel is involved exists for the PTS2 cargos.
in the process. The peroxisomal matrix protein import cascade The cargo protein needs to be separated from the receptor
can be divided into four steps: (i) cargo binding; (ii) docking; before translocating across the membrane. However, the mecha-
(iii) translocation and cargo release; and (iv) receptor ubiquit- nism of cargo–receptor dissociation is still a matter of debate.
ination and recycling [31] (Figure 12.3b). The import receptor PEX5 comprises two structurally and func-
tionally autonomous parts. The N‐terminal part interacts with
Cargo binding the docking complex at the peroxisomal membranes during
cargo translocation, whereas the C‐terminal part binds to the
Peroxisomal matrix proteins contain one of the two peroxisomal PTS1 cargos. Upon binding of its N‐terminal part to the docking
targeting signals called PTS1 or PTS2. Most peroxisomal matrix complex, the C‐terminal segment of PEX5 is thought to undergo
proteins exhibit PTS1, which consists of a tripeptide sequence a conformational change, thereby inducing cargo release [40].
(S/A/C)‐(K/R/H)‐(L/M) at the extreme C‐terminal end [32].
The PTS1 signal is recognized and bound by the soluble ­receptor
Receptor ubiquitination and recycling
PEX5, which contains a helix bundle at its N‐terminus, and a
series of tetratricopeptide repeats (TPRs) within the C‐terminal After cargo release, the receptors are recycled back to the cyto-
half of the protein. The TPRs are involved in PTS1 signal bind- sol where they either enter another import cycle or are targeted
ing; while the helix bundle is involved in binding to the docking for proteasomal degradation. This process requires the action of
complex (PEX13 and PEX14) at the peroxisomal membrane [33]. the RING complex (PEX2/PEX10/PEX12) and AAA‐type
The less common PTS2 is a nonapeptide with the consensus ATPase complex (PEX1/PEX6/PEX26). During the import
sequence (R/K)‐(L/V/I)‐X5‐(H/Q)‐(L/A) located at the N‐terminus cycle, the PTS1‐receptor PEX5 is regulated by different types
of the protein. The cytosolic receptor for PTS2‐containing of  ubiquitination [41]. Mono‐ubiquitination of PEX5 at the
142 THE LIVER:  PEROXISOMAL LIPID TRAFFICKING

(a)
19
Chaperone?
PMP

PMP PMP

19 19
PMP
3 3
16 PMP
16 PMP

Peroxisome ER lumen
lumen

Direct pathway Indirect pathway

(b)
PTS1

PTS1

5 5 Ub

II
IV

Ub 6
PTS1

13 14 14 5 5 14 2 10 12 5 26

III

PTS1

Figure 12.3  Peroxisomal protein import. (a) Peroxisomal membrane proteins can be targeted to peroxisomes either directly from the cytosol (i.e.
the direct Group II pathway) or via the ER (i.e. the indirect Group I pathway). The newly synthesized PMPs is recognized by the soluble receptor
PEX19, which is then guided to the peroxisomal membrane via its interaction with the docking factor PEX3. PEX3 itself is targeted to peroxisomes
via PEX16. The ER targeting of PMPs in mammalian cells is proposed to depend on PEX16. (b) Peroxisomal matrix protein import occurs via a
multistep cycle that is divided into four steps: cargo binding (I); docking (II); translocation and cargo release (III); and receptor ubiquitination and
recycling (IV). The peroxisomal matrix protein containing either PTS1 or PTS2 signal is recognized and bound by the soluble receptor PEX5 or
PEX7, respectively. The cargo–receptor complex is then docked at the peroxisomal membrane via the docking complex (PEX13 and PEX14). A
transient pore that is mainly constituted by PEX5 is assembled upon the docking of cargo–receptor complex, and thereby allowing the translocation
of cargo across the membrane and its subsequent release into the peroxisomal matrix. After cargo release, the receptors are then recycled back to
the cytosol with the action of the RING complex (PEX2/PEX10/PEX12) and AAA‐type ATPase complex (PEX1/PEX6/PEX26).

conserved N‐terminal cysteines is found to be required for PEROXISOMAL LIPID TRAFFICKING


receptor cycling, while poly‐ubiquitination on lysine residues is
a signal for proteasomal degradation. The RING complex is Besides proteins, peroxisomes also need to exchange lipid mol-
thought to be responsible for receptor ubiquitination. The ubiq- ecules with the surroundings to regulate and balance their pro-
uitinated PEX5 is recognized by the two AAA‐type ATPases liferation and cellular functions. Peroxisomes must acquire
PEX1 and PEX6, which are anchored to the peroxisomal mem- membrane lipids for their growth and proliferation from another
brane via their interaction with the integral membrane protein donor organelle, as they lack most of the enzymes required for
PEX26. ATP hydrolysis by PEX1 and PEX6 provides the driv- phospholipid biosynthesis. Peroxisomes also need to exchange
ing force that is required for the extraction of receptors from the metabolites with other organelles for their cellular functions.
membrane, resulting in the recycling of receptors back into the For example, metabolites are transferred between peroxisome
cytosol [31, 38]. and the mitochondria for β‐oxidation of VLCFAs. However, the
12:  Peroxisome Assembly, Degradation, and Disease 143

precise mechanism by which lipids are exchanged between is not known whether ACBD5 binds to lipid molecules directly
­peroxisomes and other organelles is not fully understood. via its acyl‐CoA‐binding domain, or whether other lipid‐­binding
In general, lipid exchange between various cellular compart- proteins are recruited to these contact sites and help to s­ huttle
ments can occur via either the vesicular transport mechanism or lipids between the two organelles.
nonvesicular mechanisms. The bulk transfer of lipids occurs
when vesicles budding from the donor membrane fuse with the
receiving membrane. In the nonvesicular trafficking pathways, Interplay between peroxisomes
specific lipid molecules are usually transferred at the MCSs, and mitochondria
which are the regions where membranes of two organelles are in Peroxisomes and mitochondria share many common charac-
close proximity (within 10–30 nm) [42]. The implication of teristics and functions. For example, the biogenesis of both
various lipid‐sorting mechanisms for lipid transfer between organelles are under the control of the pro‐lipid metabolism
­peroxisomes and other organelles will be discussed. nuclear receptors PPARs. They also cooperate with each
other in the β‐oxidation of VLCFAs in mammalian cells. Lipid
Interplay between peroxisomes and the ER exchange between the two organelles is suggested to occur at
their MCSs. A tethering complex consisting of Pex11p and the
As discussed earlier, the ER contributes to the de novo biogen-
ERMES protein Mdm34 has been identified in yeast, allowing
esis as well as the growth and division of existing peroxisomes,
for the efficient transfer of metabolites between peroxisomes
as it provides peroxisomes with a subset of PMPs. Moreover, as
and mitochondria [47]. In mammalian cells, one of the pro-
the ER is the major site of lipid synthesis, another likely role of
teins involved in the interaction between peroxisomes and
the ER in peroxisome biogenesis is to provide membrane lipids
mitochondria is identified to be ABCD1. Defects in ABCD1
for the development of peroxisomal endomembrane system.
are associated with the peroxisomal disorder X‐linked adre-
Much of our knowledge about lipid transfer between peroxi-
noleukodystrophy (X‐ALD), which is characterized by the
somes and the ER is gained from studies done in yeast; and
­accumulation of VLCFAs [48].
indeed, a nonvesicular pathway is suggested to be involved in
On the other hand, lipid transfer between peroxisomes and
this process [43]. Although the molecular mechanism underly-
mitochondria may also occur via vesicular transport, as a novel
ing this nonvesicular lipid transport is not known, close contact
vesicular trafficking pathway between them has recently been
(i.e. MCSs) between the two organelles are suggested to be
reported in the mammalian system. Mitochondria‐derived vesi-
required for this transfer.
cles (MDVs) have been shown to fuse with a subpopulation
So far, two protein complexes have been identified in yeast to
(~10% of total) of peroxisomes [49]. However, the function and
allow for the tethering between peroxisomes and the ER. The
physiological significance of these MDVs remains unclear.
first tethering complex was identified in the yeast Pichia
Whether they contribute to peroxisome biogenesis from the
­pastoris, consisting of the integral peroxisomal membrane
mitochondria or they are involved in lipid exchange between
­protein Pex30p and the ER reticulons Rnt1, Rnt2, and Yop1.
the two organelles needs to be further investigated.
The ER–peroxisome contact sites mediated by this complex
were shown to be involved in peroxisome proliferation [44]. In
the budding yeast S. cerevisiae, a tethering complex, consisting Interplay between peroxisomes
of Pex3p and the peroxisome inheritance factor Inp1p, was
identified to play a role in peroxisome inheritance. Inp1p acts as
and lipid droplets
a molecular hinge between the ER‐localized Pex3p and the per- Lipid droplets (LDs) are the lipid‐storage organelles found in all
oxisomal Pex3p, thereby anchoring peroxisomes to the ER in organisms due to their ability to accumulate neutral lipids such
the mother cells. On the other hand, peroxisomes that are as triacylglycerols (TAGs) and cholesterol ester. An imbalance
enriched for Inp2p are transported to the bud along microtu- in lipid storage and lipolysis or secretion will lead to hepatic
bules via the myosin motor protein 2 (Myo2). steatosis, which is a condition of abnormal lipid accumulation
Although no known mammalian homolog of Pex30p or Inp1p in LDs in the liver. This condition can be induced by an increase
has been identified, a novel tethering complex has recently been in fatty acid uptake/synthesis and LD biogenesis; at the same
identified in mammalian cells that is mediated by the direct time, a decrease in LD catabolism (such as β‐oxidation of fatty
interaction between the ER‐resident VAMP‐associated proteins acid) and impaired secretion can also lead to steatosis [50].
A and B (VAPA and VAPB) and the peroxisomal membrane There is growing evidence that rather than solely being a storage
protein ACBD5 [45, 46]. Both VAP proteins were found to con- organelle, LDs also have a multifunctional nature, being
centrate at specific loci on the ER close to peroxisomes, where ­implicated in protein degradation and pathogen replication as
they interact with the FFAT motif‐containing protein ACBD5 well [51].
on peroxisomal membranes to facilitate the formation of ER–­ The close proximity between peroxisomes and LDs has long
peroxisome contact sites. These contact sites bring the two orga- been seen in electron micrographs, and was later confirmed in
nelles close to each other, allowing for the transfer of materials multiple organisms, including the mammalian COS7 cells [52].
between them. Lipid molecules that may be transferred at these The association between these two metabolic organelles is
contact sites include membrane lipids and precursors of choles- thought to be linked to lipolysis in LDs and β‐oxidation of fatty
terol and plasmalogens, as disruption of this VAP–ACBD5 con- acids in peroxisomes. The fact that accumulation of LDs was
tact was shown to interfere with both peroxisome growth and seen in mouse hepatocytes lacking peroxisomes reinforced
cellular cholesterol and plasmalogen homeostasis. However, it a  connection between the two organelles [53]. Although the
144 THE LIVER:  PEROXISOME DEGRADATION: PEXOPHAGY

precise mechanism for peroxisome–LD communication is still involved in β‐oxidation, including acyl‐CoA oxidase (ACOX)
unclear, hemifusion between LD and peroxisomal membranes [59]. The pLon protease may also be involved in the degradation
(i.e. pexopodia) has been proposed to mediate the direct interac- of induced β‐oxidation enzymes upon peroxisomal proliferation
tion between the two organelles in yeast, which helps to facili- [60]. In yeast, the peroxisomal Lon promotes degradation of
tate the transfer of fatty acids from LDs to peroxisome for misfolded or damaged peroxisomal matrix proteins. Interestingly,
β‐oxidation [54]. However, whether a similar mechanism exists in the plant Arabidopsis, peroxisome degradation via autophagy
in other organisms is not known. (i.e. pexophagy) is enhanced in cells that lack Lon [61].
Remarkably, the insertion of UBXD8, a hairpin protein with In the 15‐LOX‐mediated autolysis, the peroxisomal mem-
dual localization to the ER and LD, into the cytoplasmic leaflet brane is permeabilized, leading to the release of its matrix con-
of the ER bilayer was suggested to depend on PEX3 and PEX19. tent. 15‐LOX belongs to the lipoxygenase enzyme family which
The shared protein import machinery between PMPs and LD‐ converts polyunsaturated fatty acids into conjugated hydroper-
destinated hairpin proteins suggests a coordinated relationship oxides, leading to the lipoperoxidation of organellar membranes
between the two organelles in the ER [55]. This mutual control and the synthesis of signaling molecules [62]. As a result of
for both peroxisome and LD biogenesis will allow the cell to membrane lipid peroxidation, the matrix content is released and
tightly regulate energy levels in response to metabolic stimuli, thereby degraded by the cytosolic proteases [63]. 15‐LOX was
thereby maintaining lipid homeostasis in the liver. found to be localized to some, but not all, peroxisomes in rat
hepatocytes, suggesting a potential role of 15‐LOX in pro-
grammed degradation of peroxisomes in these cells [64].
Interplay between peroxisomes and lysosomes Moreover, a novel role of 15‐LOX in regulating autophagy has
For a long time, the interaction between peroxisomes and lys- been proposed, as cells deficient in 12/15‐LOX (a homolog of
osomes were mainly focused on pexophagy, a process by which 15‐LOX) exhibit defects in autophagy [65]. Furthermore, a
peroxisomes are selectively degraded in lysosomes via the LOX‐derived oxidized phospholipid was found to act as an
autophagy pathway. However, a novel lysosome–peroxisome effective substrate for key proteins (such as yeast Atg8 and
contact site (LPMC) has recently been reported in mammalian mammalian LC3) required for autophagy, implying a linkage
cells, and its implication in cholesterol transport was discussed between phospholipid oxidation with the cellular turnover
[56]. Mammalian cells can acquire exogenous cholesterol machinery, autophagy [65].
through receptor‐mediated endocytosis of plasma low‐density The principal machinery for peroxisome degradation is the
lipoprotein (LDL). The resulting free cholesterol emerges in the autophagy‐related process called pexophagy. Like autophagy,
late endosome/lysosome, where it is further transported to vari- pexophagy can occur in two modes: macropexophagy and
ous downstream organelles, including the ER, plasma mem- micropexophagy [66, 67]. During macropexophagy, the peroxi-
brane, and mitochondria, via an unknown mechanism [57]. The some is wrapped around by a double‐membrane vesicle called
newly identified LPMC is shown to be mediated by the interac- the autophagosome, to form a new structure named the pex-
tion between the lysosomal membrane protein synaptotagmin ophagosome. The mature pexophagosome then moves along the
VII (Syt7) and phosphatidylinositol 4,5‐bisphosphate (PI(4,5) microtubules and fuses with a lysosome to degrade the seques-
P2) on the peroxisomal membrane. These LPMCs are proposed tered peroxisome. In contrast, during micropexophagy, the
to serve as a potential mechanism for cholesterol efflux from engulfment of cargo (i.e. peroxisomes) occurs by invagination
lysosomes, as evidenced by the accumulation of cholesterol in of the lysosomal membrane, resulting in the budding of vesicles
cultured fibroblasts from patients with peroxisomal disorders. into the vacuole lumen and subsequent degradation of its
After reaching peroxisomes, it is unclear whether the LDL‐ engulfed content. Macropexophagy occurs in all the eukaryotic
derived cholesterol is further processed in peroxisomes or sub- cells that have been examined, including human, whereas
sequently transported to other cellular compartments. micropexophagy has only been observed in two fungi species,
P. pastoris and Aspergillus nidulans. Therefore, only macropex-
ophagy will be discussed in this section.

PEROXISOME DEGRADATION: Ubiquitination‐mediated peroxisome


PEXOPHAGY recognition for pexophagy
Peroxisome homeostasis in cells is maintained by the balance How pexophagy is triggered or how a peroxisome is recog-
between peroxisome biogenesis (i.e. new peroxisomes are nized and targeted for pexophagy is not fully understood.
formed de novo or via growth and division) and the degradation Recent data suggest that ubiquitination of membrane proteins
of abnormal or dysfunctional peroxisomes. Peroxisomes can be of specific organelles is required for selective autophagy.
turned over via three mechanisms: (i) matrix protein degradation Consistent with this, ubiquitination of PMPs was shown to
by the Lon protease, (ii) 15‐lipoxygenese (15‐LOX)‐mediated induce pexophagy. For example, overexpression of two
autolysis, and (iii) the autophagy pathway (i.e. pexophagy) [58]. PMPs – PMP34 and PEX3 – that are fused with ubiquitin in
Lon proteases are characterized as ATP‐dependent proteases the cytosolic side, dramatically induced pexophagy in mam-
that have been found in eukaryotic mitochondria, chloroplasts, malian cells [68]. Interestingly, expression of a PEX3 mutant
and peroxisomes. The peroxisome‐specific isoform of Lon pro- that is defective in ubiquitination could also induce peroxi-
tein (pLon) in rat liver was shown to play a role in processing some ubiquitination and degradation, implying that ubiquit-
and activating several PTS1‐containing peroxisomal enzymes ination of PEX3 is dispensable for pexophagy [69]. This result
12:  Peroxisome Assembly, Degradation, and Disease 145

also suggests that other unidentified endogenous PMPs are due to either mutation or depletion of its expression, leads to
ubiquitinated upon PEX3 overexpression. increased pexophagy activity and loss of peroxisomes [76].
Another PMP that can be ubiquitinated, thereby triggering Inhibiting autophagy resulted in not only increase in peroxi-
pexophagy, is mammalian PEX5, the soluble receptor for the some number but also increase in peroxisomal functions [76].
PTS1‐containing peroxisomal matrix proteins. During the Together, these findings suggest that the loss of peroxisomes in
import cycle, the PTS1 receptor PEX5 is regulated by different AAA ATPase‐mutant cells may be due to increased degradation
types of ubiquitination [41]. The mono‐ubiquitination of PEX5 of peroxisomes and not the lack of their biogenesis.
on a conserved N‐terminal cysteine is recognized by the AAA‐
type ATPase complex (PEX1/PEX6/PEX26) that extracts PEX5
from the peroxisomal membrane, thereby recycling it back into
Autophagic receptors for pexophagy
the cytosol for further rounds of protein import. However, when Once a peroxisome is activated for degradation by accumulation
the mono‐ubiquitination‐mediated receptor recycling pathway of ubiquitin on the cytosolic side of the membrane, it will be
is blocked, poly‐ubiquitination of PEX5 on lysine residues recognized by a selective autophagy cargo receptor/adaptor
occurs, followed by its extraction from the membranes and [77]. The autophagy adaptor links the substrate (i.e. peroxi-
rapid degradation by the 26S proteasome. A potential role for some) to a nascent autophagosome by binding to both the ubiq-
mono‐ubiquitination of PEX5 as a quality‐control mechanism uitin on the substrate and the autophagy effector LC3 on the
for peroxisome degradation has been proposed in recent studies. autophagosome. At present, the mammalian autophagy recep-
For example, overexpression of an export‐deficient PEX5 tors that have been shown to play a role in pexophagy are p62
mutant was shown to trigger pexophagy in mouse embryonic and NBR1 [10]. These two proteins share similar domain com-
fibroblasts [70]. This PEX5 mutant was generated by fusing a positions, including an N‐terminal PB1 domain that allows for
bulky C‐terminal tag to the PEX5 protein. The bulky C‐terminal interaction with another p62 molecule, a UBA domain at the
tag does not interfere with PEX5 mono‐ubiquitination, but C‐terminus that binds to ubiquitin on the cargo, a ZZ‐type zinc
inhibits its export from the peroxisomal membranes, thereby finger in the middle, and an LIR motif that binds to the autophagy
allowing the mono‐ubiquitinated PEX5 to be recognized by the factor LC3 [58]. NBR1, but not p62, is shown to be necessary
autophagy machinery [70]. Moreover, mono‐ubiquitination of and sufficient for the turnover of endogenous peroxisomes [10].
PEX5 could also be induced under conditions of ROS activation P62 is not required for pexophagy when NBR1 is excess, but it
[71]. It has been shown that in response to ROS activation, works cooperatively with NBR1 to increase the efficiency of
PEX5 is phosphorylated at Ser141 by ataxia telangiectasia‐ NBR1‐mediated pexophagy [10].
mutated (ATM) kinase. This, in turn, promotes PEX5 mono‐ The difference in specificity between these two adaptors may
ubiquitination at K209, allowing it to be recognized by the be explained by the presence of an additional JUBA domain
autophagy adaptor protein p62, thereby inducing pexophagy (a  novel membrane‐interacting, amphipathic α‐helix region
[71]. However, whether ubiquitination of PMPs happens under ­preceding the UBA domain) in NBR1 [10]. The NBR1 mutants
normal physiological conditions is not known. that lack the JUBA domain (NBR1ΔJ and NBR1ΔJΔUBA)
did  not localize to peroxisome, suggesting an essential role
of  the JUBA domain for its proper subcellular localization.
E3 ubiquitin ligase for pexophagy Interestingly, the mutant that lacks only the UBA domain
Peroxisomes have four putative E3 ubiquitin ligases (PEX2, (NBR1ΔUBA) still localized to peroxisomes, but exhibited
PEX10, PEX12, and TRIM37) that are all shown to act on the increased colocalization with other organelles, such as the mito-
ubiquitination of PEX5 for its removal from the peroxisomal chondria. Taken together, these results suggest that while both
membrane during matrix protein import [72, 73]. Of these four the JUBA and UBA domains are required for its specificity for
peroxisomal E3 ubiquitin ligases, only PEX2, which is a RING peroxisomes, the UBA domain acts as a regulator of NBR1
finger E3 ligase, has been shown to be involved in the ubiquitina- binding to peroxisomes.
tion reaction required for pexophagy. PEX2 is usually rapidly
degraded and, thereby, maintained at low levels under normal
conditions. However, this protein is stabilized in stress condi-
tions, for example during amino acid starvation [74]. The stabili- PEROXISOME FUNCTIONS AND
zation of PEX2 protein expression results in a rapid increase in DISORDERS IN THE LIVER
PMP ubiquitination which is a signal for pexophagy.
Fatty acid β‐oxidation
AAA ATPase defect and pexophagy Beta‐oxidation is the process of metabolizing fatty acids to
The most common mutations in PBDs are found in the peroxi- acetyl‐CoA for ATP synthesis. While peroxisomes are the
somal AAA ATPase, which comprises PEX1, PEX6, and PEX26 sole  site of β‐oxidation in plants and yeast, it occurs in both
[75]. Mutations in these peroxins are believed to reduce peroxi- mitochondria and peroxisomes in mammalian cells. The subset
some numbers by affecting peroxisomal biogenesis. However, of lipids that can only be β‐oxidized in peroxisomes include
there is growing evidence suggesting that the AAA ATPase may VLCFAs, long‐chain dicarboxylic acids, leukotrienes, prosta-
also act in peroxisomal quality control. The main function of the glandins, isoprenoid‐derived fatty acids, and pristanic acid
peroxisomal AAA ATPase is to remove ubiquitinated PEX5. As (a  by‐product of α‐oxidation). The four steps involved
such, it has been shown that the loss of AAA ATPase function, in  ­peroxisomal β‐oxidation are similar to those found in
146 THE LIVER:  PEROXISOME FUNCTIONS AND DISORDERS IN THE LIVER

mitochondria; however, the specific enzymes involved are dif- then undergoes sequential reactions catalyzed by the p­ eroxisomal
ferent. The fatty acid substrates are first activated to CoA esters, enzymes phytanoyl‐CoA 2‐hydroxylase (PHYH), 2‐hydroxy-
before being transported across the peroxisomal membrane phytanoyl‐CoA lyase (HACL1), and pristanal dehydrogenase
by  one of the three ABC transporters (ABCD1–3) [78]. (PrDH), resulting in chain shortening of one carbon to produce
Peroxisomal β‐oxidation occurs via a sequential cycle of dehy- a 2‐methyl fatty acid, pristanoyl‐CoA. Defect in α‐oxidation in
drogenation (by acyl‐CoA oxidase 1–3, ACOX), hydration, humans leads to an autosomal recessive neurological disease
dehydrogenation (by multifunctional protein 1 and 2, MFP), called Refsume disease, which is characterized by the accumu-
and thiolytic cleavage (by sterol carrier protein x, SCPx). The lation of phytanic acid in the plasma and tissues including the
resulting chain‐shortened (by 2 carbons) fatty acids either liver [80, 87]. It is caused by mutations in the PHYH gene
enter a new cycle of β‐oxidation or are transported out of the encoding the peroxisomal enzyme involved in the first step of
peroxisome [79, 80]. α‐oxidation, or in PEX7 gene encoding the receptor for PTS2‐
Liver dysfunction was often reported in patients with single containing proteins such as PHYH [88–90]. Compared to other
peroxisomal β‐oxidation enzyme deficiencies; however, the peroxisomal disorders, the onset of Refsume disease occurs
clinical presentation is diverse. For example, patients with later in childhood with a progressive course. The clinical fea-
mutations in ABCD1 (also known as X‐lined adrenoleukod- tures include retinitis pigmentosa, anosmia, deafness, chronic
ystropy, which is the most frequent peroxisomal disease) [81] polyneuropathy, and ichthyosis. Refsum disease is one of the
and SCPx did not develop liver abnormalities [82]. Also, no few peroxisomal disorders for which a therapy has been devel-
liver disease was reported in patients with mild deficiency in oped. The current therapy calls for controlling or eliminating
MFP2 and ACOX1, who survived into adulthood [83, 84]. phytanic acid from the diet [80, 87].
However, patients with severe mutations in MFP2 possess both
hepatic and neuronal abnormalities and died within the first 2
years of life without achieving any developmental milestones.
Bile acid synthesis
Microscopic analysis on liver biopsies and post‐mortem liver In addition to the metabolism of fatty acids, peroxisomal β‐oxi-
showed enlarged peroxisomes that are reduced in number in dation also plays an important role in the synthesis of bile acids,
these patients [85]. Moreover, a single patient with a defect in as it is required for converting the C27 bile acid intermediates to
ABCD3 was reported to present with hepatosplenomegaly and the mature C24 bile acids. Cholesterol is converted into bile
severe liver disease at the age of 1.5 years. This patient was born acids via sequential reactions catalyzed by multiple enzymes
with no neonatal problems and achieved normal developmental that are predominantly expressed in the liver. After the ring
milestones. Microscopic analysis of cultured primary skin fibro- modification of cholesterol followed by oxidation and shorten-
blasts revealed that peroxisomes appeared to be reduced in ing of the sterol side‐chain, the resulting C27 bile acid interme-
number and enlarged in size. The liver disease progressed and diates di‐ and trihydroxycholestanoic acid (DHCA and THCA)
the patient developed decompensated liver cirrhosis associated are first activated to their CoA esters by the very‐long‐chain
with hepatopulmonary syndrome. The patient died shortly after acyl‐CoA synthetase (VLCS) found at both the ER and peroxi-
liver transplantation due to respiratory complications at 4 years somes or the bile acyl‐CoA synthetase (BACS) found exclu-
of age [86]. sively at the ER (and hepatic) [91]. The subsequent uptake of
these CoA esters into peroxisomes is likely via ABCD3, as a
patient with a defect in ABCD3 presented with hepatospleno-
Fatty acid α‐oxidation megaly and severe liver disease with a significant increased
Fatty acids with the presence of a methyl group at the level of peroxisomal C27 bile acid intermediates in plasma [86].
3‐position, including phytanic acid, cannot be metabolized In Abcd3−/− mice, a reduction of C24 bile acids and an accumu-
by the β‐oxidation pathway. Instead, they need to be first lation of C27 bile acid intermediates were also found in liver,
α‐oxidized in peroxisomes to pristanic acid, which then can be bile, and intestine, indicating a role of ABCD3 in transporting
further metabolized via the β‐oxidation pathway in peroxi- C27 bile acids into the peroxisomes [86]. After transporting into
somes and mitochondria. Phytanic acid is found in almost all the peroxisomes, the CoA‐ester form of DHCA and THCA (i.e.
animals, including humans; however, it is not endogenously DHC‐CoA and THC‐CoA) are first converted to chenodeoxy-
synthesized in animals but rather is a constituent of their diet. choloyl‐CoA (CDC‐CoA) and choloyl‐CoA (CA‐CoA), respec-
Phytanic acid can be efficiently absorbed by animals in the tively, via the enzyme α‐methylacyl‐CoA racemase (AMACR)
form of either phytanic acid itself or its precursors, such as [92], followed by the β‐oxidation of these substrates. The CoA‐
phytol. Phytol is widely found in the green leaves of plants and ester of the C27 bile acid intermediate is oxidized by the
trees, and it is linked to the porphytin ring of chlorophyll via branched‐chain acyl‐CoA oxidase (BCOX2) [93], and the sub-
an ester bond. It cannot be digested by humans, as it can only sequent steps of β‐oxidation are catalyzed by MFP2 and SCPx
be liberated from chlorophyll by bacteria found in the rumen [82, 85, 94]. The last step of bile acid synthesis is catalyzed by
of ruminant animals. Therefore, humans obtain phytanic acid the bile acyl‐CoA amino acid N‐acyltransferase (BAAT), which
primarily from the consumption of dairy products and meat conjugates the newly formed primary bile acids to taurine and
originating from ruminant animals [80, 87]. glycine [95, 96]. The conjugated bile acids are then transported
To initiate the α‐oxidation pathway, phytanic acid is first out of hepatic peroxisomes and excreted into the bile.
activated to phytanoyl‐CoA by the acyl‐CoA synthetases
­ Analysis of plasma bile acids from patients with peroxisomal
ACSL1 (found in mitochondria, the ER, and cytosol) and disorders, including PBDs and peroxisomal β‐oxidation disor-
ACSVL1 (found in the ER and peroxisomes). Phytanoyl‐CoA ders, invariably shows increased levels of C27 bile acid. As C27
12:  Peroxisome Assembly, Degradation, and Disease 147

bile acids are only partially conjugated by BAAT and therefore potential treatment for RCDP patients. However, the low trans-
less efficiently excreted in the bile as compared to C24 bile port efficiency through lactation and low incorporation into
acids [96], fat malabsorption is a common clinical feature in brain of the target plasmalogen limit the beneficial effects of
these patients. Interestingly, the extent of peroxisome deficiency this treatment in the mouse studies [106, 107].
has a clear correlation with the extent of the deficiency in bile
acid synthesis, as patients with the less severe forms of PBD
(i.e. infantile Refsum disease and neonatal adrenoleukodystro-
phy) presented less cholestasis and lower levels of bile acid PEROXISOMES IN ANTIVIRAL
­precursors in serum than those with the most severe form (i.e. INNATE IMMUNITY
Zellweger syndrome) [97]. Bile acid abnormalities are sug-
gested to contribute to the liver pathology in patients with Besides playing a central role in various metabolic activities as
­peroxisomal disorders. At physiological concentrations during described above, peroxisomes have also been proposed to
cholestasis, bile acids did not cause apoptosis or necrosis [98]. ­participate in the innate immune response by recent research.
Instead, they are thought to directly activate a signaling pathway Upon infection, microorganisms are detected by the pattern
in hepatocytes that promotes hepatic inflammation during ­recognition receptors (PRRs) that activate multiple signaling
­cholestasis, thereby leading to liver injury [99]. transduction pathways, leading to the activation of nuclear
factor‐κB (NF‐κB), mitogen‐activated protein kinase (MARKs),
and interferon regulatory factors (IRFs). These transcription
Ether phospholipid biosynthesis
factors will then induce the expression of various proinflamma-
Ether phospholipids are a specialized group of phospholipids tory factors, including the cytokines interleukin‐1, TNF, and
with an ether bond at the sn‐1 position of the glycerol backbone. type I and III interferons (IFNs), thereby creating an antiviral
The ether bond can be either a regular ether bond or a vinyl– state within the cell [108].
ether bond that is present in plasmanyl‐phopholipids or plasme- The adaptor protein for the RIG‐I‐like receptor (RLR) family,
nyl‐phospholipids, respectively. The most abundant form of mitochondrial antiviral signaling protein (MAVS), has been
plasmenyl‐phospholipids are plasmalogens which contain an recently identified to be present on both mitochondria and per-
ethanolamine or a choline moiety in the head group of the glyc- oxisomes in mammalian cells, including human hepatocytes
erol backbone. They are enriched in the heart, kidney, lung, and [109]. During viral infection, the peroxisomal MAVS was
skeletal muscle. Especially in the brain, they represent up to shown to trigger an immediate induction of antiviral gene
90% of the phosphatidylethanolamine fraction [100, 101]. The expression in a type I IFN‐independent manner to provide a
vinyl–ether bond at the sn‐1 position and enrichment of polyun- short‐term antiviral effect. In contrast, the mitochondrial MAVS
saturated fatty acids at the sn‐2 position provide plasmalogens activates an interferon‐dependent signaling pathway, but with
with unique features that allow them to function as: (i) mediators a  delayed kinetics, for a sustained protection later during
for maintaining membrane physical bilayer properties; (ii) ­infection [109]. The coordinate between the peroxisomal and
­sacrificial oxidants; and (iii) reservoirs for biologically active mitochondrial MAVS is proposed to occur at a subdomain on
lipid mediators [100]. the ER called mitochondria‐associated membranes (MAMs)
Plasmalogens are synthesized in concert between peroxi- [110]. During viral infection, peroxisomes and mitochondria are
somes and the ER. The peroxisomal matrix enzyme glycerone- found to interact with each other at the MAM, which acts as an
phosphate O‐acyltransferase (GNPAT) initiates the biosynthesis “innate immune synapse” that coordinates the role of peroxiso-
pathway by catalyzing the acylation of dihydroxyacetine phos- mal and mitochondrial MAVS.
phate (DHAP) at the sn‐1 position, followed by the exchange of
acyl group for an alkyl group by alkylglycerone phosphate syn-
thase (AGPS). Further modifications are completed in the ER,
resulting in the formation of mature plasmalogens [100]. Defect PEROXISOMES IN CANCER
in plasmalogen synthesis leads to the peroxisomal disorder rhi-
zomelic chondrodysplasia punctate (RCDP), which impairs the So far, not much is known about the role of peroxisomes in
normal development of multiple organs of the body, including human cancer development. Reduced peroxisome abundance
bone, brain, lens, lung, kidney, and heart [100, 101]. Depending has been observed in multiple cancer cells, including hepatocel-
on which gene is mutated, there are three forms of RCDP: lular carcinoma [111], colon carcinoma [112], breast cancer
RCDP type 1 (mutation in PEX7, encoding the receptor for [113], and renal cell carcinoma [114]. However, the mechanism
PTS2‐containing proteins, such as APGS) [102], RCDP type 2 that leads to the loss of peroxisomes in these cancer cells is still
(mutation in GNPAT, encoding the enzyme involved in the first unclear. Moreover, the protein levels of peroxisomal enzymes
step of the plasmalogen biosynthesis) [103], and RCDP type 3 involved in branched‐chain fatty acid β‐oxidation, α‐methyla-
(mutation in APGS, encoding the enzyme involved in the second cyl‐CoA racemase (AMACR) and peroxisomal multifunctional
step of the plasmalogen biosynthesis) [104]. Interestingly, protein 2, are found to be upregulated in human prostate cancer
reduced levels of plasmalogens have also been reported in (PCa) [115]. AMACR, one of the established biomarkers in
nonperoxisomal disorders, such as Alzheimer’s disease.
­ PCa, is also shown to be essential for optimal proliferation of
However, whether the reduction in plasmalogens is the cause of some PCa cell lines in vitro [116]. Monocarboxylate transporter
these diseases or a downstream effect needs to be further 2 (MCT2), a putative biomarker for PCa, has recently been
­investigated [105]. Plasmalogen replacement therapy may be a shown to localize mainly at peroxisomes in PCa cells. MCT2
148 THE LIVER:  REFERENCES

expression was upregulated significantly from nonmalignant to 24. South, S.T., Sacksteder, K.A., Li, X., Liu, Y., and Gould, S.J. Inhibitors of
malignant cells, and this increase in protein expression is COPI and COPII do not block PEX3‐mediated peroxisome synthesis. J Cell
Biol, 2000;149(7):1345–60.
directly correlated with its peroxisomal localization, implying a 25. Latruffe, N. and Vamecq, J. Peroxisome proliferators and peroxisome prolif-
possible role of peroxisomes in prostate malignant transforma- erator activated receptors (PPARs) as regulators of lipid metabolism.
tion [117]. However, it remains unclear whether peroxisome Biochimie, 1997;79(2–3):81–94.
abundance is also increased in PCa. 26. Shimizu, M., Takeshita, A., Tsukamoto, T., Gonzalez, F.J., and Osumi, T.
Tissue‐selective, bidirectional regulation of PEX11 alpha and perilipin genes
through a common peroxisome proliferator response element. Mol Cell Biol,
2004;24(3):1313–23.
27. Muntau, A.C., Mayerhofer, P.U., Paton, B.C., Kammerer, S., and Roscher,
REFERENCES A.A. Defective peroxisome membrane synthesis due to mutations in human
PEX3 causes Zellweger syndrome, complementation group G. Am J Hum
1. Rhodin, J. Correlation of ultrastructural organization and function in normal Genet, 2000;67(4):967–75.
and experimentally changed proximal convoluted tubule cells of the mouse 28. Jones, J.M., Morrell, J.C., and Gould, S.J. PEX19 is a predominantly
kidney. Thesis, Aktiebolaget Godvil Stockholm, Karolinska Institute, 1954. ­cytosolic chaperone and import receptor for class 1 peroxisomal membrane
2. De Duve, C. and Baudhuin, P. Peroxisomes (microbodies and related parti- proteins. J Cell Biol, 2004;164(1):57–67.
cles). Physiol Rev, 1966;46(2):323–57. 29. McNew, J.A. and Goodman, J.M. An oligomeric protein is imported into
3. Steinberg, S.J., Dodt, G., Raymond, G.V. et al. Peroxisome biogenesis disor- peroxisomes in vivo. J Cell Biol, 1994;127(5):1245–57.
ders. Biochim Biophys Acta, 2006;1763(12):1733–48. 30. Walton, P.A., Hill, P.E., and Subramani, S. Import of stably folded proteins
4. Erdmann, R., Veenhuis, M., Mertens, D., and Kunau WH. Isolation of per- into peroxisomes. Mol Biol Cell, 1995;6(6):675–83.
oxisome‐deficient mutants of Saccharomyces cerevisiae. Proc Natl Acad Sci 31. Platta, H.W. and Erdmann, R. Peroxisomal dynamics. Trends Cell Biol,
U S A, 1989;86(14):5419–23. 2007;17(10):474–84.
5. Tsukamoto, T., Yokota, S., and Fujiki Y. Isolation and characterization of 32. Brocard, C. and Hartig, A. Peroxisome targeting signal 1: is it really a simple
Chinese hamster ovary cell mutants defective in assembly of peroxisomes. J tripeptide? Biochim Biophys Acta, 2006;1763(12):1565–73.
Cell Biol, 1990;110(3):651–60. 33. Stanley, W.A., Filipp, F.V., Kursula, P. et al. Recognition of a functional per-
6. Fujiki, Y., Okumoto, K., Mukai, S., Honsho, M., and Tamura, S. Peroxisome oxisome type 1 target by the dynamic import receptor pex5p. Mol Cell,
biogenesis in mammalian cells. Front Physiol, 2014;5:307. 2006;24(5):653–63.
7. Prinz, W.A. Bridging the gap: membrane contact sites in signaling, metabo- 34. Lazarow, P.B. The import receptor Pex7p and the PTS2 targeting sequence.
lism, and organelle dynamics. J Cell Biol, 2014;205(6):759–69. Biochim Biophys Acta, 2006;1763(12):1599–604.
8. Shai, N., Schuldiner, M., and Zalckvar, E. No peroxisome is an island  – 35. Schliebs, W. and Kunau, W.H. PTS2 co‐receptors: diverse proteins with
Peroxisome contact sites. Biochim Biophys Acta, 2016;1863(5):1061–9. common features. Biochim Biophys Acta, 2006;1763(12):1605–12.
9. Cho, D.H., Kim, Y.S., Jo, D.S., Choe, S.K., and Jo, E.K. Pexophagy: molecu- 36. van der Klei, I.J. and Veenhuis, M. PTS1‐independent sorting of peroxisomal
lar mechanisms and implications for health and diseases. Mol Cells, matrix proteins by Pex5p. Biochim Biophys Acta, 2006;1763(12):
2018;41(1):55–64. 1794–800.
10. Deosaran, E., Larsen, K.B., Hua, R. et al. NBR1 acts as an autophagy recep- 37. Williams, C. and Distel, B. Pex13p: docking or cargo handling protein?
tor for peroxisomes. J Cell Sci, 2013;126(Pt 4):939–52. Biochim Biophys Acta, 2006;1763(12):1585–91.
11. Lazarow, P.B. and Fujiki Y. Biogenesis of peroxisomes. Annu Rev Cell Biol, 38. Meinecke, M., Bartsch, P., and Wagner, R. Peroxisomal protein import pores.
1985;1:489–530. Biochim Biophys Acta, 2016;1863(5):821–7.
12. Koch, J. and Brocard, C. PEX11 proteins attract Mff and human Fis1 to 39. Meinecke, M., Cizmowski, C., Schliebs, W. et al. The peroxisomal importomer
coordinate peroxisomal fission. J Cell Sci, 2012;125(Pt 16):3813–26. constitutes a large and highly dynamic pore. Nat Cell Biol, 2010;12(3):273–7.
13. Delille, H.K., Dodt, G., and Schrader M. Pex11pbeta‐mediated maturation of 40. Freitas, M.O., Francisco, T., Rodrigues, T.A. et al. PEX5 protein binds mon-
peroxisomes. Commun Integr Biol, 2011;4(1):51–4. omeric catalase blocking its tetramerization and releases it upon binding the
14. Honsho, M., Tamura, S., Shimozawa, N. et al. Mutation in PEX16 is causal N‐terminal domain of PEX14. J Biol Chem, 2011;286(47):40509–19.
in the peroxisome‐deficient Zellweger syndrome of complementation group 41. Platta, H.W., Hagen, S., Reidick, C., and Erdmann, R. The peroxisomal
D. Am J Hum Genet, 1998;63(6):1622–30. receptor dislocation pathway: to the exportomer and beyond. Biochimie,
15. Matsuzono, Y., Kinoshita, N., Tamura, S. et al. Human PEX19: cDNA clon- 2014;98:16–28.
ing by functional complementation, mutation analysis in a patient with 42. Helle, S.C., Kanfer, G., Kolar, K. et al. Organization and function of mem-
Zellweger syndrome, and potential role in peroxisomal membrane assembly. brane contact sites. Biochim Biophys Acta, 2013;1833(11):2526–41.
Proc Natl Acad Sci U S A, 1999;96(5):2116–21. 43. Raychaudhuri, S. and Prinz, W.A. Nonvesicular phospholipid transfer
16. Ghaedi, K., Tamura, S., Okumoto, K., Matsuzono, Y., and Fujiki, Y. The per- between peroxisomes and the endoplasmic reticulum. Proc Natl Acad Sci U
oxin pex3p initiates membrane assembly in peroxisome biogenesis. Mol Biol S A, 2008;105(41):15785–90.
Cell, 2000;11(6):2085–102. 44. David, C., Koch, J., Oeljeklaus, S. et al. A combined approach of quantita-
17. Titorenko, V.I., Chan, H., and Rachubinski RA. Fusion of small peroxisomal tive interaction proteomics and live‐cell imaging reveals a regulatory role for
vesicles in vitro reconstructs an early step in the in vivo multistep peroxisome endoplasmic reticulum (ER) reticulon homology proteins in peroxisome bio-
assembly pathway of Yarrowia lipolytica. J Cell Biol, 2000;148(1):29–44. genesis. Mol Cell Proteomics, 2013;12(9):2408–25.
18. van der Zand, A., Gent, J., Braakman, I., and Tabak, H.F. Biochemically 45. Hua, R., Cheng, D., Coyaud, E. et al. VAPs and ACBD5 tether peroxisomes
distinct vesicles from the endoplasmic reticulum fuse to form peroxisomes. to the ER for peroxisome maintenance and lipid homeostasis. J Cell Biol,
Cell, 2012;149(2):397–409. 2017;216(2):367–77.
19. Knoops, K., Manivannan, S., Cepinska, M.N. et al. Preperoxisomal vesicles 46. Costello, J.L., Castro, I.G., Hacker, C. et  al. ACBD5 and VAPB mediate
can form in the absence of Pex3. J Cell Biol, 2014;204(5):659–68. membrane associations between peroxisomes and the ER. J Cell Biol,
20. Geuze, H.J., Murk, J.L., Stroobants, A.K. et al. Involvement of the endoplas- 2017;216(2):331–42.
mic reticulum in peroxisome formation. Mol Biol Cell, 2003;14(7):2900–7. 47. Mattiazzi Usaj, M., Brloznik, M., Kaferle, P. et al. Genome‐wide localiza-
21. Hashiguchi, N., Kojidani, T., Imanaka, T. et al. Peroxisomes are formed from tion study of yeast pex11 identifies peroxisome‐mitochondria interactions
complex membrane structures in PEX6‐deficient CHO cells upon genetic through the ERMES complex. J Mol Biol, 2015;427(11):2072–87.
complementation. Mol Biol Cell, 2002;13(2):711–22. 48. McGuinness, M.C., Lu, J.F., Zhang, H.P. et  al. Role of ALDP (ABCD1)
22. Baes, M., Gressens, P., Baumgart, E. et  al. A mouse model for Zellweger and  mitochondria in X‐linked adrenoleukodystrophy. Mol Cell Biol,
syndrome. Nat Genet, 1997;17(1):49–57. 2003;23(2):744–53.
23. Hua, R., Gidda, S.K., Aranovich, A., Mullen, R.T., and Kim PK. Multiple 49. Neuspiel, M., Schauss, A.C., Braschi, E. et  al. Cargo‐selected transport
domains in PEX16 mediate its trafficking and recruitment of peroxisomal from  the mitochondria to peroxisomes is mediated by vesicular carriers.
proteins to the ER. Traffic, 2015;16(8):832–52. Curr Biol, 2008;18(2):102–8.
12:  Peroxisome Assembly, Degradation, and Disease 149

50. Gluchowski, N.L., Becuwe, M., Walther, T.C., and Farese, R.V., Jr. Lipid 77. Johansen, T. and Lamark, T. Selective autophagy mediated by autophagic
droplets and liver disease: from basic biology to clinical implications. adapter proteins. Autophagy, 2011;7(3):279–96.
Nat Rev Gastroenterol Hepatol, 2017;14(6):343–55. 78. Morita, M. and Imanaka, T. Peroxisomal ABC transporters: structure, func-
51. Beller, M., Thiel, K., Thul, P.J., and Jackle, H. Lipid droplets: a dynamic tion and role in disease. Biochim Biophys Acta, 2012;1822(9):1387–96.
organelle moves into focus. FEBS Lett, 2010;584(11):2176–82. 79. Baes, M. and Van Veldhoven, P.P. Hepatic dysfunction in peroxisomal
52. Schrader, M. Tubulo‐reticular clusters of peroxisomes in living COS‐7 cells: ­disorders. Biochim Biophys Acta, 2016;1863(5):956–70.
dynamic behavior and association with lipid droplets. J Histochem Cytochem, 80. Waterham, H.R., Ferdinandusse, S., and Wanders, R.J. Human disorders
2001;49(11):1421–29. of  peroxisome metabolism and biogenesis. Biochim Biophys Acta,
53. Dirkx, R., Vanhorebeek, I., Martens, K. et  al. Absence of peroxisomes in 2016;1863(5):922–33.
mouse hepatocytes causes mitochondrial and ER abnormalities. Hepatology, 81. Kemp, S., Berger, J., and Aubourg, P. X‐linked adrenoleukodystrophy:
2005;41(4):868–78. clinical, metabolic, genetic and pathophysiological aspects. Biochim
54. Binns, D., Januszewski, T., Chen, Y. et al. An intimate collaboration between Biophys Acta, 2012;1822(9):1465–74.
peroxisomes and lipid bodies. J Cell Biol, 2006;173(5):719–31. 82. Ferdinandusse, S., Kostopoulos, P., Denis, S. et al. Mutations in the gene
55. Schrul, B. and Kopito RR. Peroxin‐dependent targeting of a lipid‐droplet‐ encoding peroxisomal sterol carrier protein X (SCPx) cause leukencepha-
destined membrane protein to ER subdomains. Nat Cell Biol, lopathy with dystonia and motor neuropathy. Am J Hum Genet,
2016;18(7):740–51. 2006;78(6):1046–52.
56. Chu, B.B., Liao, Y.C., Qi, W. et al. Cholesterol transport through lysosome‐ 83. Lines, M.A., Jobling, R., Brady, L. et al. Peroxisomal D‐bifunctional pro-
peroxisome membrane contacts. Cell, 2015;161(2):291–306. tein deficiency: three adults diagnosed by whole‐exome sequencing.
57. Chang, T.Y., Chang, C.C., Ohgami, N., and Yamauchi Y. Cholesterol sensing, Neurology, 2014;82(11):963–8.
trafficking, and esterification. Annu Rev Cell Dev Biol, 2006;22:129–57. 84. Ferdinandusse, S., Barker, S., Lachlan, K. et al. Adult peroxisomal acyl‐
58. Zientara‐Rytter, K. and Subramani S. Autophagic degradation of peroxi- coenzyme A oxidase deficiency with cerebellar and brainstem atrophy. J
somes in mammals. Biochem Soc Trans, 2016;44(2):431–40. Neurol Neurosurg Psychiatry, 2010;81(3):310–12.
59. Omi, S., Nakata, R., Okamura‐Ikeda, K., Konishi, H., and Taniguchi H. 85. Ferdinandusse, S., Denis, S., Mooyer, P.A. et al. Clinical and biochemical spec-
Contribution of peroxisome‐specific isoform of Lon protease in sorting trum of D‐bifunctional protein deficiency. Ann Neurol, 2006;59(1):92–104.
PTS1 proteins to peroxisomes. J Biochem, 2008;143(5):649–60. 86. Ferdinandusse, S., Jimenez‐Sanchez, G., Koster, J. et al. A novel bile acid
60. Yokota, S., Haraguchi, C.M., and Oda T. Induction of peroxisomal Lon pro- biosynthesis defect due to a deficiency of peroxisomal ABCD3. Hum Mol
tease in rat liver after di‐(2‐ethylhexyl)phthalate treatment. Histochem Cell Genet, 2015;24(2):361–70.
Biol, 2008;129(1):73–83. 87. Wanders, R.J., Komen, J., and Ferdinandusse, S. Phytanic acid metabolism
61. Bartel, B., Farmer, L.M., Rinaldi, M.A. et al. Mutation of the Arabidopsis in health and disease. Biochim Biophys Acta, 2011;1811(9):498–507.
LON2 peroxisomal protease enhances pexophagy. Autophagy, 88. Jansen, G.A., Waterham, H.R., and Wanders, R.J. Molecular basis of
2014;10(3):518–19. Refsum disease: sequence variations in phytanoyl‐CoA hydroxylase
62. Maccarrone, M., Melino, G., and Finazzi‐Agro, A. Lipoxygenases and their (PHYH) and the PTS2 receptor (PEX7). Hum Mutat, 2004;23(3):209–18.
involvement in programmed cell death. Cell Death Differ, 2001;8(8):776–84. 89. Mihalik, S.J., Morrell, J.C., Kim, D. et  al. Identification of PAHX, a
63. van Leyen, K., Duvoisin, R.M., Engelhardt, H., and Wiedmann M. A func- Refsum disease gene. Nat Genet, 1997;17(2):185–9.
tion for lipoxygenase in programmed organelle degradation. Nature, 90. van den Brink, D.M., Brites, P., Haasjes, J. et al. Identification of PEX7 as
1998;395(6700):392–5. the second gene involved in Refsum disease. Am J Hum Genet,
64. Yokota, S., Oda, T., and Fahimi, H.D. The role of 15‐lipoxygenase in disrup- 2003;72(2):471–7.
tion of the peroxisomal membrane and in programmed degradation of per- 91. Mihalik, S.J., Steinberg, S.J., Pei, Z. et al. Participation of two members of
oxisomes in normal rat liver. J Histochem Cytochem, 2001;49(5):613–22. the very long‐chain acyl‐CoA synthetase family in bile acid synthesis and
65. Morgan, A.H., Hammond, V.J., Sakoh‐Nakatogawa, M. et al. A novel role recycling. J Biol Chem, 2002;277(27):24771–9.
for 12/15‐lipoxygenase in regulating autophagy. Redox Biol, 2015;4:40–7. 92. Schmitz, W., Albers, C., Fingerhut, R., and Conzelmann, E. Purification
66. Feng, Y., He, D., Yao, Z., and Klionsky, D.J. The machinery of macroau- and characterization of an alpha‐methylacyl‐CoA racemase from human
tophagy. Cell Res, 2014;24(1):24–41. liver. Eur J Biochem, 1995;231(3):815–22.
67. Gatica, D., Lahiri, V., and Klionsky, D.J. Cargo recognition and degradation 93. Vanhove, G.F., Van Veldhoven, P.P., Fransen, M. et al. The CoA esters of
by selective autophagy. Nat Cell Biol, 2018;20(3):233–42. 2‐methyl‐branched chain fatty acids and of the bile acid intermediates di‐
68. Kim, P.K., Hailey, D.W., Mullen, R.T., and Lippincott‐Schwartz, J. Ubiquitin and trihydroxycoprostanic acids are oxidized by one single peroxisomal
signals autophagic degradation of cytosolic proteins and peroxisomes. Proc branched chain acyl‐CoA oxidase in human liver and kidney. J Biol Chem,
Natl Acad Sci U S A, 2008;105(52):20567–74. 1993;268(14):10335–44.
69. Yamashita, S., Abe, K., Tatemichi, Y., and Fujiki Y. The membrane peroxin 94. Novikov, D., Dieuaide‐Noubhani, M., Vermeesch, J.R. et  al. The human
PEX3 induces peroxisome‐ubiquitination‐linked pexophagy. Autophagy, peroxisomal multifunctional protein involved in bile acid synthesis: activity
2014;10(9):1549–64. measurement, deficiency in Zellweger syndrome and chromosome map-
70. Nordgren, M., Francisco, T., Lismont, C. et al. Export‐deficient monoubiq- ping. Biochim Biophys Acta, 1997;1360(3):229–40.
uitinated PEX5 triggers peroxisome removal in SV40 large T antigen‐­ 95. He, D., Barnes, S., and Falany, C.N. Rat liver bile acid CoA:amino acid
transformed mouse embryonic fibroblasts. Autophagy, 2015;11(8):1326–40. N‐acyltransferase: expression, characterization, and peroxisomal localiza-
71. Zhang, J., Tripathi, D.N., Jing, J. et  al. ATM functions at the peroxisome tion. J Lipid Res, 2003;44(12):2242–9.
to  induce pexophagy in response to ROS. Nat Cell Biol, 2015; 96. Solaas, K., Ulvestad, A., Soreide, O., and Kase BF. Subcellular organiza-
17(10):1259–69. tion of bile acid amidation in human liver: a key issue in regulating the
72. Okumoto, K., Noda, H., and Fujiki, Y. Distinct modes of ubiquitination of biosynthesis of bile salts. J Lipid Res, 2000;41(7):1154–62.
peroxisome‐targeting signal type 1 (PTS1) receptor Pex5p regulate PTS1 97. Van Eldere, J.R., Parmentier, G.G., Eyssen, H.J. et al. Bile acids in peroxi-
protein import. J Biol Chem, 2014;289(20):14089–108. somal disorders. Eur J Clin Invest, 1987;17(5):386–90.
73. Wang, W., Xia, Z.J., Farre, J.C., and Subramani, S. TRIM37, a novel E3 98. Zhang, Y., Hong, J.Y., Rockwell, C.E. et al. Effect of bile duct ligation on
ligase for PEX5‐mediated peroxisomal matrix protein import. J Cell Biol, bile acid composition in mouse serum and liver. Liver Int,
2017;216(9):2843–58. 2012;32(1):58–69.
74. Sargent, G., van Zutphen, T., Shatseva, T. et  al. PEX2 is the E3 ubiquitin 99. Allen, K., Jaeschke, H., and Copple BL. Bile acids induce inflammatory
ligase required for pexophagy during starvation. J Cell Biol, genes in hepatocytes: a novel mechanism of inflammation during obstruc-
2016;214(6):677–90. tive cholestasis. Am J Pathol, 2011;178(1):175–86.
75. Nazarko, T.Y. Pexophagy is responsible for 65% of cases of peroxisome bio- 100. Braverman, N.E. and Moser, A.B. Functions of plasmalogen lipids in health
genesis disorders. Autophagy, 2017;13(5):991–4. and disease. Biochim Biophys Acta, 2012;1822(9):1442–52.
76. Law, K.B., Bronte‐Tinkew, D., Di Pietro, E. et  al. The peroxisomal AAA 101. da Silva, T.F., Sousa, V.F., Malheiro, A.R., and Brites, P. The importance of
ATPase complex prevents pexophagy and development of peroxisome bio- ether‐phospholipids: a view from the perspective of mouse models. Biochim
genesis disorders. Autophagy, 2017;13(5):868–84. Biophys Acta, 2012;1822(9):1501–8.
150 THE LIVER:  REFERENCES

102. Purdue, P.E., Skoneczny, M., Yang, X., Zhang, J.W., and Lazarow, P.B. 110. Horner, S.M., Liu, H.M., Park, H.S., Briley, J., and Gale, M., Jr.
Rhizomelic chondrodysplasia punctata, a peroxisomal biogenesis disorder Mitochondrial‐associated endoplasmic reticulum membranes (MAM) form
caused by defects in Pex7p, a peroxisomal protein import receptor: a mini- innate immune synapses and are targeted by hepatitis C virus. Proc Natl
review. Neurochem Res, 1999;24(4):581–6. Acad Sci U S A, 2011;108(35):14590–5.
103. Ofman, R., Hettema, E.H., Hogenhout, E.M. et  al. Acyl‐CoA: 111. Litwin, J.A., Beier, K., Volkl, A., Hofmann, W.J., and Fahimi, H.D.
dihydroxyacetonephosphate acyltransferase: cloning of the human cDNA Immunocytochemical investigation of catalase and peroxisomal lipid beta‐
and resolution of the molecular basis in rhizomelic chondrodysplasia punc- oxidation enzymes in human hepatocellular tumors and liver cirrhosis.
tata type 2. Hum Mol Genet, 1998;7(5):847–53. Virchows Arch, 1999;435(5):486–95.
104. de Vet, E.C., Ijlst, L., Oostheim, W., Wanders, R.J., and van den Bosch, H. 112. Lauer, C., Volkl, A., Riedl, S., Fahimi, H.D., and Beier, K. Impairment of
Alkyl‐dihydroxyacetonephosphate synthase. Fate in peroxisome biogene- peroxisomal biogenesis in human colon carcinoma. Carcinogenesis,
sis disorders and identification of the point mutation underlying a single 1999;20(6):985–9.
enzyme deficiency. J Biol Chem, 1998;273(17):10296–301. 113. el Bouhtoury, F., Keller, J.M., Colin, S., Parache, R.M., and Dauca M.
105. Han, X., Holtzman, D.M., and McKeel, D.W., Jr. Plasmalogen deficiency in Peroxisomal enzymes in normal and tumoral human breast. J Pathol,
early Alzheimer’s disease subjects and in animal models: molecular charac- 1992;166(1):27–35.
terization using electrospray ionization mass spectrometry. J Neurochem, 114. Frederiks, W.M., Bosch, K.S., Hoeben, K.A., van Marle, J., and Langbein,
2001;77(4):1168–80. S. Renal cell carcinoma and oxidative stress: the lack of peroxisomes. Acta
106. Brites, P., Ferreira, A.S., da Silva, T.F. et al. Alkyl‐glycerol rescues plasm- Histochem, 2010;112(4):364–71.
alogen levels and pathology of ether‐phospholipid deficient mice. PLoS 115. Zha, S., Ferdinandusse, S., Hicks, J.L. et al. Peroxisomal branched chain
One, 2011;6(12):e28539. fatty acid beta‐oxidation pathway is upregulated in prostate cancer.
107. Das, A.K., Holmes, R.D., Wilson, G.N., and Hajra, A.K. Dietary ether lipid Prostate, 2005;63(4):316–23.
incorporation into tissue plasmalogens of humans and rodents. Lipids, 116. Zha, S., Ferdinandusse, S., Denis, S. et  al. Alpha‐methylacyl‐CoA race-
1992;27(6):401–5. mase as an androgen‐independent growth modifier in prostate cancer.
108. Odendall, C. and Kagan, J.C. Peroxisomes and the antiviral responses of Cancer Res, 2003;63(21):7365–76.
mammalian cells. Subcell Biochem, 2013;69:67–75. 117. Valenca, I., Pertega‐Gomes, N., Vizcaino, J.R. et al. Localization of MCT2
109. Dixit, E., Boulant, S., Zhang, Y. et al. Peroxisomes are signaling platforms at peroxisomes is associated with malignant transformation in prostate can-
for antiviral innate immunity. Cell, 2010;141(4):668–81. cer. J Cell Mol Med, 2015;19(4):723–33.
Organelle–Organelle
13 Contacts: Origins and
Functions
Uri Manor
Waitt Advanced Biophotonics Center, Salk Institute for Biological Studies, La Jolla, CA, USA

INTRODUCTION implications of this evolutionary history and the insight this pro-
vides towards the mechanisms for organelle–organelle interac-
Eons ago, before the rise of the eukaryotic cell, single‐celled tions in physiology and pathophysiology.
protobacteria and archaeal cells dominated the biosphere. The
revolutionary theory of endosymbiosis, first posited by Lynn
Margulis, dictates that what we today call mitochondria were
actually prokaryotic cells (“protomitochondria”) that were THE ORIGIN STORY OF THE ER, THE
engulfed by a larger prokaryotic host cell in a symbiotic rela- CHIEF OPERATING OFFICER OF
tionship [1]. Similarly, the nucleus has been proposed to be the THE NUCLEUS
evolutionary progeny of an ancient endosymbiont that was
engulfed by our ancestral host cell. The precise ordering or As briefly mentioned above, until recently, all endosymbiotic
mechanism of these ancient endosymbiotic marriages is unclear: origin stories of the modern eukaryotic cell traditionally cast
Was it first mitochondria or the nucleus? How did the archaeal the nucleus and mitochondria as the descendants of smaller
cell engulf other cells long before the evolution of phago/endo- ancestral cells engulfed by a larger ancestral cell. This could be
cytotic machinery? Nonetheless, these critical events are now described as an “outside‐in” model, in which endosymbionts
accepted to be a prerequisite and the basis for the evolution of are engulfed by and evolved within the host cell. A recent study
multicellular life and all its complexities. It would be difficult to strongly supports a conceptual inversion of this model to an
overstate the impact endosymbiosis had on life and, indeed, the “inside‐out” model, wherein “ectosymbiotic” protomitochon-
planet Earth. Today, all eukaryotic life is both shaped and con- dria closely interacted and exchanged materials with membrane
strained by these early evolutionary events in ways that have protrusions on the surface of a larger archaeal host cell [2]. For
only recently started to become apparent. For this chapter, I will many generations these two cell types continually exchanged
first explain how recent work shows that early endosymbiosis as materials in mutually beneficial transactions that reinforced
traditionally understood requires a simple but profound modifi- symbiotic coevolution. That archaeal host cell is homologous
cation; it turns out that “ectosymbiosis” provides a much more to what we now call the nucleus (aka “protonucleus”), with the
feasible and parsimonious model [2, 3]. Then I will explain how bleb extrusion sites representing what are now nuclear pore
this novel model for the evolution of the modern cell sets the complexes. The ancestral archaeal cell already evolved proteo-
stage for everything we currently know about organelle biology, lytic, N‐glycosylation, membrane scission, and cytoskeletal
in particular how interactions between the endoplasmic reticu- machineries necessary for the cargo trafficking and modifica-
lum (ER) and the rest of our membrane‐bound organelles (i.e. tion that exemplifies the Golgi and lysosomal compartments
mitochondria, lysosomes, peroxisomes, endosomes, and the [2]. The protonucleus and protomitochondria continued to
Golgi) reflect our ancestral archaea’s evolution into a modern interact and the protonucleus’s blebs eventually expanded and
eukaryotic cell. With this context, we will then explore the fused around the protomitochondria until they all formed an

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
152 THE LIVER:  THE ORIGIN STORY OF THE ER, THE CHIEF OPERATING OFFICER OF THE NUCLEUS

interconnected web that became what we now call the ER, This model is appealing in both its simplicity and its parsi-
mitochondria, cytoplasm, endomembrane system, and the mony. It is a more plausible model, in that it no longer requires
plasma membrane (Figure  13.1) [2]. In short, the modern the highly complicated, unlikely, and energetically expensive
eukaryotic cell. engulfment and subsequent endosymbiosis of a separate cell

(a) (b)
Epibiotic bacterium
(future mitochondrion)

Eocyte
(future nucleus)

(c) (d)

(e) (f)

original weakened half full nuclear


S-layer S-layer pore pore

Figure 13.1  Inside‐out model for the evolution of eukaryotic cell organization. Model showing the stepwise evolution of eukaryotic cell organiza-
tion from (a) an eocyte ancestor with a single bounding membrane and a glycoprotein rich cell wall (S‐layer) interacting with epibiotic α‐proteo-
bacteria (protomitochondria). (b) The eocyte cell forms protrusions, aided by protein–membrane interactions at the protrusion neck. These
protrusions facilitated material exchange with protomitochondria. (c) Selection for a greater area of contact between the symbionts would have led
to bleb enlargement and the eventual loss of the S‐layer from the protrusions. (d) Blebs would have then been further stabilized by the development
of a symmetric nuclear pore outer ring complex (Figure 13.2) and through the establishment of LINC complexes that, following the gradual loss of
the S‐layer, physically connected the original cell body (the nascent nuclear compartment) to the inner bleb membranes. (e) With the expansion of
blebs to enclose the protomitochondria, a process that would have facilitated the acquisition of bacterial lipid biosynthesis machinery by the host,
the site of cell growth would have progressively shifted to the cytoplasm, facilitated by the development of regulated traffic through the nuclear
pore. At the same time, the spaces between blebs would have enabled the gradual maturation of proteins secreted into the environment via the peri-
nuclear space through glycosylation and proteolytic cleavage. (f) Finally, bleb fusion would have connected cytoplasmic compartments and driven
the formation of an intact plasma membrane, perhaps through a process akin to phagocytosis whereby one bleb enveloped the whole. This simple
topological transition would have isolated the endoplasmic reticulum from the outside world, driven the full development of a system of vesicular
trafficking, and established strict vertical transmission of mitochondria, leading to a cell with modern eukaryotic cell organization. Reproduced
from Baum 2014 [2], https://bmcbiol.biomedcentral.com/articles/10.1186/s12915‐014‐0076‐2. Licensed under CCBY 4.0.
13:  Organelle–Organelle Contacts: Origins and Functions 153

type for the protonucleus. Instead, the host cell is the nucleus, Thus, it is not terribly surprising that the ER interacts with all of
meaning the only symbiotic partner left to account for is the the other organelles in the cell. Nonetheless, the apposition of
mitochondria. The inside‐out model of the coevolution of the the ER with other organelles ultimately represents an opportu-
host cell and protomitochondria also makes more sense from an nity for these organelle pairs to synergistically carry out novel
evolutionary perspective: The exchange of materials with ectos- functions that would otherwise be impossible. With this frame-
ymbiotic protomitochondria across generations could facilitate work in mind, let us explore the physiological roles of these
incremental advantages on evolutionary time‐scales necessary ER–organelle contacts.
for natural selection to develop a strong symbiotic relationship.
Most relevant to this chapter and textbook, the inside‐out model
is parsimonious because it contextualizes so much of modern
cell and organelle biology. THE ER AND MITOCHONDRIA:
Conceptualizing the ancestral ER as an outgrowth of the host AN ANCIENT MARRIAGE
cell’s probing interactions with its environment, it is not difficult
to see how it became the chief operating officer of our ancestral Perhaps not coincidentally, the most well‐characterized orga-
host cell, what we today call the nucleus. The fact that the nelle–organelle contact in cell biology is the ER–mitochondria
nuclear membrane is continuous with the ER, that all the orga- junction (commonly referred to as mitochondria‐associated
nelles are almost always in contact with the ER (Figure 13.2), (ER) membranes, or “MAMs”). When considering the inside‐
and that the ER plays a key role in mitochondrial and endo- out origin story of the eukaryotic cell, it is clear that ER–­
membrane dynamics are natural outcomes of this evolutionary mitochondria interactions were preordained long before they
­history. Similarly, the essential roles of ER–organelle contacts existed in their current form – protomitochondria and the host
make much more sense in the context of having existed since the cell (i.e. the protonucleus or, in this specific context, proto‐ER)
eukaryotic primordial era. The largest organelle in the cell, the were interacting long before eukaryotic cells or organelles
ER physically occupies more of the cell than any other existed. Thus, it should be unsurprising these two organelles
­organelle  –  approximately 35%. When considering ER move- are highly coordinated, working together to control key physi-
ment over time, over 90% of the cellular volume is in contact ological processes such as lipid synthesis, calcium signaling,
with the ER at one point or another in less than 5 minutes [4, 5]. and even cell death. But what was not obvious until recently is

Plasma
membrane

ER tubules

ER sheet

Endosomes

Mitochondria

Peroxisomes

Lipid droplets

Golgi

MCS

Nuclear envelope

Figure 13.2  Structure of endoplasmic reticulum (ER) membrane contact sites (MCSs). The ER consists of the nuclear envelope (outlined with a
dashed line) and the peripheral ER, which spreads into the cytosol as a network of sheets and tubules. The peripheral ER forms MCSs with the plasma
membrane, mitochondria, endosomes, peroxisomes, lipid droplets, and the Golgi. Reproduced from [16] with permission of Springer Nature.
154 THE LIVER:  THE ER AND MITOCHONDRIA: AN ANCIENT MARRIAGE

the ER also plays a key role in regulating mitochondrial Calcium signaling


­biogenesis and motility, dictating when and where new mito-
chondria are formed, and where they are going [6]. One of the most profound manifestations of the integral role of
In both yeast and animal cells, MAMs have been character- mitochondria in the evolution of multicellular life is apoptosis,
ized with high‐resolution microscopy, revealing an average dis- wherein mitochondria can control whether a cell will live or die.
tance of 10–30 nm, often with protein tethers visible by electron This is essential for multicellular life and a classic example is in
microscopy appearing at these junctions [7]. Live cell micros- development wherein “pruning” removes structures unnecessary
copy reveals that ER and mitochondria will often move together, later on in life. That apoptosis is manifested by what was once an
illustrating that these two organelles can remain tethered as they entirely different organism raises deep philosophical questions
move about the cell [4]. The majority of MAMs involve extended about the nature of the modern cell. Who is really in charge? Are
regions of membrane apposition, in which the two membranes we but slaves to our protomitochondrial ancestors, whose only
(ER and the mitochondrial outer membrane) are parallel to one function is to gather resources and protect our mitochondrial
another. A smaller fraction of MAMs involve ER tubules wrap- overlords? Although the precise mechanisms are beyond the
ping around mitochondria orthogonally to their longer axis (most scope of this chapter, the essential role of apoptosis in health and
mitochondria adopt a tubular morphology). The two most well‐ disease highlights yet another key role of MAMs  –  regulating
studied functions of MAMs – lipid exchange and calcium signaling – calcium transfer from the ER to mitochondria, which in turn can
are thought to be largely mediated at the parallel types of trigger apoptosis. Interestingly, calcium influx from the ER to
contacts, by virtue of the larger amount of contact area available mitochondria appears to occur in specific subdomains, highlight-
in this conformation. The orthogonally conjoined MAMs were ing the existence of a highly conserved, highly organized molec-
only recently discovered to play a key role in mitochondrial ular interaction that facilitates the process [14, 15].
dynamics and will be discussed at the end of this section. There are at least three known physiological purposes for
calcium transfer from the ER to mitochondria. The calcium
concentration in the ER is usually somewhere between 100 and
Phospholipid synthesis 500 μM, whereas cytoplasmic calcium is usually around 100 nM.
The vast majority of lipid synthesis enzymes are localized to the Thus, MAMs can provide a source of highly concentrated
ER membrane, but several key enzymes reside within the mito- ­calcium, which may be necessary for the activation of calcium‐
chondrial outer membrane. Some phospholipid synthesis dependent mitochondrial proteins and processes [6, 16]. For
requires several key enzymes, some of which are on the ER, and example, elevated calcium levels in the mitochondrial matrix
others on mitochondria. Long before microscopes were capable are required to activate the tricarboxylic acid cycle for the gen-
of live imaging of MAMs, biochemical analyses revealed the eration of ATP, and the highly selective mitochondrial calcium
presence of phosphatidylserine (PS) synthase at MAMs [8]. Two uniporter on the mitochondrial inner membrane is unlikely to
additional phospholipids – phosphatidylcholine (PC) and phos- pass high enough concentrations of calcium anywhere other
phatidylethanolamine (PE) – are also known to be dependent on than near MAMs that facilitate direct transfer of calcium from
and synthesized at MAMs. PE and PC synthesis illustrates the the ER [6, 16–18].
tight coordination and exchange of material between the ER and As mentioned above, calcium influx from the ER to the mito-
mitochondria: PS, first synthesized in the ER, is then transferred chondria also plays a role in apoptosis. Calcium stimulates the
to the mitochondrial outer membrane and then from there to the opening of the mitochondrial permeability transition pore,
mitochondrial inner membrane, whereupon it is enzymatically which leads to release of cytochrome c, which then leads to the
converted to PE. To generate PC, PE is then translocated from apoptotic caspase signaling cascade. Blocking the ER calcium
the mitochondrial inner membrane through the outer membrane channel Ins(1,4,5)P3R confers apoptotic resistance in several
and then back to the ER, whereupon ER‐resident enzymes enzy- cell lines [19, 20]. Blocking key ER–mitochondria tethering
matically convert PE to PC. Fascinatingly, PC is also found on complexes, such as the mitochondrial outer membrane protein
the mitochondrial outer membrane, indicating that PC is translo- FIS1 (also implicated in mitochondrial fission) and the ER pro-
cated from ER back to mitochondria [8–12]. This ping‐pong pat- tein BAP31, also confers protection against apoptosis [21].
tern of phospholipid translocation and modification likely began Finally, the key GTPase protein involved in mitochondrial fis-
to occur prior to the evolution of the eukaryotic cell and may sion, dynamin‐related protein 1 (Drp1), is also important in
have served as a key selective factor for the eventual ectosymbi- mitochondrial outer membrane permeabilization (MOMP) that
otic relationship between protomitochondria and the protonuclei. is required for cytochrome c release and consequent apoptosis
In fact, evidence exists that all eukaryotic phospholipids were [22]. When calcium is released from the ER to mitochondria,
inherited from protomitochondria [13]. Thus, today the steady‐ the apoptotic proteins BAX and BAK facilitate MOMP while
state levels of key phospholipids in each of our organelles is also facilitating stable localization of Drp1 to the outer mito-
clearly demarcated by MAM machinery that was constructed chondrial membrane [23–25]. Taken together, these findings
during the courting period of the eukaryotic cell’s ancestors [2]. demonstrate a strong link between calcium release, apoptosis,
The manifold roles of phospholipids in many human diseases are and mitochondrial fission.
beyond the scope of this chapter other than to point out that this
is largely a consequence of the foundational importance of the
Mitochondrial fission
proper functioning of MAMs, which in turn is a consequence of
their evolutionary history, without which the modern eukaryotic Interestingly, mitochondrial fission (a precursor to both mito-
cell as we know it would not exist. chondrial biogenesis and degradation) is also mediated by
13:  Organelle–Organelle Contacts: Origins and Functions 155

calcium [26]. The precise mechanisms regulating mitochondrial extracellular space to/from the Golgi and the lysosome. Contacts
fission are both beyond the scope of this chapter, and not yet between the ER and endosomes play multiple roles in regulating
completely understood. However, recent work has revealed an endosomal function, depending on the context and subcellular
entirely new role for ER–mitochondria interactions in regulat- location. The ER contains phosphatases that can dephosphoryl-
ing mitochondrial fission. The first key finding was that ER ate important cargo such as epidermal growth factor receptor
tubules are almost always orthogonally oriented with respect to (EGFR) [36]. ER–endosome contacts can regulate endosomal
mitochondrial fission sites, and that these intersections repre- motility in a cholesterol‐dependent fashion [37], and cholesterol
sent mitochondrial constrictions [27]. These constrictions are may be trafficked from endosomes to the ER at contact sites
thought to be necessary for Drp1 oligomers to circumscribe [38]. Given the importance of these cellular processes and in
mitochondria since Drp1 oligomer rings are ~110 nm in diam- particular cholesterol, it is perhaps not surprising that multiple
eter, whereas a non‐constricted mitochondrion has a diameter diseases are caused by mutations in ER–endosome contact pro-
closer to 350 nm [27]. Interestingly, the actin cytoskeleton was teins [39–42]. Along the same lines, targeting ER–endosome
soon shown to play a key role in this process [26, 28–31], both contacts presents an attractive target for potential therapies for
by providing the force necessary for constriction (how else metabolic disorders.
could a 50 nm ER tubule produce the force necessary to con- More recently, it was shown that, similarly to mitochondria,
strict the double‐membraned 350 nm mitochondrion?) and by endosomal fission is also mediated by ER tubules wrapping
providing a scaffold for and activation of Drp1 GTPase activity around and constricting endosomes at their fission sites [43].
[32]. These studies also led to the observation that, prior to The multi‐destination character of endosomal trafficking neces-
mitochondrial fission, Drp1 proteins are not just diffusing sitates the existence of spatiotemporal regulation of endosomal
around in the cytoplasm as previously thought, but rather are fission such that cargo with different destinies are segregated at
localized to the cytoplasmic surface of ER tubules [33]. The the appropriate location and time. The specific destination and
actin polymerization involved in this process is regulated by an direction of cargo traffic is of course regulated by many molecu-
ER‐anchored member of the formin protein family called INF2 lar components, but the precise spatial location of endosomal
(which is implicated in Charcot–Marie–Tooth disease and focal fission was recently shown to be regulated by no less than the
segmental glomerulosclerosis) and the mitochondrial outer ER, in a fashion quite similar to that observed for mitochondria:
membrane‐anchored protein Spire1C [26, 28, 30, 34]. ER tubules wrap around endosomes at fission sites, at which
The processes described above delineate a (relatively) clear point actin regulatory proteins on the endosomal surface (which
mechanism for outer membrane fission, but how inner mem- includes endosomal isoforms of Spire family proteins) and
brane fission is regulated by the ER is less clear in this scenario. dynamin proteins (related to Drp1) cooperate to drive endoso-
Until you consider calcium. Follow‐up studies quickly showed mal fission [43]. It has not yet been shown whether ER‐anchored
that calcium influx from the ER to mitochondria via the mito- INF2 similarly plays a role in Spire‐mediated endosomal fis-
chondrial calcium uniporter (MCU) was able to drive mitochon- sion, but disruption of either of these proteins does appear to
drial inner membrane fission, even in the absence of outer result in decreased endosomal fission (unpublished data), sug-
membrane fission [26]. Interestingly, this calcium exchange gesting a conserved molecular mechanism for ER‐mediated
appears to be somewhat dependent on the presence of INF2, ­fission of both mitochondria and endosomes. Complicating the
indicating that both inner and outer membrane fission can be story are studies that indicate that endosome‐associated Rab
controlled by a complementary set of proteins. In a process that GTPase proteins also regulate mitochondrial fission [44, 45],
remains less well understood, it appears that mitochondrial directly coupling the endolysosomal machinery to the mito-
DNA (mtDNA) nucleoids also appear to divide at ER–mito- chondrial fission machinery. At this point, it is perhaps reason-
chondria fission sites, providing an ER‐regulated mechanism able to wonder whether puppet‐master ER brings all the
for coupling mtDNA and mitochondrial biogenesis, or, in the necessary components together to regulate fission of all the
case of mitophagy, the degradation of dysfunctional mtDNA organelles.
and mitochondria components while preserving functional
mitochondria [35]. In conclusion, mitochondrial fission is an
intricately coordinated process that involves multiple compo-
nents with complementary, yet completely independent func- THE ER AND GOLGI: CLOSE BUT
tions in each step, all dependent on interactions between the ER, MYSTERIOUS
mitochondrial inner and outer membranes and mtDNA, and the
actin cytoskeleton. The Golgi complex is the central hub of the cellular secretory
pathway. Immediately adjacent to the ER, the Golgi is known
to receive proteins and lipid cargos from the ER, process those
cargos further with post‐translational modifications, and then
THE ER AND ENDOSOMES: HINTS OF sort them on towards their final destination. The Golgi is com-
A UNIVERSAL MECHANISM pletely dependent on input from the ER; blocking traffic from
the ER to the Golgi eventually results in the total disintegra-
Given that the ER interacts with every organelle in the cell, it tion of the Golgi. Trafficking is bidirectional and cargos can
should not be surprising that the ER interacts with endosomes in be transported from the Golgi to the ER as well [46–51].
multiple contexts. Endosomes are membrane vesicles involved These well‐known, well‐characterized functional interactions
in the trafficking of cargos to/from the plasma membrane and between the Golgi and ER are one of the first things we are
156 THE LIVER:  ER, MITOCHONDRIA, AND PEROXISOMES: SHARED CUSTODY?

taught when introduced to cell biology. So you might expect ER AND LIPID DROPLETS: OLD
that the physical interactions between Golgi and ER mem- “BUD” DIES?
branes are among the most well characterized. But you would
be wrong. Lipid droplets (LDs) were thought to be passive lipid‐storage
What is known about direct ER–Golgi membrane interac- facilities until the discovery of multiple proteins and active
tions is that they facilitate lipid exchange. One well‐studied functions for LDs. Although LDs have been shown to make con-
example of this is the transfer of ceramide from the ER to the tacts with the ER in electron microscopy, much like ER–
Golgi, as a first step towards glycosphingolipid and sphingomy- mitochondria contacts, it is thought that ER–LD contacts are
elin synthesis in the Golgi [52]. Although the proteins involved unique for a number of reasons, mainly that LDs are known to
in ceramide transfer from Golgi to ER are known, whether they originate from the ER. It is thought that when lipid esters in the
are the key proteins involved in maintaining ER–Golgi contacts ER–lipid bilayer reach a critical concentration, they are no
remains unclear. Nor do we know how ER–Golgi contacts are longer stable, and cause the lipid membrane to bulge and even-
regulated. Interestingly, regulation of the cytoskeleton by INF2 tually bud off into a nascent LD [67]. Thus, ER–LD contacts
and the ER seem to be important for regulating normal Golgi have been speculated to contain continuous membranes at some
structure [53], once again tempting one to think there are con- point in the LD life cycle but have not yet been directly observed.
served mechanisms involved in ER‐ and actin‐mediated regula- New high‐resolution cellular cryoelectron microscopy techniques
tion of organelle structure and dynamics. may finally be able to conclusively answer this question.
The importance of ER–LD contacts is highlighted by the fact
that mutations in one of the key ER–LD tethering proteins,
seipen, cause Berardinelli–Seip congenital lipodystrophy. This
THE ER AND AUTOPHAGOSOMES: disease causes a lack of adipose tissue and fat deposition in the
ANOTHER UNIVERSAL MECHANISM liver [68]. Cells with mutations in seipin have an abnormally
high number of cells with either small LDs or greatly oversized
Autophagy is the main mechanism by which the cell degrades LDs [69]. Fascinatingly, sometimes LDs are observed inside the
intracellular components. The autophagosome is assembled at nucleus, contacting the inner nuclear reticular membrane [70],
least in part by the ER, often triggered by stress. Assembly of highlighting once more the historical evolutionary framework
the autophagosome depends on the coordinated recruitment of viewing the ER as an extension of the nucleus itself.
membrane proteins to the ER at the autophagosome assembly Taking us even deeper into our evolutionary framework, we
site. Interestingly, ER–mitochondria contact sites were recently must also consider recent work demonstrating the importance of
shown to be a preferred location for the recruitment of these contacts between LDs and mitochondria (LD–mitos). Although
proteins – mitochondria even appear to contribute membrane to LD–mito contacts have been observed in brown adipose tissue
autophagosomes [54–58], and this process is likely to play a (BAT) [71], heart [72], and type I skeletal muscle [73], their
role in maintaining homeostasis in hepatocytes [59–61]. It is function remained unclear. Two recent studies shed new light on
perhaps only fitting that the same organelle involved in dictat- how LD–mitos cooperate to regulate metabolism and energetic
ing the timing of the cell’s death (i.e. apoptosis) also controls flux in the cell. First, Rambold et al. in 2015 showed that fatty
the cell’s self‐cannibalism. That said, while autophagy is often acids are trafficked from LDs to mitos during starvation, ena-
associated with cell death, it is also known to promote survival, bling the cell to switch from glycolysis to β‐oxidation for ATP
and thus so‐called “autophagic cell death” may in fact be energy production [74]. More recently, an elegant study from
reversing causality and blaming ambulances for accidents. ER– the Shirihai lab showed that mitochondria directly contacting
mitochondria contacts are not necessary for autophagosome LDs have distinct features that facilitate LD biogenesis and tria-
assembly: ER–plasma membrane contacts have also been cylglyceride synthesis [75]. Although this leaves some ambigu-
shown to be the site of autophagosome biogenesis [62, 63]. ity about whether LD–mito contacts facilitate catabolism or
Interestingly, VAPs, one of the key protein families mediating biogenesis of LDs (it almost certainly depends heavily on vari-
ER–autophagosome contacts and autophagosome biogenesis, ations in cell and tissue type and context), the role of LD–mito
also mediate contacts with multiple other organelles, including contacts in metabolism is unimpeachable.
the plasma membrane, mitochondria, lysosomes, endosomes,
and Golgi [64, 65].
Although the physical interaction of ER and autophagosomes is
clear – they literally serve as hubs for autophagosome biogenesis – ER, MITOCHONDRIA, AND
the functional purpose (or origin) of this intimate relationship PEROXISOMES: SHARED CUSTODY?
is less so. The answer maybe lies in the fact that the ER is
­perhaps the central player in the cell stress response. This fram- Although specific tethers between peroxisomes and the ER or
ing is supported by the finding that disrupting autophagy leads mitochondria are not clearly defined, a discussion on organelle–
to increased ER stress and subsequent cell death. It is thus organelle interactions would be incomplete without discussing
thought that ER stress‐induced autophagy may be a last‐ditch peroxisome biogenesis, which has been proposed to originate
effort by the cell to survive before succumbing to stress‐induced both from the ER and from mitochondria [76, 77]. That the ER
apoptosis [66]. contacts all organelles at high spatiotemporal frequency makes
13:  Organelle–Organelle Contacts: Origins and Functions 157

it difficult to discern specific functions of ER–organelle con- points to a model wherein peroxisomes evolved from a need to
tacts. This difficulty is further confounded by the fact that all keep mitochondrial ROS generation in check. Thus, the physical
membrane‐bound organelles will contain proteins that at one biogenesis of peroxisomes from both ER and mitochondrial
point or another were localized on the ER. So‐called pre‐peroxi- membranes provides a strong testament to the entire theory of
somal vesicles are thought to shuttle proteins from the ER to symbiosis, in particular our favored “inside‐out” model. For an
peroxisomes, presumably facilitating peroxisomal biogenesis interesting discussion of the symbiotic theory of peroxisome
without any direct contacts or tethering [76, 77]. However, ER– evolution, see the 2017 essay by D. Speijer [79].
peroxisome tethers have indeed been reported (mediated by our
aforementioned friends in the VAP family), and these tethers are
now known to be important for phospholipid synthesis, regula-
tion of cholesterol levels, and peroxisome growth [78]. SOME FINAL REMARKS
At the same time, new evidence is emerging that peroxisomes
are derived from both the ER and mitochondria. The acceptance Given the number of organelle types, the number of permuta-
of this model, almost exclusively pioneered by the McBride tions of transorganelle contacts can get high, especially when
laboratory [76, 77], was severely hampered by yeast studies that considering tri‐ or even quad‐organelle contacts (e.g. LDs, ER
showed no involvement of mitochondria. Intriguingly, this both and mitochondria, or endosomes, ER, autophagosomes, and
highlights a key evolutionary step in higher eukaryotes and also mitochondria). There is new evidence for co‐transport of orga-
explains the physiological role of peroxisomes. Specifically, nelles by “piggybacking” on one another [80, 81], and it is
peroxisomes have been posited to be indispensable for regulat- highly likely there will be new instances of such noncanonical
ing mitochondrial reactive oxygen species (ROS) generated by transport revealed in the near future. So this chapter is by no
fatty acid metabolism. Notably, yeast mitochondria long ago means comprehensive. The ER–plasma membrane contacts
discarded their ability to β‐oxidize fatty acids – hence the dis- were barely mentioned, and I intentionally focused on mamma-
crepancy. In mammalian cells, however, all evidence currently lian‐specific organelles, so no mention of chloroplasts either.

Table 13.1  List of known organelle membrane contact proteins in metazoans


Contact site Protein name Description Reference
ER–PM VAPs ER receptors for numerous proteins containing FFAT motif [65]
STIM1, Orai Dynamic tether: PM Ca2+ channel Orai binds to ER integral STIM1 at low [83]
luminal Ca2+
E‐Syt1/2/3 SMP domain‐containing tethers; E‐Syt1 is a Ca2+‐dependent dynamic tether [84]
Junctophylin1/2/3/4 ER residents, bind PM via MORN domains [85]
DHPR, RyR PM and ER Ca2+ channels, interact and function in a concerted way [86]
ORP5, ORP8 LTPs, dynamic tethers, contain TMD and PH domain [87]
ER–mitochondria MFN1/2 Mitochondrial fusion GTPase; an ER MFN2 pool mediates tethering by [88]
interacting with mitochondrial MFN1/2
IP3R, VDAC, Grp75 ER Ca2+ release channel IP3R and mitochondrial metabolite channel VDAC [89]
are connected. Grp75 may be involved
Fis1, BAP31 Mitochondrial Fis1 and ER BAP31 interact for transmission of apoptotic [21]
signals
PTPIP51, VAP PTPIP51 is a mitochondrial LTP structurally equipped for tethering via [90]
VAP binding
ER–endosome VAPs ER receptors for numerous proteins containing FFAT motif [65]
StARD3, StARD3NL Integral endosomal proteins, interact with ER proteins VAP via FFAT motif [91]
ORP1L, ORP5 LTPs active at the ER–endosome interface [37, 38]
PTP1B, EGFR, Annexin A1 Components mediating interplay between ER and multivesicular bodies [36, 92]
Protrudin, Rab7 ER protrudin interacts with Rab7 and phosphatidylinositol‐3‐phosphate on [93]
late endosomes
ER–Golgi VAPs ER receptors for numerous proteins containing FFAT motif [65]
OSBP LTP, dynamic phosphatidylinositol‐4‐phosphate‐dependent tether, contains [94]
FFAT motif and PH domain
CERT LTP, putative dynamic tether, contains FFAT motif and PH domain [95]
FAPP2 LTP, structurally equipped for tethering (FFAT motif, PH domain) [52]
Nir2 Phosphatidylinositol transfer protein, contains FFAT motif [95]
Lysosome–peroxisome Synaptotagmin‐7 Mediates lysosome–peroxisome tethering important for cholesterol transfer [96]
ER–lipid droplet DGAT2, FATP1 Lipid droplet‐resident DGAT2 and ER‐resident FATP1 interact and [97]
coordinate lipid droplet biogenesis
Mitochondria–lipid Perilipin‐5 Lipid droplet scaffold protein involved in interaction with mitochondria [98]
droplet
Mitochondrial Mic60/27/26/25/19/10, Mitochondrial contact site and cristae organizing system (MICOS), an [99]
IM–OM Qil1 integral protein complex of the inner membrane
SAMM50, Metaxin 1/2 MICOS inteaction partners in the outer mitochondrial membrane
Reproduced from [64] with permission of Elsevier.
FFAT, phenylalanine in an acidic tract; SMP, synaptotagmin‐like mitochondrial lipid‐binding protein; MORN, membrane occupation and recognition nexus; LTP, lipid transfer
protein; TMD, transmembrane domain; PH, pleckstrin homology; IM, mitochondrial inner membrane; OM, mitochondrial outer membrane.
158 THE LIVER:  REFERENCES

Needless to say, the transorganelle contact field is large and still 19. Jayaraman, T. and Marks, A.R. T cells deficient in inositol 1,4,5‐trisphosphate
in its infancy. Indeed, there is at least one entire book dedicated receptor are resistant to apoptosis. Mol Cell Biol, 1997;17(6):3005–12.
20. Khan, A.A. et al. Lymphocyte apoptosis: mediation by increased type 3 ino-
to organelle contact sites [82]. In lieu of discussing all the sitol 1,4,5‐trisphosphate receptor. Science, 1996;273(5274):503–7.
known proteins involved in transorganelle contacts in this one 21. Iwasawa, R. et al. Fis1 and Bap31 bridge the mitochondria‐ER interface to
chapter, readers can refer to Table 13.1 (taken from [64]) and the establish a platform for apoptosis induction. EMBO J, 2011;30(3):556–68.
references therein for additional details [21, 36–38, 52, 65, 22. Chan, D.C. Fusion and fission: interlinked processes critical for mitochon-
drial health. Annu Rev Genet, 2012;46:265–87.
83–99]. Regularly searching for “organelle contacts,” “transor-
23. Lucken‐Ardjomande, S. and Martinou, J.C. Regulation of Bcl‐2 proteins and
ganelle,” and “membrane contact sites” on PubMed or Google of the permeability of the outer mitochondrial membrane. C R Biol,
Scholar will likely reveal new publications every week. Novel 2005;328(7):616–31.
proteomics methods with increasing spatiotemporal resolution 24. Montessuit, S. et al. Membrane remodeling induced by the dynamin‐related
promise to allow for mapping of novel proteins involved in protein Drp1 stimulates Bax oligomerization. Cell, 2010;142(6):889–901.
25. Wasiak, S., Zunino, R., and McBride, H.M. Bax/Bak promote sumoylation
organelle–organelle contacts in different conditions (e.g. cell
of DRP1 and its stable association with mitochondria during apoptotic cell
stress, in the presence of disease‐associated mutations, etc.) death. J Cell Biol, 2007;177(3):439–50.
[100], and subtractive analysis thereof will provide detailed 26. Chakrabarti, R. et al. INF2‐mediated actin polymerization at the ER stimu-
insight towards how these complexes regulate and are regulated lates mitochondrial calcium uptake, inner membrane constriction, and divi-
by physiological cues. The bad news is this means we have sion. J Cell Biol, 2018;217(1):251–68.
27. Friedman, J.R. et  al. ER tubules mark sites of mitochondrial division.
much more work to do. The good news is this means novel drug Science, 2011;334(6054):358–62.
targets and therapies are likely to multiply in the coming years. 28. Korobova, F., Ramabhadran, V., and Higgs, H.N. An actin‐dependent step in
With some luck and much effort, this golden age of organelle mitochondrial fission mediated by the ER‐associated formin INF2. Science,
biology will lead to a corresponding revolution in medicine. 2013;339(6118):464–7.
29. Moore, A.S. et al. Dynamic actin cycling through mitochondrial subpopula-
tions locally regulates the fission‐fusion balance within mitochondrial net-
works. Nat Commun, 2016;7:12886.
30. Manor, U. et  al. A mitochondria‐anchored isoform of the actin‐nucleating
REFERENCES spire protein regulates mitochondrial division. Elife, 2015;4.
31. Korobova, F., Gauvin, T.J., and Higgs, H.N. A role for myosin II in mam-
1. Sagan, L. On the origin of mitosing cells. J Theor Biol, 1967;14(3):255–74. malian mitochondrial fission. Curr Biol, 2014;24(4):409–14.
2. Baum, D.A. and Baum, B. An inside‐out origin for the eukaryotic cell. BMC 32. Ji, W.K. et  al. Actin filaments target the oligomeric maturation of the
Biol, 2014;12:76. dynamin GTPase Drp1 to mitochondrial fission sites. Elife, 2015;4:e11553.
3. Dey, G., Thattai, M., and Baum, B. On the archaeal origins of eukaryotes and 33. Ji, W.K. et  al. Receptor‐mediated Drp1 oligomerization on endoplasmic
the challenges of inferring phenotype from genotype. Trends Cell Biol, reticulum. J Cell Biol, 2017;216(12):4123–39.
2016;26(7):476–85. 34. Curchoe, C.L. and Manor, U. Actin cytoskeleton‐mediated constriction of
4. Valm, A.M. et al. Applying systems‐level spectral imaging and analysis to membrane organelles via endoplasmic reticulum scaffolding. ACS Biomater
reveal the organelle interactome. Nature, 2017;546(7656):162–7. Sci Eng, 2017;3(11):2727–32.
5. Cohen, S., Valm, A.M., and Lippincott‐Schwartz, J. Interacting organelles. 35. Lewis, S.C., Uchiyama, L.F., and Nunnari, J. ER‐mitochondria contacts cou-
Curr Opin Cell Biol, 2018;53:84–91. ple mtDNA synthesis with mitochondrial division in human cells. Science,
6. Rowland, A.A. and Voeltz, G.K. Endoplasmic reticulum‐mitochondria 2016;353(6296):aaf5549.
contacts: function of the junction. Nat Rev Mol Cell Biol,
­ 36. Eden, E.R. et al. Membrane contacts between endosomes and ER provide
2012;13(10):607–25. sites for PTP1B‐epidermal growth factor receptor interaction. Nat Cell Biol,
7. Csordas, G. et al. Structural and functional features and significance of the 2010;12(3):267–72.
physical linkage between ER and mitochondria. J Cell Biol, 2006; 37. Rocha, N. et al. Cholesterol sensor ORP1L contacts the ER protein VAP to
174(7):915–21. control Rab7‐RILP‐p150 Glued and late endosome positioning. J Cell Biol,
8. Vance, J.E. Phospholipid synthesis in a membrane fraction associated with 2009;185(7):1209–25.
mitochondria. J Biol Chem, 1990;265(13):7248–56. 38. Du, X. et al. A role for oxysterol‐binding protein‐related protein 5 in endo-
9. Voelker, D.R. Interorganelle transport of aminoglycerophospholipids. somal cholesterol trafficking. J Cell Biol, 2011;192(1):121–35.
Biochim Biophys Acta, 2000;1486(1):97–107. 39. Ueno, S. et al. The gene encoding a newly discovered protein, chorein, is
10. van Meer, G., Voelker, D.R., and Feigenson, G.W. Membrane lipids: where mutated in chorea‐acanthocytosis. Nat Genet, 2001;28(2):121–2.
they are and how they behave. Nat Rev Mol Cell Biol, 2008;9(2):112–24. 40. Belvedere, R. et al. Annexin A1 contributes to pancreatic cancer cell pheno-
11. Osman, C., Voelker, D.R., and Langer, T. Making heads or tails of phospho- type, behaviour and metastatic potential independently of formyl peptide
lipids in mitochondria. J Cell Biol, 2011;192(1):7–16. receptor pathway. Sci Rep, 2016;6:29660.
12. Stone, S.J. and Vance, J.E. Phosphatidylserine synthase‐1 and ‐2 are localized 41. Kondo, I. et al. COH1 analysis and linkage study in two Japanese families
to mitochondria‐associated membranes. J Biol Chem, 2000;275(44): with Cohen syndrome. Clin Genet, 2005;67(3):270–2.
34534–40. 42. Hashimoto, Y. et al. Protrudin regulates endoplasmic reticulum morphology
13. Lykidis, A. Comparative genomics and evolution of eukaryotic phospholipid and function associated with the pathogenesis of hereditary spastic paraple-
biosynthesis. Prog Lipid Res, 2007;46(3–4):171–99. gia. J Biol Chem, 2014;289(19):12946–61.
14. Rizzuto, R. et al. Close contacts with the endoplasmic reticulum as determi- 43. Rowland, A.A. et al. ER contact sites define the position and timing of endo-
nants of mitochondrial Ca2+ responses. Science, 1998;280(5370):1763–6. some fission. Cell, 2014;159(5):1027–41.
15. Rizzuto, R. et  al. Microdomains with high Ca2+ close to IP3‐sensitive 44. Wong, Y.C., Ysselstein, D., and Krainc, D. Mitochondria‐lysosome contacts
channels that are sensed by neighboring mitochondria. Science, 1993.
­ regulate mitochondrial fission via RAB7 GTP hydrolysis. Nature,
262(5134):744–7. 2018;554(7692):382–6.
16. Phillips, M.J. and Voeltz, G.K. Structure and function of ER membrane con- 45. Landry, M.C. et al. A functional interplay between the small GTPase Rab11a
tact sites with other organelles. Nat Rev Mol Cell Biol, 2016;17(2):69–82. and mitochondria‐shaping proteins regulates mitochondrial positioning and
17. Rizzuto, R. et al. Mitochondria as sensors and regulators of calcium signal- polarization of the actin cytoskeleton downstream of Src family kinases. J
ling. Nat Rev Mol Cell Biol, 2012;13(9):566–78. Biol Chem, 2014;289(4):2230–49.
18. Baughman, J.M. et al. Integrative genomics identifies MCU as an essential 46. Lippincott‐Schwartz, J. et al. Rapid redistribution of Golgi proteins into the
component of the mitochondrial calcium uniporter. Nature, 2011; ER in cells treated with brefeldin A: evidence for membrane cycling from
476(7360):341–5. Golgi to ER. Cell, 1989;56(5):801–13.
13:  Organelle–Organelle Contacts: Origins and Functions 159

47. Lippincott‐Schwartz, J. et al. Microtubule‐dependent retrograde transport of mitochondrial enzyme activity. Am J Physiol Regul Integr Comp Physiol,
proteins into the ER in the presence of brefeldin A suggests an ER recycling 2007;292(3):R1271–8.
pathway. Cell, 1990;60(5):821–36. 74. Rambold, A.S., Cohen, S., and Lippincott‐Schwartz, J. Fatty acid traffick-
48. Lippincott‐Schwartz, J. Bidirectional membrane traffic between the endo- ing in starved cells: regulation by lipid droplet lipolysis, autophagy, and
plasmic reticulum and Golgi apparatus. Trends Cell Biol, 1993;3(3):81–8. mitochondrial fusion dynamics. Dev Cell, 2015;32(6):678–92.
49. Presley, J.F. et al. ER‐to‐Golgi transport visualized in living cells. Nature, 75. Benador, I.Y. et al. Mitochondria bound to lipid droplets have unique bio-
1997;389(6646):81–5. energetics, composition, and dynamics that support lipid droplet expansion.
50. Ward, T.H. et al. Maintenance of Golgi structure and function depends on the Cell Metab, 2018;27(4):869–85 e6.
integrity of ER export. J Cell Biol, 2001;155(4):557–70. 76. Mohanty, A. and McBride, H.M. Emerging roles of mitochondria in the
51. Sengupta, P. et al. ER trapping reveals Golgi enzymes continually revisit the evolution, biogenesis, and function of peroxisomes. Front Physiol,
ER through a recycling pathway that controls Golgi organization. Proc Natl 2013;4:268.
Acad Sci U S A, 2015;112(49):E6752–61. 77. Sugiura, A. et al. Newly born peroxisomes are a hybrid of mitochondrial
52. D’Angelo, G. et al. Glycosphingolipid synthesis requires FAPP2 transfer of and ER‐derived pre‐peroxisomes. Nature, 2017;542(7640):251–4.
glucosylceramide. Nature, 2007;449(7158):62–7. 78. Hua, R. et al. VAPs and ACBD5 tether peroxisomes to the ER for peroxi-
53. Ramabhadran, V. et  al. Splice variant‐specific cellular function of the some maintenance and lipid homeostasis. J Cell Biol, 2017;216(2):367–77.
formin  INF2 in maintenance of Golgi architecture. Mol Biol Cell, 79. Speijer, D. Evolution of peroxisomes illustrates symbiogenesis. Bioessays,
2011;22(24):4822–33. 2017;39(9).
54. Hamasaki, M. et al. Autophagosomes form at ER‐mitochondria contact sites. 80. Salogiannis, J. and Reck‐Peterson, S.L. Hitchhiking: a non‐canonical mode
Nature, 2013;495(7441):389–93. of microtubule‐based transport. Trends Cell Biol, 2017;27(2):141–50.
55. Gomez‐Suaga, P. et  al. The ER‐mitochondria tethering complex VAPB‐ 81. Salogiannis, J., Egan, M.J., and Reck‐Peterson, S.L. Peroxisomes move by
PTPIP51 regulates autophagy. Curr Biol, 2017;27(3):371–85. hitchhiking on early endosomes using the novel linker protein PxdA. J Cell
56. Rambold, A.S. and Lippincott‐Schwartz, J. Mechanisms of mitochondria Biol, 2016;212(3):289–96.
and autophagy crosstalk. Cell Cycle, 2011;10(23):4032–8. 82. Tagaya, M. Organelle Contact Sites: From Molecular Mechanism to
57. Hailey, D.W. et  al. Mitochondria supply membranes for autophagosome Disease. Springer, New York, 2017.
­biogenesis during starvation. Cell, 2010;141(4):656–67. 83. Liou, J. et  al. Live‐cell imaging reveals sequential oligomerization and
58. Rambold, A.S. and Lippincott‐Schwartz, J. Starved cells use mitochondria local plasma membrane targeting of stromal interaction molecule 1 after
for autophagosome biogenesis. Cell Cycle, 2010;9(18):3633–4. Ca2+ store depletion. Proc Natl Acad Sci U S A, 2007;104(22):9301–6.
59. Kang, S.W. et al. AMPK activation prevents and reverses drug‐induced mito- 84. Giordano, F. et al. PI(4,5)P(2)‐dependent and Ca2+‐regulated ER‐PM interac-
chondrial and hepatocyte injury by promoting mitochondrial fusion and tions mediated by the extended synaptotagmins. Cell, 2013;153(7):1494–509.
function. PLoS One, 2016;11(10):e0165638. 85. Takeshima, H., Hoshijima, M., and Song, L.S. Ca2+ microdomains organ-
60. Fu, D. et al. Coordinated elevation of mitochondrial oxidative phosphoryla- ized by junctophilins. Cell Calcium, 2015;58(4):349–56.
tion and autophagy help drive hepatocyte polarization. Proc Natl Acad Sci U 86. Rebbeck, R.T. et  al. The beta(1a) subunit of the skeletal DHPR binds to
S A, 2013;110(18):7288–93. skeletal RyR1 and activates the channel via its 35‐residue C‐terminal tail.
61. Fu, D., Lippincott‐Schwartz, J., and Arias, I.M. Increased mitochondrial Biophys J, 2011;100(4):922–30.
fusion and autophagy help isolated hepatocytes repolarize in collagen sand- 87. Chung, J. et  al. Intracellular transport. PI4P/phosphatidylserine counter-
wich cultures. Autophagy, 2013;9(12):2154–5. transport at ORP5‐ and ORP8‐mediated ER‐plasma membrane contacts.
62. Zhao, Y.G. et  al. The ER contact proteins VAPA/B interact with multiple Science, 2015;349(6246):428–32.
autophagy proteins to modulate autophagosome biogenesis. Curr Biol, 88. de Brito, O.M. and Scorrano, L. Mitofusin 2 tethers endoplasmic reticulum
2018;28(8):1234–45 e4. to mitochondria. Nature, 2008;456(7222):605–10.
63. Zhao, Y.G. and Zhang, H. The ER‐localized autophagy protein EPG‐3/ 89. Szabadkai, G. et al. Chaperone‐mediated coupling of endoplasmic reticu-
VMP1 regulates ER contacts with other organelles by modulating ATP2A/ lum and mitochondrial Ca2+ channels. J Cell Biol, 2006;175(6):901–11.
SERCA activity. Autophagy, 2018;14(2):362–3. 90. De Vos, K.J. et al. VAPB interacts with the mitochondrial protein PTPIP51
64. Eisenberg‐Bord, M. et al. A tether is a tether is a tether: tethering at mem- to regulate calcium homeostasis. Hum Mol Genet, 2012;21(6):1299–311.
brane contact sites. Dev Cell, 2016;39(4):395–409. 91. Alpy, F. et al. STARD3 or STARD3NL and VAP form a novel molecular
65. Murphy, S.E. and Levine, T.P. VAP, a versatile access point for the endoplas- tether between late endosomes and the ER. J Cell Sci, 2013;126(Pt
mic reticulum: review and analysis of FFAT‐like motifs in the VAPome. 23):5500–12.
Biochim Biophys Acta, 2016;1861(8 Pt B):952–61. 92. Eden, E.R. et al. Annexin A1 tethers membrane contact sites that mediate
66. Molino, D. et al. ER‐driven membrane contact sites: evolutionary conserved ER to endosome cholesterol transport. Dev Cell, 2016;37(5):473–83.
machineries for stress response and autophagy regulation? Commun Integr 93. Raiborg, C. et  al. Repeated ER‐endosome contacts promote endosome
Biol, 2017;10(5–6):e1401699. translocation and neurite outgrowth. Nature, 2015;520(7546):234–8.
67. Khandelia, H. et al. Triglyceride blisters in lipid bilayers: implications for 94. Mesmin, B. et  al. A four‐step cycle driven by PI(4)P hydrolysis
lipid droplet biogenesis and the mobile lipid signal in cancer cell mem- directs  sterol/PI(4)P exchange by the ER‐Golgi tether OSBP. Cell,
branes. PLoS One, 2010;5(9):e12811. 2013;155(4):830–43.
68. Magre, J. et  al. Identification of the gene altered in Berardinelli‐Seip 95. Peretti, D. et al. Coordinated lipid transfer between the endoplasmic reticu-
congenital lipodystrophy on chromosome 11q13. Nat Genet, 2001;
­ lum and the Golgi complex requires the VAP proteins and is essential for
28(4):365–70. Golgi‐mediated transport. Mol Biol Cell, 2008;19(9):3871–84.
69. Szymanski, K.M. et al. The lipodystrophy protein seipin is found at endo- 96. Chu, B.B. et al. Cholesterol transport through lysosome‐peroxisome mem-
plasmic reticulum lipid droplet junctions and is important for droplet mor- brane contacts. Cell, 2015;161(2):291–306.
phology. Proc Natl Acad Sci U S A, 2007;104(52):20890–5. 97. Xu, N. et al. The FATP1‐DGAT2 complex facilitates lipid droplet expan-
70. Ohsaki, Y. et al. PML isoform II plays a critical role in nuclear lipid droplet sion at the ER‐lipid droplet interface. J Cell Biol, 2012;198(5):895–911.
formation. J Cell Biol, 2016;212(1):29–38. 98. Wang, H. et  al. Perilipin 5, a lipid droplet‐associated protein, provides
71. Boutant, M. et  al. Mfn2 is critical for brown adipose tissue thermogenic physical and metabolic linkage to mitochondria. J Lipid Res,
function. EMBO J, 2017;36(11):1543–58. 2011;52(12):2159–68.
72. Wang, H. et al. Analysis of lipid droplets in cardiac muscle. Methods Cell 99. Zerbes, R.M. et al. Mitofilin complexes: conserved organizers of mitochon-
Biol, 2013;116:129–49. drial membrane architecture. Biol Chem, 2012;393(11).
73. Tarnopolsky, M.A. et al. Influence of endurance exercise training and sex on 100. Hung, V. et al. Spatially resolved proteomic mapping in living cells with the
intramyocellular lipid and mitochondrial ultrastructure, substrate use, and engineered peroxidase APEX2. Nat Protoc, 2016;11(3):456–75.
Gap and Tight Junctions
14 in Liver: Structure, Function,
and Pathology
John W. Murray1,2 and David C. Spray1,3,4
1
Marion Bessin Liver Research Center, Albert Einstein College of Medicine, Bronx, New York, NY, USA
2
Department of Anatomy and Structural Biology, Albert Einstein College of Medicine, Bronx,
New York, NY, USA
3
Department of Neuroscience, Albert Einstein College of Medicine, Bronx, New York, NY, USA
4
Department of Medicine, Albert Einstein College of Medicine, Bronx, New York, NY, USA

STRUCTURE OF JUNCTIONS intramembrane particles that form tight junctions, gap junc­
IN THE LIVER tion particles are clustered together to form discoid areas or
gap junction plaques (Figure  14.1). Gap junctions occupy a
This chapter is an update that integrates new findings on both large fraction, as much as 3%, of the total surface area of
tight junctions and gap juntions that have emerged since the pre­ hepatocytes [5]. In stained thin sections of liver tissue (and of
vious edition of this volume [1]. For more detailed historical hepatocyte cell pairs), gap junctions are recognized as septi­
background on gap junctions in particular, the reader is referred laminar linear membrane appositions separated by an extra­
to an even earlier edition [2]. cellular gap (Figure 14.1b,d). The seven lamina consist of the
Thin‐section and freeze‐fracture electron microscopy applied transparent extracellular gap sandwiched on either side by
to liver has revealed elaborate junctional complexes near the each cell’s three‐lamina membranes, consisting of two stained
apical cellular domains at appositional areas between hepato­ and thus electron‐opaque lipid head groups on each side of the
cytes (Figure 14.1). Tight junctions surround the bile canaliculi, electron‐lucent membrane interior. The overall thickness of
where they seal the paracellular spaces between hepatocytes, this double membrane specialization is approximately 15–18
regulating movement of solutes, ions, and water through the nm, and the extracellular space in the region of gap junctional
extracellular space, and act as a fence to block lateral diffusion contact is approximately 2–4 nm wide.
of membrane‐embedded molecules between basolateral and In freeze‐fracture images, gap junctions of liver and hepato­
apical membrane, thereby establishing and maintaining cellular cyte cell pairs are recognizable as arrays or plaques of intram­
polarity. In thin section, tight junctions appear as small, very embranous particles approximately 9 nm across present in the
close sites of membrane contact or “kisses” between the cell P‐fracture face with complementary pits on the E fracture face.
surfaces. In replicas of freeze‐fractured membranes, tight junc­ These plaques are generally round or oval and can be quite large
tions appear as continuous, branching, or anastomosing strands in hepatocytes (Figure 14.1c,d), commonly exceeding 1 μm in
of 10 nm particles in the P‐fracture (protoplasm, i.e. cytosol) diameter and containing more than 10 000 particles. The parti­
face (Figure  14.1a), with complementary grooves in the cles seen in freeze‐fracture replicas and the bridges across the
E‐(extracellular) face [4]. The resistance of the tight junctional extracellular space seen in thin section are believed to represent
barrier is proportional to the number of tight junction strands, channels with hydrophilic walls extending from the cytoplasmic
which in liver can be quite high. aspect of one cell to that of another (Figure 14.1).
Gap junctions are found in the basolateral membrane High‐resolution ultrastructural studies on isolated liver gap
nearby the tight junctions. In contrast to the linear strands of junctions using techniques of X‐ray diffraction and low‐angle

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
14:  Gap and Tight Junctions in Liver: Structure, Function, and Pathology 161

(a) (b) (e)

(c) (d)

(f) (g)

Figure 14.1  Fine structure of appositional membranes between hepatocytes, showing tight junctions and gap junctions. (a, c, f, g) Freeze‐fracture;
(b, d) thin sections; (e) immunofluorescence. (a) Below the bile canaliculi, tight junction webs are seen that seal off the canalicular surfaces from
the appositional membrane, as seen by freeze‐fracture electron micrograph. (b) Thin‐section micrograph illustrates tight junctions (Tj) sealing the
canaliculus (arrowhead) and a large gap junction (Gj) below it. Freeze‐fracture (c) and thin‐section (d) views show larger gap junction plaques
isolated from tight junctional strands. (e) Immunofluorescence of mouse liver showing ZO‐1 (tight junction protein‐1, green) lining the bile
­canaliculi of the liver and the gap junction protein, connexin 32, outside of but in close proximity to the bile canaliculi. White arrows show bile
­canaliculi. Nuclei are shown in blue. (f) Freeze‐fracture micrographs of the adluminal plasma membrane where tight junction strands form well‐
developed networks and small gap junction plaques (g), observed peripherally or within the tight junction network. Bars = 0.1 nm or as indicated.
Modified from [3] with permission of Rockefeller University Press.

scattering provided early details to the model of the gap junction interconnects with the connexon of an apposing cell through an
channel [6]. It is now generally accepted that liver gap junction interdigitated double‐walled extracellular cavity.
channels are composed of 12 subunits, six contributed by each
cell, forming what is termed a hemichannel or connexon
(Figure 14.2). The subunits are radially symmetrical around the
central pore, and each is believed to be tilted slightly relative to MOLECULAR COMPONENTS AND GENES
the plane of the membrane. Cryoelectron microscopy and com­ OF TIGHT AND GAP JUNCTIONS
puter modeling have provided evidence that the formation of rat
liver gap junctions requires a 30 degree rotation between
hemichannels for proper docking [7]. The structure of a gap
Tight junctions
junction channel consisting of connexin 26 in an open confirma­ Rodent liver membrane fractions served as the source for the
tion has been determined to 0.35 nm resolution [8]. This depicts initial isolation of tight junction‐associated proteins. The first of
the Cx26 gap junction connexon as a positively charged hexago­ these was a high molecular weight (~225 kDa) molecule, which
nal funnel that narrows to a 0.14 nm central pore, which then was termed zonula occludens 1 (ZO‐1, also known as tight
162 THE LIVER:  MOLECULAR COMPONENTS AND GENES OF TIGHT AND GAP JUNCTIONS

(a)
Tight junction proteins Gap junction proteins

1 2 3 4 1 2 3 4 1 2 3 4 1 1 2 3 4 1 2 3 4

CLDN1 CLDN2 OCLN JAM1

GJB1 GJB2

ZO1

(b) (c)
Tight junction channel Gap junction channel
(CLDN2)

Cell 1

Cell 2

Figure 14.2  Molecular components and topologies of liver tight and gap junctions. (a) Tight junction proteins are primarily tetraspan claudins
(Cldn1/2) and occludin (OCLN), but single‐pass proteins such as junction adhesion molecule (JAM1) are also components. Gap junction proteins
between healthy hepatocytes are primarily connexins 32 and 26 (Cx32 or GJB1 and Cx26 or GJB2), although Cx43 (GJA1) is also present in some
cells and may be induced following injury. The zonula occludens proteins ZO‐1 (shown) and ZO‐2/3, which do not contain membrane‐spanning
domains, are also prominent in liver and bind to both tight junction and gap junction proteins. Diagrams represent amino acid sequences of human
isoforms from Protter software. (b) Secondary structure of tight and gap junctions reveal contrasting ways in which these proteins form paracellular
and intercellular channels. On left is a conceptual drawing of a tight junction formed of Cldn2, illustrating the permeability pathway between the
cells proposed to be formed by hydrophilic residues in the extracellular beta barrels. To the right are paired hexameric connexons or hemichannels
forming bidirectional permeation pathways connecting interiors of adjacent cells.

junction protein 1, Tjp1), and discovery of the other family major surprise that the Ocln‐null mouse still had similar struc­
members ZO‐2 and ZO‐3 (Tjp2 and 3) quickly followed (see tures [10]. Re‐examining the purified chick membrane fractions
[4]). These proteins were soon realized to be members of the led to discovery of two additional somewhat smaller proteins
membrane‐associated guanylate kinase (MAGUK) superfamily, (~23 kDa), termed claudin‐1 and claudin‐2 (Cldn1/2), that
which contains domains (termed SH‐2, GUK, and PDZ) that are became the founding members of what is now known to be a
specialized for binding to other proteins. Although ZO‐1/2/3 are large family of 24 members in the human genome.
commonly associated with tight junctions, they are actually Although liver tight junctions are mainly composed of Ocln
scaffolding proteins that also bind to protein components of and members of the Cldn family, other core molecular tight junc­
other junctional types, including gap junctions. The first true tion components include the junctional adhesion molecules
core component of tight junctions was identified as a 65 kDa (JAMs), coxsackie adenovirus receptor (CAR), tricellulin, and a
protein in junctional membrane fractions purified from chicken few other identified or suspected molecules [11]. Tight junctions
liver and termed occludin (Ocln) [9]. Ocln was shown to form in various tissues have distinct complements of these core pro­
tight junction‐like webs of intramembrane particles in trans­ teins, which can form both homo‐ and heterophilic interactions.
fected cells and it seemed to satisfy criteria as the universal This diversity reflects in part a tissue‐specific organization, but it
transmembrane protein of tight junctions. It thus came as a also achieves various degrees of “tightness,” a topic considered
14:  Gap and Tight Junctions in Liver: Structure, Function, and Pathology 163

in more detail in subsequent sections dealing with normal and particularly in extracellular and transmembrane domains.
pathological functions. Homologous extracellular domains are thought to account for
The major tight junction components occludin and the claudins the ability of many connexins to form heterotypic channels (i.e.
are tetraspan proteins with intracellular N‐ and C‐termini and cells expressing one type of connexin forming gap junctions
two  extracellular loop domains (Figure  14.2). By contrast, the with cells expressing a different connexin). The existence of
JAMs and CAR are single‐pass membrane proteins. The first heteromeric connexons (i.e. hexameric connexons that contain
loop of Cldns is twice the length of the second loop and its amino different connexin proteins) in the liver was initially suggested
acid sequence influences the paracellular charge selectivity, from biochemistry experiments that fractionated detergent‐solu­
whereas the second extracellular loop acts as a receptor for a bac­ bilized gap junctions [24], and this was supported by the
terial toxin, and the C‐terminus binds cytoplasmic proteins, electrophysiological properties of hepatocytes isolated from
­
including cytoskeletal elements via a PDZ motif. Discrete Cx32‐deficient and wild‐type mice [25], and by studies on
­residues within the first extracellular loop of Cldn1 in liver are reconstituted vesicles in which differences in permeability for
critical for hepatitis C virus (HCV) entry [12, 13]. Like claudins, second messenger molecules were obtained when the ratio of
Ocln possesses intracellularly located PDZ‐binding domains for Cx32 to Cx26 was varied [26]. The greatest dissimilarities
scaffold association proteins such as ZO‐1 that link it to the among connexins occur in the amino acid sequence of the C‐ter­
cytoskeleton. minus, which is the domain thought to play the major role in
Numerous tight junction proteins are expressed in murine regulation of gap junction channels. It is within this cytoplasmic
­livers, including claudin‐1, ‐2, ‐3, ‐5, ‐7, ‐8, ‐12, ‐14, occludin, region of the proteins that most of the potential phosphorylation
JAM‐A, CAR and tricellulin; claudin‐1, ‐2, ‐3 are expressed in sites are found and against which most of the useful connexin‐
the bile canaliculus region of hepatocytes [14, 15]. Claudin‐2 specific antibodies have been raised. Moreover, the connexin
immuno‐localization shows a lobular gradient increasing from C‐terminus contains binding sites for numerous protein p­artners,
periportal to pericentral hepatocytes, whereas claudin‐1 and ‐3 including kinases and structural proteins in a scaffolding
are expressed uniformly throughout the liver lobule. Further­ ­complex that has been termed the Nexus [27]. The tight junc­
more, claudin‐2 expression induces cation‐selective channels in tion‐associated protein ZO‐1 interacts directly with the extreme
tight junctions of epithelial cells, as do claudin‐7 and ‐15 [16]. C‐terminus of Cx43, utilizing the second PDZ domain of ZO‐1,
and Cx32 interacts with the SH‐3 domain of the PDZ‐scaffold­
ing protein discs large homolog 1 (Dlgh 1) in the liver [28, 29].
Gap junctions
Furthermore, small gap junction plaques are associated with
The first gap junction protein isolated from detergent‐­solubilized tight junction strands in certain cell types, including hepatocytes
gap junctions displayed an apparent molecular weight of [5], and Cx32 is partially colocalized with Ocln and Cldn1 in
27 kDa, although a 21 kDa protein was also present (for histori­ hepatocytes [30]. Such protein–protein interactions likely pro­
cal overview, see [4]). Cloning of the cDNA encoding the more vide platforms localizing intracellular signal transduction to
slowly migrating protein [2, 17] indicated that its predicted contacts between the cells. In addition to the binding of
molecular weight is about 32 000, and this protein is generally cytoskeletal proteins to liver connexins, immunoprecipitation
referred to as connexin32 (Cx32). In an alternative nomencla­ pulldown studies have identified a number of mitochondrial
ture this gap junction protein is referred to as beta1 and its gene proteins bound to Cx32, suggesting the possibility of gap junc­
as Gjb1 in rodents and GJB1 in humans [18]. The cDNA encod­ tion involvement in mitochondrial signaling in the liver [31].
ing the 21 kDa protein was subsequently cloned and found to
encode a protein of predicted molecular weight of 26 000 [19].
This connexin is now termed Cx26 (or beta2) and its gene is
Gjb2 or GJB2. FUNCTIONS AND REGULATION
Cx32 and Cx26, the core components of hepatocyte gap junc­ OF TIGHT JUNCTIONS IN THE LIVER
tions, are also found in a variety of other cell types (for review,
see [20]). Moreover, another gap junction protein, Cx43 (or
The barrier: permeability regulation
alpha1; gene Gja1 or GJA1) is prominent between other liver
cell types, including stellate cells, Kupffer cells, and endothelial Tight junctions between hepatocytes provide a high‐resistance
cells (for review, see [21]). The connexin gene family now com­ barrier to leakage of water and solutes into and out of the bile
prises at least 20 proteins in vertebrates. Connexins are absent in canaliculus, termed the blood–biliary barrier. The number of
nonchordate genomes, where gap junction channels are formed tight junctional strands encircling the bile canaliculus is nor­
by a separate gene family encoding the innexin proteins [20]. mally high, indicating that this permeability is low, which is
Although chordate homologs of innexin proteins have been validated by resistance measurements from microelectrodes
identified, only one of these so‐called pannexin (Panx1/2/3) inserted into the canalicular space [32]. As a consequence, this
proteins has been shown to form channels, and the channels that barrier functions to maintain the composition of the canalicular
Panx1 forms appear to be exclusively involved in formation of fluid, allowing the accumulation of high concentrations of bile
nonjunctional, rather than junctional channels [22]. Pannexins acids, phospholipids, and other preferentially secreted organic
are found in the liver, where they have been proposed to play anions within this intercellular compartment.
roles in acetaminophen‐induced injury and other disorders [23]. Approaches commonly used to investigate the barrier func­
Like Ocln and Cldns, connexins are tetraspan membrane pro­ tion of tight junctions involve measurement of the permeability
teins. They exhibit very high homology among family members, of the paracellular pathway to ions and uncharged hydrophilic
164 THE LIVER:  FUNCTIONS AND REGULATION OF TIGHT JUNCTIONS IN THE LIVER

(a) Fence Barrier


(d1) (e)

I1
V1
V2
(d2) –50

100 pA
I2

(b1) 5 sec
(c)
I1

(f)
(b2)

20 pA
I2

2 sec
Figure 14.3  Functions of tight and gap junctions. (a) Function and molecular components of tight junctions. The major tight junction functions
are termed “fence,” where membrane‐embedded molecules are segregated to apical and basolateral domains, and “barrier,” where aqueous mole­
cules in canalicular fluid are excluded from the blood. (b) Fence function images of tight junctions in liver. The hepatocytes are labeled with
BODIPY‐sphingomyelin, which is effectively retained from flowing past the apical domain (b1, arrow). In hepatocytes treated with 3 mM EGTA
for 5 minutes, the probe diffused through the tight junction, labeling the basolateral face (b2, arrow). (c) Barrier function illustrated with thin‐sec­
tion image. Tight junctions flanking the microvillus‐filled canaliculus exclude lanthanum, which had been introduced into the extracellular spaces
via the vascular system and is seen to penetrate the extracellular spaces but is denied access to the canalicular lumen by the tight junctions (arrows).
(d) Gap junctions permit diffusion of molecules up to a size limit of about 1 kDa between cell cytoplasms. Light (d1) and epifluorescence (d2)
photos showing that Lucifer Yellow CH (5% wt/vol) injected iontophoretically into one cell of a pair of hepatocytes passes detectably to a coupled
neighboring cell within 30 seconds. (e) Voltage‐clamp experiment on a pair of hepatoma cells stably transfected with Cx32. A 50 mV hyperpolariza­
tion (V2) of one cell produces an initial current in the other cell (I1), which decreases over a period of seconds. This voltage sensitivity is charac­
teristic of liver gap junction channels. (f) Single gap junction channel openings and closures can be seen in this example of a high‐again recording
from a poorly coupled pair of SKHep1 cells transfected with rat Cx32. In response to a driving force of 50 mV, channels continuously open and
close, as indicated by abrupt transitions of equal size but opposite polarity recorded in each cell’s voltage‐clamp circuit.

macromolecules. Such permeability is visualized using elec­ is quantified through measurements of ratio of proteins in the
tron‐dense (e.g. ruthenium red, lanthanum) or fluorescent dyes two domains. One straightforward assay involves the use of
and can also be quantified as the transepithelial resistance (TER) fluorescent lipids or lipophilic probes (such as BODIPY‐­
measured between the apical and basolateral compartments sphingomyelin). When added to medium bathing specifically
[33]. Determination of the crystal structure of Cldn15 revealed the canalicular or basolateral surface, the lipids insert into the
that the β‐helical extracellular domain forms a dimer that could outer leaflet of the plasma membrane. Restriction of dye to one
provide a paracellular permeation pathway if lined with hydro­ compartment or the other can be compared under conditions in
philic amino acid residues but would otherwise be impermeable which tight junctions are present or disrupted by Ca2+ chelation
to charged molecules [34]. Of particular relevance to the liver, (Figure 14.3b) [30].
Cldn2 has such hydrophilic residues in its extracellular domain,
whereas Cldn1 does not. Thus, Cldn2 provides an ion permea­
tion pathway across the paracellular barrier that is analogous to Regulation of tight junctions
the cell–cell channels connecting cell interiors provided by gap Both barrier and fence functions of tight junctions are highly
junctions (Figure 14.2b). In the liver Cldn2 shows a distribution dependent on the integrity of the cytoskeleton. For example,
that is graded across the lobule. This expression profile is under optimal culture conditions, primary rat hepatocytes dis­
hypothesized to generate directional bile flow from the center to play well‐developed networks of tight junction strands, with
the periphery of the lobule, as evidenced by deficient flow in the circumferential actin filaments near the tight junction regions
Cldn2‐null mouse [35]. [30] (Figure 14.1). The tight junction molecules occludin, clau­
din‐1, ZO‐1, and ZO‐2 are observed by immunofluorescence at
The fence: establishing and maintaining the cell borders, and the fence function of tight junctions in the
cells is well maintained, as evidenced by retention of labeled
cell polarity sphingomyelin (Figure  14.3b). Treatment of the cells with an
Tight junctions also separate the apical from the basolateral cell actin depolymerizing drug (mycalolide B) caused disappear­
surface domains, thereby isolating membrane proteins to one ance of both the circumferential actin filaments and occludin,
region or the other and establishing and maintaining cell polar­ while tight junction strands remained virtually intact. These
ity. This is termed the “fence” function of tight junctions, and it results provide evidence that occludin may be especially critical
14:  Gap and Tight Junctions in Liver: Structure, Function, and Pathology 165

for providing strong linkage between the actin cytoskeleton and of claudin‐1 are also observed, and the lack of claudin‐1 may
tight junctions in hepatocytes [30]. lead to increased paracellular permeability between bile duct
A number of signaling molecules have been associated with epithelial cells [41].
the regulation of tight junctions, including tyrosine kinases, HCV gains access to hepatocytes through binding to co‐
cAMP, Ca2+, protein kinase C (PKC), heteromeric G proteins, receptors on hepatocytes that include extracellular loop domains
and phospholipase C [36, 37]. A protocol of switching between of Cldn1 and Ocln and the tetraspanin membrane protein CD81
Ca2+‐containing and Ca2+‐free solutions has demonstrated that in [12, 13]. The potential utility of targeting claudin tight junction
monolayers incubated in low calcium medium, tight junction proteins to block infection by HCV is highlighted by a report
proteins are disassembled, dephosphorylated, and are less that monoclonal Cldn1 antibodies can clear HCV infection in
tightly associated with actin filaments [30]. This indicates that humanized mice [42]. Other tight junction components may
the function of tight junctions may be locally regulated by sign­ also bind viruses, and HCV appears to target multiple claudin
aling events within the tight junction plaque or may regulate protein family members. Moreover, reovirus coat protein binds
some aspect of intracellular signaling. Cytoplasmic signaling to extracellular JAM‐A homodimerization domains, and CAR is
pathways are believed to directly control the barrier function of related to immunoglobulin junctional adhesion molecules and is
tight junctions. ATP‐depletion experiments result in altered localized to junctional domains.
tight junction structure, decreased TER, and an increased asso­
ciation of tight junction proteins with the actin filaments [38].
The PKC‐activating phorbol esters induce a rapid decrease in
tight junction permeability and alter perijunctional actomyosin FUNCTIONS OF GAP JUNCTIONS
[39]. Viral and bacterial pathogens have exploited core tight IN THE LIVER
junction proteins and their adaptors to disrupt the paracellular
barrier, and these strategies are in many examples associated As described above, hepatocytes express high levels of Cx32
with a reorganization of perijunctional actin [13]. and somewhat lower levels of Cx26. Kupffer cells, stellate cells,
cells of Glisson’s capsule, cholangiocytes, and sinusoidal
endothelial cells all express Cx43. Interestingly, chronic or
Tight junction disease acute liver injury can induce Cx43 expression in hepatocytes but
Hepatocyte tight junctions are believed to provide the primary the significance of this is unclear [43]. Endothelium of the por­
intercellular barrier between the sinusoidal and the canalicular tal vein and arteries, however, express Cx37 and Cx40, which
space. This barrier, often referred to as the blood–bile barrier, are commonly expressed in endothelia and smooth muscle.
is compromised under experimentally induced liver injury in Hepatocytes and Kupffer cells also express hemichannel‐form­
mice, with accompanying impaired bile secretion and jaun­ ing Panx1. As mentioned above, these are related to the inverte­
dice [40]. In tissue thin sections, the tight junctions of hepato­ brate gap junction proteins innexins, but in the chordate animal
cytes can be seen to block access of the experimentally phylum (which includes vertebrates) it appears that gap junction
administered extracellular tracer (lanthanum) to the canalicu­ function has been taken over by connexins and that pannexins
lar lumen (Figure  14.3c). Irregularities in the structure and do not form communicating channels between cells [21, 22].
distribution of tight junctions accompanied by increased Gap junctions closely align with the functional structure of
blood–bile barrier permeability have been observed in experi­ hepatocytes and therefore to the functional structure of the
mental mouse models including extrahepatic cholestasis after entire liver. Gap junctions form in close association with tight
common bile duct ligation, intrahepatic choleostasis after junctions and adherens junctions, and these all come together to
ethinylestradiol treatment, partial hepatectomy, choline‐defi­ organize the apical membrane of the hepatocyte, the bile canali­
cient diet, hepatocyte injury following hepatotoxic drug culus (Figure  14.1). As is seen elsewhere in this volume, bile
administration, and experimental colitis [30, 40]. Loss of gap canaliculi take the form of narrow tubules that cross the lateral
junctions, leaky tight junctions, and disorganized actin bun­ surface of hepatocytes. They contain microvilli on their luminal
dles are considered to be sufficient to cause cholestasis and the surface that provide high surface area to the crowded bile cana­
eventual development of jaundice. With advances in studies of licular interior. Bile canaliculi give a characteristic chickenwire
tight junctions, it has become clear that these three causes are appearance across the liver acini when seen by fluorescence
never independent of each other. Disruption of actin filaments microscopy (Figures 14.1 and 14.4). Cx32 and Cx26 are seen to
clearly deteriorates barrier function of tight junctions, and the align along this chickenwire, forming irregular disk‐shaped
interactions among tight junction proteins has not been fully plaques, rather than tubules. Since gap junctions are associated
clarified. Tight junction proteins have also been shown to with the apical junctions of hepatocytes [5], injuries or disease
mediate entry of numerous hepatotropic viruses, and func­ that can cause loss of hepatocyte polarity or loss of functional
tional and structural abnormalities of tight junctions are impli­ organization of the hepatic acinus can also alter gap junction
cated in liver cancer. localization and expression [21].
Regarding genetic diseases of human tight junction proteins,
missense mutations in ZO‐2 have been identified in patients
Zonation of the liver
with familial hypercholanemia (FHCA: MIM phenotype 607748)
and in progressive familial intrahepatic cholestasis 4 (PFIC4: Gap junctions couple cells electrically by allowing ions to pass
615878). In a syndrome associating ichthyosis and neonatal between the channels such that a change membrane potential
sclerosing cholangitis (NISCH syndrome: 607626), mutations experienced by a cell can be shared by its neighbor through a
166 THE LIVER:  FUNCTIONS OF GAP JUNCTIONS IN THE LIVER

Hepatic lobules

P.triad

P.triad

Figure 14.4  Exposure to toxic doses of acetaminophen leads to centrilobular necrosis and increased expression of connexin 43. Immunofluorescence
microscopy images of connexins 32 and 43 in C57BL6 mouse livers treated with 500 mg kg−1 acetaminophen (APAP) or control solvent alone (Ctl).
Twenty‐four hours after IP administration of APAP, livers were removed, frozen and subsequently sliced and co‐stained with antibodies to Cx32
and Cx43 followed by fluorescent secondary antibodies to allow visualization of both proteins in the same microscopy fields. Control livers (top
panels) show punctate plaques of Cx32 (left panels) that dot the borders of the hepatocytes in a chickenwire‐like appearance, while Cx43 (right
panels) staining is minimal except in the fibroblast cells of Glisson’s capsule (white arrow). APAP‐treated livers (lower panels) show autofluores­
cent necrotic cells (nec.) and punctate staining of both Cx32 and Cx43 throughout the liver. The cartoon at the right depicts the lobular organization
of the liver along with its metabolic gradient zones (1, 2, 3) and portal triads (P. Triad), central veins (CV), and centrilobular necrosis (nec.) that
occurs with APAP toxicity. The approximate locations of these zones of the liver are also indicated in the lower panels.

flow of ions. In neurons, gap junctions are referred to as elec­ zonation, with stronger staining in the periportal region (zone
trical synapses, and they function to transmit impulses and syn­ 1), but the significance of this is still unclear [46]. Zonal locali­
chronize the cells. The physiological consequences of high zation of metabolic enzymes and the activity of gap junctions
levels of gap junction expression that is seen in hepatocytes presents a conceptual dilemma; if substrates can freely diffuse
includes not only ionic coupling, but also transfer of cellular through the liver via gap junctions then why would it be benefi­
signals, especially those involved in growth control and metab­ cial to localize enzymes to specific zones? There may be
olism. Relaying of metabolic signals is important not only numerous answers to this question depending the enzymes
because of the anatomical arrangement of the hepatic lobule being examined, but one explanation is that the substrates and
but also because many physiologic functions show gradations various molecules can still form gradients within the acinus
along the lobule, from the periportal region (containing the even though they can diffuse between cells. For instance, it has
portal triad) to pericentral regions (containing the terminal been shown that changing the direction of flow of an isolated
hepatic venule). That is, the liver contains functional metabolic perfused liver will change the zonal location of gluconeogen­
zones 1, 2, and 3 that reflect the level of oxygen and the flow esis [47]. Nevertheless, the strong electrical coupling mediated
blood through the sinusoid from perportal region to the central by gap junctions serves to equilibrate the membrane potential
vein (Figure  14.4). Interestingly, bile flows in the opposite among hepatocytes. In isolated rat livers, perfusion with gluca­
direction. Many metabolic enzymes exhibit ­asymmetric zona­ gon leads to hyperpolarization of all hepatocytes across the
tion. For example, hepatocytes with higher gluconeogenic hepatic acinus. However, in the presence of octanol, a gap
activity are found periportally rather than pericentrally [44, junction blocker, glucagon‐induced hyperpolarization is higher
45], and glutamine synthetase, which removes ammonia from in periportal hepatocytes [48] Furthermore, heptanol blockade
blood, is strictly localized to hepatocytes surrounding the cen­ of gap junctions has been shown to abolish metabolic and
tral vein (zone 3). Connexin 26 also exhibits asymmetric hemodynamic effects of nerve stimulation within the liver [49].
14:  Gap and Tight Junctions in Liver: Structure, Function, and Pathology 167

Calcium signaling two thirds of the liver is surgically removed, gap junctions in the
recovering liver undergo a strikingly coordinated cycle of disap­
Calcium is a major intracellular signaling molecule and the pearance and reappearance over a time course of 48 hours (for
­permeability of Ca2+ ions and the second messenger, inositol reviews see [5, 58]). For the first 20–24 hours there is no change
trisphosphate (IP3), has important roles in coordinating the
­ in the gap junctions, but by 26 hours most hepatocytes have lost
response of hepatocytes to environmental fluctuations. Imaging these junctions. By 36 hours, gap junctions begin to reappear
studies on hepatocyte pairs has shown that injection of Ca2+ into and then fully recover by 48 hours. The disappearance corre­
one cell leads to a rise in Ca2+ in the adjacent cell [50]. This sponds to a peak in mitosis. Electrical measurements indicate
intercellular spread can be blocked by treatment of the cells that junctional conductance decreases and increases with a simi­
with the gap junction inhibitor heptanol, indicating that move­ lar time course, although increased nonjunctional resistance
ment is by way of gap junction channels. Although both IP3 and contributes to the coupling measurements. Using immunoblot
Ca2+ ions can pass through gap junctions, most models indicate techniques, Cx32 was also found to undergo a similar cycle of
that IP3 is the primary propagating agent. When injected into a decrease and reappearance, with the decrease coinciding with a
cell, IP3 increases intracellular Ca2+ in injected cells and their time of maximal DNA synthesis. However, Cx26 mRNA was
coupled partners. Because the Ca2+ rise in the recipient cell can found to be selectively increased before the onset of S‐phase
occur far from the junctional region, and because the traveling after partial hepatectomy as well as after bile duct ligation.
waves are very fast, diffusion of Ca2+ itself cannot account for its Freshly isolated primary hepatocytes, which generally do not
propagation. In fact, Ca2+ is strongly buffered by the intracellu­ undergo cell division, also undergo changes in the expression of
lar environment and cannot easily diffuse. Instead the system gap junction proteins and electrical coupling that are similar to
relies on signaling molecules such as IP3 and its metabolites those that are seen following tissue injury in situ; gap junctions
[51] that can trigger rapid release in distal areas. Intercellular initially disappear and then reappear over a time course of days
Ca2+ signaling is commonly used by cells to coordinate and [59].The disappearance of gap junctions in hepatocyte cell cul­
regulate a wide range of cellular functions including cell growth ture can be delayed by treatment with cAMP or glucagon or
and differentiation [52]. Ca2+ waves also contribute to the vari­ with protein synthesis inhibitors, and connexin mRNA stability
ous functions of liver associated with vasopressin and norepi­ is increased under these conditions. These findings indicate that
nephrine stimulation [53], and increased Ca2+ induces bile that reduction in gap junction expression and function correlates
canalicular contraction and secretion of bile by hepatocytes and with cellular reprogramming and is permissive for proliferation
cholangiocytes [54]. Gap junction‐mediated regenerative waves in the liver, though not absolutely required. In this regard it is
of elevated Ca2+ appear to provide long‐range signaling for the also possible that loss of gap junctions could provide a selective
propagation of these contractions. advantage for preneoplastic liver cells as they develop into rap­
In addition to transmission through gap junction channels, idly proliferating tumor cells.
Ca2+ waves may also propagate via rises in extracellular nucleo­
tides that are detected by adjacent and non‐adjacent cells. This
latter mechanism utilizes the activation of purinergic receptors Regulation of coupling between hepatocytes
by ATP and related nucleotides that are released from the stimu­ The extent to which signals spread in the liver depends on the
lated cells at least partly through Panx1 channels [55]. This type number of open gap junction channels. Gap junction channels
of signaling can be inhibited by purinergic receptor antagonists between hepatocytes are closed by, among other things, cyto­
or apyrase, and hepatocytes can be desensitized to this signaling plasmic acidification, cellular injury, high extracellular Ca2+,
by inclusion of ATP itself [56]. Cholangiocytes also utilize CC14, certain alcohols, and large voltage gradients (for reviews,
purinergic signaling for bile secretion, and this appears to be see [2, 60]).
activated by ATP released from within bile [54]. Intracellular
Ca2+ signaling within the liver represents a confluence for
responses to changes in the environment, and gap junctions as Phosphorylation of gap junctions
well as purinergic signaling participate in how these responses Another important avenue of gap junction regulation is phos­
are propagated. phorylation. All connexins, with the exception of Cx26, appear
to be substrates for phosphorylation, and Cx43 has been shown
to be phosphorylated at 16 different sites [57]. Connexin 32 is
Growth control
phosphorylated in its cytoplasmic tail at Ser233 by cAMP‐
Control over the growth of tissues has long been proposed as an dependent protein kinase (PKA), and this correlates with
important function of gap junctions. Evidence supporting this increased conductance [61]. Ca2+‐calmodulin‐dependent protein
association include loss of functional coupling in liver tumor kinase II and protein kinase C can also phosphorylate Cx32, and
cells, reduced gap junction expression in hepatomas compared these can produce tryptic digestion patterns that are distinct
with that in adjacent normal tissue, reduced rate of tumor growth from those obtained with cAMP‐dependent protein kinase.
in cells transfected with connexins, and loss of gap junction Phosphorylation by protein kinase C appears to inhibit some
communication in response to chemical tumor promoters [57]. types of proteolysis [61]. Although in other cell types inhibition
Models that have been useful for understanding the association of junctional communication has been associated with the
between gap junction expression and growth control in liver expression of tyrosine kinases, Cx32 in isolated rat liver gap
include the regeneration of liver following surgical or chemical junctions is not phosphorylated by purified pp60v‐src tyrosine
insults. Following partial hepatectomy, whereby approximately kinase or the insulin receptor.
168 THE LIVER:  FUNCTIONS OF GAP JUNCTIONS IN THE LIVER

Oligomerization of gap junctions fibroblasts from Cx43‐knockout mice show aberrant cell
­polarity and reduced wound closure, and transfection with dom­
As with other transmembrane proteins, gap junction proteins are inant‐negative Cx43 that lacks a putative microtubule‐binding
co‐translationally inserted into the endoplasmic reticulum (ER). domain reconstitutes this phenotype [65]. Interestingly, in
The nascent connexin protein must fold correctly and oli­ hepatocytes vesicle delivery of biosynthetic components such
gomerize with five other subunits to form a hexameric complex, as lipoprotein occurs at the sinusoidal face, near to plus ends of
the connexon. Connexins have their N‐ and C‐termini located microtubules, while gap junction plaques are found in close
on the cytoplasmic face and recently, compatibility motifs have proximity to the apical domain where microtubule minus ends
been identified that roughly predict which connexins can het­ are abundant. Liver injury or other conditions that can disrupt
ero‐oligomerize. The compatibility motifs were identified by hepatocyte polarity can induce expression of Cx43 (Figure 14.4),
amino acid sequence homology clustering, and allow categori­ and the above findings suggest that Cx43 may take advantage of
zation of connexins as either R (arginine) or W (tryptophan) or altered cell polarity during such situations.
other, based on the amino acids that are present. The motif is Microfilaments form another important cytoskeletal compo­
found at the junction between the cytosolic intracellular loop nent for junctional proteins. The filamentous actin network is
and the third transmembrane domain. Connexins 26 and 32 are integral to the structure of cell junctions and the shape, polarity,
W types, whereas Cx43 is R type. Cx43 is unusual in that its and contractile activity of cells in general. In hepatocytes, bile
monomer form is stabilized in the ER, whereas Cx32 and Cx26 canaliculi are frequently identified by the abundance of actin
appear to oligomerize within the ER. These form hetero‐ that surrounds them. For delivery of junctional proteins, actin is
oligomers as is predicted by their compatibility motif. unlikely to serve as a railroad track, since actin filaments are
Oligomerization and trafficking through the various compart­ much shorter than microtubules and frequently form an iso­
ments in the cell appear to be assisted by chaperones, and the tropic network. However, actin may participate in anchoring
regulation of oligomerization in the ER and Golgi apparatus vesicles containing connexins to particular regions of the cell
suggests the presence of patrolling complexes that prevent inter‐ and they may provide a directional bias to allow vesicles and
connexin association [62]. various macromolecules to form contacts and, for example, pro­
mote fusion of vesicles with the plasma membrane. Additionally,
Trafficking of and cytoskeletal interactions there is considerable literature linking gap junction protein
of gap junctions expression with the regulation of cell motility, cell adhesion,
and migration. These studies have primarily focused on Cx43,
Gap junction trafficking pathways beyond the ER and Golgi which has an unusually long C‐terminal tail, but recent studies
compartments are not clearly defined and it is plausible that also implicate Cx32 and Cx26 in cell migration [66]. Neurons
connexons made up of different connexin subunits use alternate from developing mouse brain as well as embryonic fibroblasts
routes. Growing evidence is solidifying the view that microtu­ from Cx43‐knockout animals exhibit reduced and disorgan­
bules play a major role in regulation of connexin trafficking ized cell migration, whereas overexpression of Cx43 has led
[63]. The dependence of gap junction regulation on cytoskeletal to  enhanced migration of multiple cell types in both wound
components such as microtubules and microfilaments presents a healing and transwell migration assays. Short hairpin RNA
mechanism for directed trafficking within the cell. In most cell (shRNA)‐mediated knockdown of Cx43 has also led to disor­
types, microtubules stretch across the length of the cell and form ganized cell polarity and reduced migration. Although the
railroad tracks for the movement of organelles and macromole­ mechanisms by which connexins may affect cell polarity and
cules that may otherwise be trapped by the crowded nature of motility are not understood, it has been suggested that Cx43 and
cytoplasm. In hepatocytes, like in other polarized cell types, other connexins can act as dynamic scaffolding proteins that
microtubules extend from the microtubule organizing center as organize and attract actin‐modifying proteins to cell–cell junc­
well as from the apical domain itself. Therefore, the slow‐ tions and to sites of cell adhesion and protrusion [67]. As indi­
growing minus ends of microtubules are found near the bile cated earlier, Cx43 has also been shown to bind microtubules,
canaliculus and the fast‐growing plus ends are found near the and the actin‐organizing and microtubule‐binding activities may
basolateral plasma membrane, although some microtubules are work together to control cell shape and movement. In this regard
found in other orientations. Studies from our group have confirmed it is interesting that adult hepatocytes, which are typically non‐
that Cx32‐containing vesicles isolated from liver and within migratory and do not express Cx43, can be induced to express
cultured cells traverse along microtubules in the plus end direc­ this connexin during tissue injury, whereas migratory cells of
tion using the microtubule motor kinesin 1 [64]. This suggests the liver, such as Kupffer and stellate cells, express Cx43.
that connexins are delivered to the basolateral membrane, where Once within the plasma membrane, connexons are believed
they form large gap junction plaques that localize to the borders to diffuse laterally and dock with opposing connexons from
of the apical–basolateral domains (see Figure 14.1). Connexins, others cells. Hepatocytes form gap junctions with all neighbor­
and Cx43 in particular have been shown to not only bind micro­ ing hepatocytes, making approximately six cell–cell contacts.
tubules, but to traffic to adherens junction‐associated mem­ It is unclear whether this process is active or passive; however,
branes containing N‐cadherin and beta‐catenin. Knockdown of cell adhesion molecules are known to play a role in the forma­
EB1 protein has been shown to reduce gap junction plaque for­ tion of gap junctions. Studies using tetracysteine and GFP‐
mation, and EB1 may function to deliver connexins and other tagged Cx43 have demonstrated that the insertion to membrane
junctional proteins to the plasma membrane through its micro­ plaques occurs laterally onto the edges, whereas removal takes
tubule plus end tracking activity [63]. Furthermore, embryonic place from the center of the plaque [68, 69] in a process that
14:  Gap and Tight Junctions in Liver: Structure, Function, and Pathology 169

may involve clathrin/actin‐mediated endocytosis. Because serum‐free medium and found a markedly higher proliferation
docked connexons cannot be separated by most physiological rate of Cx32‐KO hepatocytes [78]. Interestingly, however, the
conditions, junctional removal appears to be a phagocytic‐ proliferation rate of Cx32‐KO hepatocytes after partial hepatec­
like  event which involves the formation of double membrane tomy was significantly lower [79]. Experiments on transgenic
vesicles, termed annular junctions. It is unclear whether this is mice expressing a dominant‐negative mutant of Cx32 have
the only form of internalization for gap junctions since annu­ found delayed liver regeneration and an increase of susceptibil­
lar membrane profiles are not commonly observed in normal ity to chemical hepatocarcinogen [80]. Despite all of this, non­
liver [5]. functioning Cx32 in CMTX patients is not associated with liver
The half‐lives of connexins are unusually short for trans­ abnormalities. The lack of reported liver complications in
membrane proteins, with measurements of 3–6 hours for mouse CMTX patients could be due to greater innervation of human
liver connexins [2]. It has been shown that internalized gap liver compared to that of the mouse, which perhaps overcomes
junctions are degraded by lysosomes and proteasomes, and via some of the functions of gap junctions, but it is also unclear
autophagic processes [68]. Connexins have also been found to whether liver cancer rates or other properties have been meas­
inhibit autophagy, whereas nutrient starvation can promote ured in these patients.
release of this inhibition resulting in degradation of connexins
and further upregulation of autophagy [70].
Connexin deficiency and liver cancer
Studies using Cx32‐KO mice have supported the notion that
intact gap junction cellular communication functions to sup­
PATHOLOGY ASSOCIATED WITH GAP press tumors and that loss of gap junction communication can
JUNCTIONS AND TIGHT JUNCTIONS lead to tumors [21]. Aged Cx32‐KO mice have also been shown
IN LIVER to have more spontaneous liver tumors in male animals [81]. As
we have noted, hepatocellular carcinoma is typically accompa­
Junctional proteins within the liver are critical for proper organ nied by reduced gap junction activity and loss of Cx26 or accu­
development and maintenance of biological function and over­ mulation of Cx32 in the cytosol rather than plasma membrane
all health. Genetic deficiencies in gap junction proteins can can increase the cancer potential of liver cells [22]. Connexins
result in deafness, neuropathy, cataracts and other eye disorders, themselves can become cytoplasmically localized instead of
and heart, skin, and connective tissue disease [71]. However, the forming cell–cell junctions and there is evidence that cytoplas­
liver does not appear to absolutely require communication mic localization and downregulation may promote hepatocel­
through gap junctions in the adult. Although, as described ear­ lular carcinoma, invasion, and metastasis [82]. Furthermore, in
lier in this chapter, hepatocytes of the liver primarily express important early studies on gap junction function in liver, loss of
connexins Cx32 and Cx26, they also express pannexins [20]. intercellular coupling was shown to be a hallmark of cancer
Pannexins have only been observed to form hemichannels, cells [57]. Induction of connexin expression in many cancer
which link the cytosol to the cell exterior rather than another cell lines has been shown to reduce cell growth and invasion
cell, but they may still allow indirect cell–cell communication phenotypes and regulate epithelial‐to‐mesenchymal transition
through the extracellular space. [57]. Mouse models of cancer have shown variable effects. In
some cases it appears that gap junctions can promote tumori­
genesis by increasing blood vessel growth and distributing
Genetic animal models of connexin deficiency
chemokines [57], while many studies now indicate that
Gap junction beta 1 (GJB1) is notable as a genetic target for increased gap junction expression in late‐stage tumors can
mutations that cause X‐linked Charcot–Marie–Tooth disease ­promote metastatic features.
(CMTX). GJB1 encodes Cx32 and more than 300 distinct muta­ Non‐gap junction activities are also thought to play a role in
tions of Cx32 have been found to cause CMTX [72]. In an possible tumor‐promoting phenotypes of connexins. When
attempt to mimic CMTX disease in a rodent model, Cx32‐­ unpaired, connexins form hemichannels at the cell surface that
deficient knockout mice (Cx32‐KO) were generated through can allow distribution of cell contents into the extracellular
homologous recombination [73]. Although these mice demon­ space. Under normal conditions, hemichannels are believed to
strated progressive demyelination typical of the human disease, remain closed. However, under situations of decreased extracel­
they also showed deficiencies that have not been observed in lular Ca2+ or pHi, hemichannels may open and for instance
CMTX patients. Liver function was compromised, as evidenced release ATP and other signaling molecules that can promote
by reduced glucose release from liver in response to s­ ympathetic inflammation, which in turn can promote carcinogenesis.
nerve stimulation or hormone stimulation [74]. Cx26 levels Gap junctions exchange small molecules between cells and
were decreased, as were the areas of gap junction plaques; therefore are expected to contribute to cancer treatments that
­electrophysiological conductance and permeability to IP3 was involve small molecules or any soluble effectors that can pass
also decreased compared to wild type [75]. Additionally, the through gap junctions (i.e. that have a molecular weight of
proliferation rate of Cx32‐KO hepatocytes in vivo was high, and approximately 1500 Da or less). This includes ions, nucleotides,
spontaneous and chemically induced liver tumors were more metabolites, bile acids, and even peptides and miRNAs.
prevalent [76], and the mice showed increased carcinogenesis However, the effect of gap junctions toward cell survival to toxic
following X‐ray radiation [77]. Our laboratory compared cul­ small molecules is complicated by the “kiss of death, kiss of
tured hepatocytes from Cx32‐KO and wild‐type mice using life” conundrum, sometimes referred to as the “bystander”
170 THE LIVER:  PATHOLOGY ASSOCIATED WITH GAP JUNCTIONS AND TIGHT JUNCTIONS IN LIVER

versus “Good Samaritan” effect [83]: gap junctions might serve to study and likely dependent on multiple converging variables,
to distribute toxins such that all cells are affected, or they might whereas the study of intrinsically toxic drugs has been crucial in
serve to distribute toxins such that the toxin is no longer c­ ritically understanding how liver toxicity unfolds. Some of the most stud­
damaging. Additionally, gap junctions can distribute protective ied intrinsically toxic drugs in rodents are acetaminophen,
antioxidants, such as glutathione, that can detoxify cells. In the thioacetamide, carbon tetrachloride, dimethylnitrosamine, and
liver, for instance, a toxic drug might distribute through a hepatic d‐galactosamine, with acetaminophen being the most prominent
cord in such a way as to allow cytochrome P450 or conjugating and directly clinically relevant [86]. Acetaminophen is nontoxic
enzymes to modify the drug and allow it to be excreted or fur­ at moderate dose but severely toxic and potentially life threaten­
ther metabolized. Some studies have shown that gap junction ing at high dose (e.g. >6 g in adults). Like many drugs, acetami­
inhibition can block paracrine signaling between cancer cells nophen is metabolized by the liver where it reacts with
and non‐cancer cells, and mouse liver and lung metastases have cytochromes and is conjugated to glutathione, sulfate, and glucu­
been reduced and survival prolonged following treatment with ronic acid, allowing for further metabolism and excretion.
the gap junction inhibitor oleamide in combination with an anti‐ However, at high doses glutathione becomes depleted, and a toxic
VEGF antibody [84]. reactive metabolite, N‐acetyl‐p‐benzoquinoneimine (NAPQI),
It remains an important question whether therapeutics to pro­ accumulates and can instigate a chain reaction of hepatocyte
mote or inhibit gap junction function could be beneficial for the necrosis that initiates around central veins of liver lobules
treatment of cancers, including liver cancers. It appears that (Figure 14.4). Notably, cytochromes CYP2E1 and CYP3A4 that
such treatments must take into effect the complex role that gap catalyze the conversion of acetaminophen into NAPQI also
junctions are now thought to play in the cancer process. strongly localize around central veins. The propagating cycle of
Historically, gap junctions have been considered tumor suppres­ necrosis involves reactive oxygen species (ROS), mitochondrial
sors. Changes in gap junctions during fibrosis progression have oxidative stress and dysfunction, activation of Jun N‐terminal
also been well documented in the liver. Hepatic stellate cells kinase, and an inflammatory response [21, 86].
(also known as Ito cells) play a major role in fibrosis of the liver Gap junctions and gap junction proteins are expected to con­
under many settings and a reduction in their activation and tribute to this in multiple ways. For instance, although APAP
growth in the liver represents a major goal in the prevention of can enter cells by diffusion, gap junction channels could distrib­
liver fibrosis. During stellate cell activation the level of Cx43 ute cellular signals, ions, and critical molecules, and hemichan­
increases on these cells as well as on hepatocytes and blockage nels could extrude inflammatory signals such as ATP and other
of gap junction and hemichannel signaling can reduce fibrosis in and damage‐associated molecular pattern molecules. On the
mice [85]. However, induced expression of Cx43 in hepatocytes other hand, gap junctions could also distribute detoxifying com­
appears to be beneficial during liver injury and its inhibition can pounds such as glutathione or UDP‐glucuronic acid to hepato­
be detrimental [21]. Also interestingly, there is a significant cytes where these may be depleted. A critical step in the
interaction of connexin proteins with mitochondria. In models progression of toxicity appears to be the spreading of centri­
of ischemic injury in the heart and brain, Cx43 has been found lobular necrosis to the rest of the liver. In 2004 Asamoto and
associated with mitochondria and provides protection during colleagues [87] described the generation of transgenic rats that
reperfusion. Cx32 has also been found associated with rat liver express hepatocyte‐specific dominant‐negative Cx32 under
mitochondria and may provide a tethering link between the control of the albumin promoter. This resulted in decreased gap
hepatocyte plasma membrane and mitochondria [31]. It there­ junction activity, reduced Cx32 and Cx26 membrane localiza­
fore appears that understanding the precise function of gap junc­ tion, and resistance to the hepatic toxins carbon tetrachloride
tions in specific cells and at specific times during injury will be and d‐galactosamine. A second study then found that these
critical for potentially alleviating liver injury. With regard to rats were resistant to acetaminophen toxicity, showing lowered
uncontrolled cell growth and cancer, gap junction communica­ serum aminotransferase levels and improved liver histology
tion has for many years been viewed as limiting and tumor sup­ compared to wild type [88]. In addition, it was found that aceta­
pressive. This view is now being replaced by more complex minophen‐induced cell death in isolated hepatocytes from
models where gap junctions may be tumor suppressors in early Cx32‐KO mice was no longer synchronized as it is in wild type.
stages of cancer but can become tumor promoters at later stages. Connexin expression was also required to allow cell‐to‐cell pro­
tection of coupled hepatocytes from female to male animals.
Another study observed that Cx32‐KO mice themselves were
The role of gap junctions in drug‐induced resistant to acetaminophen and thioacetamide toxicity as
observed by aminotransferase levels, histology, and death of the
liver injury
animals. Toxicity could also be reduced by the gap junction‐
Drug‐induced liver injury is the most common cause of acute blocking drug 2‐aminoethoxy‐diphenyl‐borate (2‐APB) [89].
liver failure in the United States and a large portion of the cases Although these studies appeared compelling, a number of
involve the widely available pain and fever‐reducing drug aceta­ doubts have been raised. As described earlier in the chapter,
minophen (acetyl‐para‐aminophenol or APAP). Liver injury is Cx32‐KO mice display increased levels of chemically induced
also a primary concern in drug development and a principal rea­ liver tumors, suggesting that they have increased susceptibility
son for the removal of drugs from the clinical pipeline. Drugs that to drugs. Another study indicated that Cx32‐KO mice have
cause harm to the liver are described as either intrinsically toxic, reduced levels of centrilobular glutathione, and that they are in
with predictable dose‐dependent effects, or idiopathically toxic, fact more susceptible to acetaminophen. Maes and colleagues,
with rare and unpredictable effects. Idiopathic drugs are difficult who have studied acetaminophen toxicity extensively, weighed
14:  Gap and Tight Junctions in Liver: Structure, Function, and Pathology 171

in on these issues by performing their own studies. They found form the apical–basolateral junction, it is possible that this loca­
that Cx32‐KO mice have similar levels of cell death, inflamma­ tion is a protective niche for viruses or that viruses can exploit
tion, and oxidative stress in response to acetaminophen, with temporary disruptions of the blood–biliary barrier for their
somewhat lower levels of protein adduct formation [90]. The entry. Future studies are likely to reveal new functions of
gap junction blocker 2‐APB was also shown to exert protective connexins, occludin, and claudins that affect the processes
­
effects through inhibition of CYP450s and c‐Jun N‐terminal of  viral entry as well as downstream signaling and eventual
kinase. Follow‐up reports by some of these investigators have induction of hepatitis.
now found that connexin and pannexin hemichannels may be
important, as specific inhibition of hemichannels provided pro­
tection from drug‐induced liver injury [91]. Collectively, these
studies indicate that all of the contributors to acetaminophen REFERENCES
toxicity are not well resolved. Many technical details can influ­
1. Kojima T., Sawada, N., Yamaguchi H., Fort A.G., and Spray D.C. Gap and
ence the results, such as the solubilizing agent DMSO, and the tight junctions in liver: composition, regulation, and function, in The Liver:
species of mouse or rodent; for instance, rats are less susceptible Biology and Pathobiology, 5th edn (eds. I.M. Arias et al.), Wiley‐Blackwell,
than mice [92]. Connexin proteins and the activity of gap junc­ Chichester, 2009, pp. 201–20.
tions (and also Panx1) appear to feature in these processes, but 2. Spray, D.C., Saez, J.C., Herzberg, E.L. et al. Gap junctions in liver: composi­
at present a clear protective or damaging role in this toxicity tion, function and regulation, in The Liver: Biology and Pathobiology, 3rd
edn (eds. I.M. Arias et al.), Raven Press, New York, 1994, pp. 951–67.
cannot be assumed and it remains to be seen whether gap block­ 3. Spray, D.C., Ginzberg, R.D., Morales, E.A., Gatmaitan, Z., and Arias, I.M.
age can be useful in the clinical setting. Electrophysiological properties of gap junctions between dissociated pairs of
rat hepatocytes. J Cell Biol, 1986;103(1):135–44.
4. Franke, W.W. Discovering the molecular components of intercellular
­junctions  –  a historical view. Cold Spring Harb Perspect Biol, 2009;1(3):
a003061.
PROSPECTS AND PERSPECTIVES 5. Revel, J.P., Yancey, S.B., and Meyer, D.J. Cellular architecture and intercel­
lular communication in rat liver. JAMA, 1981;245(9):958–9.
The two most prominent junctional types in liver are the gap 6. Unwin, P.N. and Ennis, P.D. Two configurations of a channel‐forming mem­
junctions, which provide direct intercellular communication, brane protein. Nature, 1984;307(5952):609–13.
7. Perkins, G.A., Goodenough, D.A., and Sosinsky, G.E. Formation of the gap
and the tight junctions, which serve to partition membrane
junction intercellular channel requires a 30 degree rotation for interdigitating
domains of individual cells and to occlude extracellular space, two apposing connexons. J Mol Biol, 1998;277(2):171–7.
restricting pericellular diffusion. The integral membrane pro­ 8. Maeda, S., Nakagawa, S., Suga, M. et al. Structure of the connexin 26 gap
teins of gap junctions, the connexins, have been structurally junction channel at 3.5 A resolution. Nature, 2009;458(7238):597–602.
well characterized, but their newly recognized roles in binding 9. Furuse, M., Hirase, T., Itoh, M. et  al. Occludin: a novel integral mem­
brane  protein localizing at tight junctions. J Cell Biol, 1993;123(6 Pt 2):
to and organizing cytoplasmic proteins suggest that they may 1777–88.
form nucleation sites for intracellular as well as extracellular 10. Saitou, M., Fujimoto, K., Doi, Y. et al. Occludin‐deficient embryonic stem
signaling. Connexins interact with actin and microtubules as cells can differentiate into polarized epithelial cells bearing tight junctions.
well as mitochondria, and they participate in recovery from J Cell Biol, 1998;141(2):397–408.
injury in many tissues, although their precise role as both gap 11. Chiba, H., Osanai, M., Murata, M. et al. Transmembrane proteins of tight
junctions. Biochim Biophys Acta, 2008;1778(3):588–600.
junctions and hemichannels (where Panx1 may also contribute) 12. Evans, M.J., von Hahn, T., Tscherne, D.M. et al. Claudin‐1 is a hepatitis C virus
in drug‐induced liver injury remains enigmatic. By contrast, co‐receptor required for a late step in entry. Nature, 2007;446(7137):801–5.
peripheral tight junction proteins were identified prior to the 13. Guttman, J.A. and Finlay, B.B. Tight junctions as targets of infectious
core proteins, and one area of intense interest is the diversity of agents. Biochim Biophys Acta, 2009;1788(4):832–41.
14. Rahner, C., Mitic, L.L., and Anderson, J.M. Heterogeneity in expression and
“tightness” profiles provided by the large claudin family of
subcellular localization of claudins 2, 3, 4, and 5 in the rat liver, pancreas,
integral membrane proteins. Although gap and tight junctions and gut. Gastroenterology, 2001;120(2):411–22.
perform different functions, there are numerous points at which 15. Yamamoto, T., Kojima, T., Murata, M. et al. p38 MAP‐kinase regulates func­
these junctional types appear to overlap. Indeed, the findings tion of gap and tight junctions during regeneration of rat hepatocytes.
that the traditionally tight junction‐associated protein ZO‐1 J Hepatol, 2005;42(5):707–18.
16. Amasheh, S., Meiri, N., Gitter, A.H. et  al. Claudin‐2 expression induces
also binds to connexins, and that occludin colocalizes with
cation‐selective channels in tight junctions of epithelial cells. J Cell Sci,
Cx32 indicate the possibility for either coordinate or reciprocal 2002;115(Pt 24):4969–76.
regulation of macromolecular complexes containing gap and 17. Paul, D.L. Molecular cloning of cDNA for rat liver gap junction protein.
tight junction proteins. Studies of protein–protein interactions J Cell Biol, 1986;103(1):123–34.
and of coordinate and subordinate regulation of gene families 18. Kumar, N.M. and Gilula, N.B. The gap junction communication channel.
Cell, 1996;84(3):381–8.
may elucidate the intricacies of inter‐ and intracellular signal­ 19. Zhang, J.T. and Nicholson, B.J. Sequence and tissue distribution of a second
ing concerning growth control by gap junctions and mainte­ protein of hepatic gap junctions, Cx26, as deduced from its cDNA. J Cell
nance of the “blood–biliary barrier” formed by tight junctions Biol, 1989;109(6 Pt 2):3391–401.
[30, 93]. As with the HCV receptor CD81, the major gap junc­ 20. Dahl, G. and Muller, K.J. Innexin and pannexin channels and their signaling.
tion components (connexins) and the major tight junction com­ FEBS Lett, 2014;588(8):1396–402.
21. Maes, M., Crespo Yanguas, S., Willebrords, J., Cogliati, B., and Vinken M.
ponents (occludin and claudins) are tetraspan proteins with Connexin and pannexin signaling in gastrointestinal and liver disease. Transl
critical intracellular and extracellular domains. The appropria­ Res, 2015;166(4):332–43.
tion of tight junction proteins as portals for virus entry suggests 22. Sosinsky, G.E., Boassa, D., Dermietzel, R. et al. Pannexin channels are not
special properties for these proteins. As tight junctions actively gap junction hemichannels. Channels (Austin), 2011;5(3):193–7.
172 THE LIVER: REFERENCES

23. Willebrords, J., Maes, M., Crespo Yanguas, S., and Vinken M. Inhibitors of 49. Seseke, F.G., Gardemann, A., and Jungermann K. Signal propagation via gap
connexin and pannexin channels as potential therapeutics. Pharmacol Ther, junctions, a key step in the regulation of liver metabolism by the sympathetic
2017;180:144–60. hepatic nerves. FEBS Lett, 1992;301(3):265–70.
24. Stauffer, K.A. The gap junction proteins beta 1‐connexin (connexin‐32) and 50. Saez, J.C., Connor, J.A., Spray, D.C., and Bennett, M.V. Hepatocyte gap
beta 2‐connexin (connexin‐26) can form heteromeric hemichannels. J Biol junctions are permeable to the second messenger, inositol 1,4,5‐trisphos­
Chem, 1995;270(12):6768–72. phate, and to calcium ions. Proc Natl Acad Sci U S A, 1989;86(8):2708–12.
25. Valiunas, V., Niessen, H., Willecke, K., and Weingart, R. Electrophysiological 51. Nathanson, M.H. and Burgstahler, A.D. Coordination of hormone‐induced
properties of gap junction channels in hepatocytes isolated from connexin32‐ calcium signals in isolated rat hepatocyte couplets: demonstration with con­
deficient and wild‐type mice. Pflugers Arch, 1999;437(6):846–56. focal microscopy. Mol Biol Cell, 1992;3(1):113–21.
26. Locke, D., Liu, J., and Harris, A.L. Lipid rafts prepared by different methods 52. Clapham, D.E. Calcium signaling. Cell, 2007;131(6):1047–58.
contain different connexin channels, but gap junctions are not lipid rafts. 53. Bartlett, P.J., Gaspers, L.D., Pierobon, N., and Thomas, A.P. Calcium‐
Biochemistry, 2005;44(39):13027–42. dependent regulation of glucose homeostasis in the liver. Cell Calcium,
27. Duffy, H.S., Delmar, M., and Spray, D.C. Formation of the gap junction 2014;55(6):306–16.
nexus: binding partners for connexins. J Physiol Paris, 2002;96(3–4): 54. Guerra, M.T. and Nathanson, M.H. Calcium signaling and secretion in chol­
243–9. angiocytes. Pancreatology, 2015;15(4 Suppl):S44–8.
28. Duffy, H.S., Fort, A.G., and Spray, D.C. Cardiac connexins: genes to nexus. 55. Scemes, E. and Spray, D.C. Extracellular K+ and astrocyte signaling via con­
Adv Cardiol, 2006;42:1–17. nexin and pannexin channels. Neurochem Res, 2012;37(11):2310–16.
29. Duffy, H.S., Iacobas, I., Hotchkiss, K. et al. The gap junction protein con­ 56. Schlosser, S.F., Burgstahler, A.D., and Nathanson, M.H. Isolated rat hepato­
nexin32 interacts with the Src homology 3/hook domain of discs large cytes can signal to other hepatocytes and bile duct cells by release of nucleo­
homolog 1. J Biol Chem, 2007;282(13):9789–96. tides. Proc Natl Acad Sci U S A, 1996;93(18):9948–53.
30. Kojima, T., Yamamoto, T., Murata, M. et al. Regulation of the blood‐biliary 57. Aasen, T., Mesnil, M., Naus, C.C., Lampe, P.D., and Laird, D.W. Gap junc­
barrier: interaction between gap and tight junctions in hepatocytes. Med tions and cancer: communicating for 50 years. Nat Rev Cancer,
Electron Microsc, 2003;36(3):157–64. 2016;16(12):775–88.
31. Fowler, S.L., Akins, M., Zhou, H., Figeys, D., and Bennett, S.A. The liver 58. Yamasaki, H., Krutovskikh, V., Mesnil, M. et al. Role of connexin (gap junc­
connexin32 interactome is a novel plasma membrane‐mitochondrial tion) genes in cell growth control and carcinogenesis. C R Acad Sci III,
­signaling nexus. J Proteome Res, 2013;12(6):2597–610. 1999;322(2–3):151–9.
32. Graf, J. and Boyer, J.L. The use of isolated rat hepatocyte couplets in 59. Saez, J.C., Gregory, W.A., Watanabe, T. et al. cAMP delays disappearance of
­hepatobiliary physiology. J Hepatol, 1990;10(3):387–94. gap junctions between pairs of rat hepatocytes in primary culture. Am J
33. Gunzel, D., Zakrzewski, S.S., Schmid, T. et al. From TER to trans‐ and para­ Physiol, 1989;257(1 Pt 1):C1–11.
cellular resistance: lessons from impedance spectroscopy. Ann N Y Acad Sci, 60. Spray, D.C., Rozental, R., and Srinivas, M. Prospects for rational develop­
2012;1257:142–51. ment of pharmacological gap junction channel blockers. Curr Drug Targets,
34. Suzuki, H., Nishizawa, T., Tani, K. et al. Crystal structure of a claudin pro­ 2002;3(6):455–64.
vides insight into the architecture of tight junctions. Science, 2014;344(6181): 61. Saez, J.C., Nairn, A.C., Czernik, A.J. et al. Phosphorylation of connexin 32,
304–7. a hepatocyte gap‐junction protein, by cAMP‐dependent protein kinase, pro­
35. Matsumoto, K., Imasato, M., Yamazaki, Y. et  al. Claudin 2 deficiency tein kinase C and Ca2+/calmodulin‐dependent protein kinase II. Eur J
reduces bile flow and increases susceptibility to cholesterol gallstone disease Biochem, 1990;192(2):263–73.
in mice. Gastroenterology, 2014;147(5):1134–45 e10. 62. Koval, M., Isakson, B.E., and Gourdie, R.G. Connexins, pannexins and
36. Buckley, A. and Turner, J.R. Cell biology of tight junction barrier regulation innexins: protein cousins with overlapping functions. FEBS Lett,
and mucosal disease. Cold Spring Harb Perspect Biol, 2018;10(1). 2014;588(8):1185.
37. Fasano, A., Fiorentini, C., Donelli, G. et  al. Zonula occludens toxin 63. Epifantseva, I. and Shaw, R.M. Intracellular trafficking pathways of Cx43
­modulates tight junctions through protein kinase C‐dependent actin reor­ gap junction channels. Biochim Biophys Acta, 2018;1860(1):40–7.
ganization, in vitro. J Clin Invest, 1995;96(2):710–20. 64. Fort, A.G., Murray, J.W., Dandachi, N. et al. In vitro motility of liver con­
38. Tsukamoto, T. and Nigam, S.K. Tight junction proteins form large com­ nexin vesicles along microtubules utilizes kinesin motors. J Biol Chem,
plexes and associate with the cytoskeleton in an ATP depletion model for 2011;286(26):22875–85.
reversible junction assembly. J Biol Chem, 1997;272(26):16133–9. 65. Francis, R., Xu, X., Park, H. et al. Connexin43 modulates cell polarity and
39. Turner, J.R., Angle, J.M., Black, E.D. et  al. PKC‐dependent regulation of directional cell migration by regulating microtubule dynamics. PLoS One,
transepithelial resistance: roles of MLC and MLC kinase. Am J Physiol, 2011;6(10):e26379.
1999;277(3 Pt 1):C554–62. 66. Polusani, S.R., Kalmykov, E.A., Chandrasekhar, A., Zucker, S.N., and
40. Pradhan‐Sundd, T., Vats, R., Russell, J.M. et al. Dysregulated bile transport­ Nicholson, B.J. Cell coupling mediated by connexin 26 selectively con­
ers and impaired tight junctions during chronic liver injury in mice. tributes to reduced adhesivity and increased migration. J Cell Sci,
Gastroenterology, 2018;155(4):1218–32.e24. 2016;129(23):4399–410.
41. Gissen, P. and Arias, I.M. Structural and functional hepatocyte polarity and 67. Matsuuchi, L. and Naus, C.C. Gap junction proteins on the move: connexins,
liver disease. J Hepatol, 2015;63(4):1023–37. the cytoskeleton and migration. Biochim Biophys Acta, 2013;1828(1):
42. Mailly, L., Xiao, F., Lupberger, J. et al. Clearance of persistent hepatitis C 94–108.
virus infection in humanized mice using a claudin‐1‐targeting monoclonal 68. Falk, M.M., Bell, C.L., Kells Andrews, R.M., and Murray, S.A. Molecular
antibody. Nat Biotechnol, 2015;33(5):549–54. mechanisms regulating formation, trafficking and processing of annular gap
43. Crespo Yanguas, S., Willebrords, J., Maes, M. et al. Connexins and pannex­ junctions. BMC Cell Biol, 2016;17(Suppl 1):22.
ins in liver damage. EXCLI J, 2016;15:177–86. 69. Gaietta, G., Deerinck, T.J., Adams, S.R. et al. Multicolor and electron micro­
44. Gumucio, J.J. and Miller, D.L. Functional implications of liver cell heteroge­ scopic imaging of connexin trafficking. Science, 2002;296(5567):503–7.
neity. Gastroenterology, 1981;80(2):393–403. 70. Bejarano, E., Yuste, A., Patel, B. et al. Connexins modulate autophagosome
45. Jungermann, K. and Katz N. Functional specialization of different hepato­ biogenesis. Nat Cell Biol, 2014;16(5):401–14.
cyte populations. Physiol Rev, 1989;69(3):708–64. 71. Laird, D.W. Syndromic and non‐syndromic disease‐linked Cx43 mutations.
46. Berthoud, V.M., Iwanij, V., Garcia, A.M., and Saez, J.C. Connexins and FEBS Lett, 2014;588(8):1339–48.
glucagon receptors during development of rat hepatic acinus. Am J Physiol, 72. Scherer, S.S. and Kleopa, K.A. X‐linked Charcot‐Marie‐Tooth disease.
1992;263(5 Pt 1):G650–8. J Peripher Nerv Syst, 2012;17(Suppl 3):9–13.
47. Kinugasa, A. and Thurman, R.G. Differential effect of glucagon on gluco­ 73. Nelles, E., Butzler, C., Jung, D. et al. Defective propagation of signals gener­
neogenesis in periportal and pericentral regions of the liver lobule. Biochem ated by sympathetic nerve stimulation in the liver of connexin32‐deficient
J, 1986;236(2):425–30. mice. Proc Natl Acad Sci U S A, 1996;93(18):9565–70.
48. Lee, S.M. and Clemens, M.G. Subacinar distribution of hepatocyte mem­ 74. Stumpel, F., Ott, T., Willecke, K., and Jungermann K. Connexin 32 gap junc­
brane potential response to stimulation of gluconeogenesis. Am J Physiol., tions enhance stimulation of glucose output by glucagon and noradrenaline
1992;263(3 Pt 1):G319–26. in mouse liver. Hepatology, 1998;28(6):1616–20.
14:  Gap and Tight Junctions in Liver: Structure, Function, and Pathology 173

75. Niessen, H. and Willecke K. Strongly decreased gap junctional permeability 83. Spray, D.C., Hanstein, R., Lopez‐Quintero, S.V. et  al. Gap junctions and
to inositol 1,4, 5‐trisphosphate in connexin32 deficient hepatocytes. FEBS bystander effects: Good Samaritans and executioners. Wiley Interdiscip Rev
Lett, 2000;466(1):112–14. Membr Transp Signal, 2013;2(1):1–15.
76. Schwarz, M., Wanke, I., Wulbrand, U., Moennikes, O., and Buchmann A. 84. Zibara, K., Awada, Z., Dib, L. et al. Anti‐angiogenesis therapy and gap junc­
Role of connexin32 and beta‐catenin in tumor promotion in mouse liver. tion inhibition reduce MDA‐MB‐231 breast cancer cell invasion and metas­
Toxicol Pathol, 2003;31(1):99–102. tasis in vitro and in vivo. Sci Rep, 2015;5:12598.
77. King, T.J. and Lampe, P.D. Altered tumor biology and tumorigenesis in 85. Crespo Yanguas, S., da Silva, T.C., Pereira, I.V.A. et al. TAT‐Gap19 and car­
irradiated and chemical carcinogen‐treated single and combined con­ benoxolone alleviate liver fibrosis in mice. Int J Mol Sci, 2018;19(3).
nexin32/p27Kip1‐deficient mice. Cell Commun Adhes, 2005;12(5–6): 86. Iorga, A., Dara, L., and Kaplowitz, N. Drug‐induced liver injury: cascade of
293–305. events leading to cell death, apoptosis or necrosis. Int J Mol Sci, 2017;18(5).
78. Kojima, T., Fort, A., Tao, M., Yamamoto, M., and Spray, D.C. Gap junction 87. Asamoto, M., Hokaiwado, N., Murasaki, T., and Shirai, T. Connexin 32
expression and cell proliferation in differentiating cultures of Cx43 KO dominant‐negative mutant transgenic rats are resistant to hepatic damage by
mouse hepatocytes. Am J Physiol Gastrointest Liver Physiol, 2001;281(4): chemicals. Hepatology, 2004;40(1):205–10.
G1004–13. 88. Naiki‐Ito, A., Asamoto, M., Naiki, T. et al. Gap junction dysfunction reduces
79. Temme, A., Ott, T., Dombrowski, F., and Willecke, K. The extent of synchro­ acetaminophen hepatotoxicity with impact on apoptotic signaling and con­
nous initiation and termination of DNA synthesis in regenerating mouse liver nexin 43 protein induction in rat. Toxicol Pathol, 2010;38(2):280–6.
is dependent on connexin32 expressing gap junctions. J Hepatol, 2000;32(4): 89. Patel, S.J., Milwid, J.M., King, K.R. et al. Gap junction inhibition prevents
627–35. drug‐induced liver toxicity and fulminant hepatic failure. Nat Biotechnol,
80. Dagli, M.L., Yamasaki, H., Krutovskikh, V., and Omori, Y. Delayed liver 2012;30(2):179–83.
regeneration and increased susceptibility to chemical hepatocarcinogenesis 90. Maes, M., McGill, M.R., da Silva, T.C. et al. Connexin32: a mediator of aceta­
in transgenic mice expressing a dominant‐negative mutant of connexin32 minophen‐induced liver injury? Toxicol Mech Methods, 2016;26(2):88–96.
only in the liver. Carcinogenesis, 2004;25(4):483–92. 91. Maes, M., Crespo Yanguas, S., Willebrords, J. et al. Connexin hemichannel
81. Igarashi, I., Makino, T., Suzuki, Y. et  al. Background lesions during a inhibition reduces acetaminophen‐induced liver injury in mice. Toxicol Lett,
­24‐month observation period in connexin 32‐deficient mice. J Vet Med 2017;278:30–7.
Sci, 2013;75(2):207–10. 92. Maes, M. and Vinken M. Connexin‐based signaling and drug‐induced hepa­
82. Li, Q., Omori, Y., Nishikawa, Y. et  al. Cytoplasmic accumulation of con­ totoxicity. J Clin Transl Res, 2017;3(Suppl 1):189–98.
nexin32 protein enhances motility and metastatic ability of human hepatoma 93. Lee, N.P. The blood‐biliary barrier, tight junctions and human liver diseases.
cells in vitro and in vivo. Int J Cancer, 2007;121(3):536–46. Adv Exp Med Biol, 2012;763:171–85.
Ribosome Biogenesis and its
15 Role in Cell Growth and
Proliferation in the Liver
Katherine I. Farley‐Barnes1 and Susan J. Baserga1,2,3
1
Department of Molecular Biophysics & Biochemistry, Yale University School of Medicine, New Haven, CT, USA
2
Department of Genetics, Yale University School of Medicine, New Haven, CT, USA
3
Department of Therapeutic Radiology, Yale University School of Medicine, New Haven, CT, USA

INTRODUCTION Saccharomyces cerevisiae. Only recently have scientists made


advances into understanding how this process works in humans.
Ribosomes, the cellular machinery responsible for all protein Additionally, how ribosome biogenesis changes among differ-
synthesis, are essential for cells to grow and to divide. The bio- ent tissues is largely unknown. One of the best models for
genesis of ribosomes is thus a highly regulated process that understanding ribosome biogenesis at the tissue level, however,
coordinates a number of cellular cues and depends especially has been the liver. Foundational studies on the liver have pro-
upon nutrient availability. Nowhere is the relationship between vided insights into how ribosomes are made and how the cell
nutrient availability, cellular signaling, and ribosome biogenesis responds to various stimuli to regulate the production of ribo-
more critical than in the liver. Liver size decreases dramatically somes. Indeed, nucleoli, the non‐membrane‐bound organelles
after fasting in rats and mice [1, 2]. When refed, however, the responsible for ribosome production, were first biochemically
liver size recovers within 24 hours [1, 2]. Total liver proteins and isolated from liver cells in 1956 [8]. It is likely that in the future,
RNA follow the same fluctuation, indicating that the liver the liver will continue to play a large role in our understanding
responds to limited nutrients by downregulating ribosome pro- of ribosome biogenesis and its relation to human disease.
duction and upregulating it again once the organism is presented
with an adequate diet [1–3].
Understanding ribosome biogenesis is not only essential to
understanding basic cellular biology, but also to probing the OVERVIEW OF RIBOSOME BIOGENESIS
pathogenesis of a number of human diseases. Such diseases
include cancer, where an increase in cell growth and prolifera- Ribosome biogenesis begins in a non‐membrane‐bound orga-
tion goes hand‐in‐hand with an increase in ribosome biogenesis nelle inside the cell nucleus called the nucleolus. It starts with
(reviewed in [4]). Ribosome dysfunction has also been transcription of the tandemly repeated ribosomal DNA (rDNA)
implicated in neurodegenerative diseases (reviewed in [5]).
­ by RNA polymerase I (RNAPI). As the rDNA is being tran-
Additionally, aberrant ribosome biogenesis causes a number of scribed, the nucleolus forms around it. Thus, the repeats of
tissue‐specific disorders, termed ribosomopathies (reviewed in rDNA on the chromosomes are called nucleolar organizing
[6, 7]). It is intriguing that dysregulation of such an essential regions, or NORs. The NORs are located on five different chro-
process like ribosome biogenesis would result in a viable organ- mosomes in humans (13, 14, 15, 21, and 22). Therefore, human
ism. Indeed, how changes in ribosome biogenesis, a process that cells have the potential to form 10 nucleoli per cell, although far
is required for every cell type, would create tissue‐specific fewer nucleoli are often observed [9]. The number of nucleoli in
pathologies is an outstanding question in the field. a diploid mouse hepatocyte ranges from 3 to 6 (mice have the
Much of the current knowledge of the steps in ribo- potential to form up to 12 nucleoli per cell), with 3 nucleoli per
some ­biogenesis comes from studies in the budding yeast, nucleus being seen most often [10].

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
15:  Ribosome Biogenesis and its Role in Cell Growth and Proliferation in the Liver 175

The human nucleolus consists of three different compart- In addition to the modifications made to the pre‐rRNA, the pre‐
ments: the fibrillar center (FC), the dense fibrillar component rRNA is also extensively cleaved and processed. In humans, pre‐
(DFC), and the granular component (GC) (Figure 15.1). A distinct rRNA processing occurs through multiple pathways (Figure 15.2).
function in making ribosomes can be ascribed to each of these The difference between the two major processing pathways
compartments. The nucleolus forms around the FC, and at the begins with the choice of cleavage at either site 1 or site 2. If site
interface of the FC and the DFC, where rDNA transcription 1 is cleaved first, the 41S pre‐rRNA is made first until a later
takes place. After the transcription of the rDNA, modification of cleavage at site 2 separates the 21S and 32S pre‐rRNAs. If site 2
the pre‐ribosomal RNA (pre‐rRNA), processing of the pre‐ is cleaved first, the small and large subunit pre‐rRNAs are sepa-
rRNA, and pre‐ribosome assembly proceed outwards from the rated into the 30S and 32S pre‐rRNAs, respectively (Figure 15.2,
DFC into the third nucleolar compartment, the GC. In the GC, and reviewed in [14]). Regardless of the pathway used, the end
ribosomal proteins (r‐proteins) assemble with the pre‐rRNA to result is the mature 18S, 5.8S, and 28S rRNAs.
form the pre‐small subunit (pre‐SSU) and pre‐large subunit In addition to the snoRNAs and r‐proteins, a number of trans‐
(pre‐LSU) of the ribosome [11]. acting factors are also important for optimal pre‐rRNA process-
One explanation for the tripartite organization of the nucleo- ing. Best characterized in yeast, these factors include endo‐ and
lus uses the physics of liquid–liquid phase separations. Just like exonucleases, helicases, AAA‐ATPases, GTPases, and more [15].
an oil droplet in water, the proteins of the nucleolus separate Many of the functions of these enzymes are still being defined. In
into multiple compartments due to differences in surface tension yeast, over 200 proteins are involved in ribosome assembly [15],
and hydrophobicity. For example, a structure similar to the a number that is likely to be greatly increased in humans. Indeed,
DFC/GC compartments can be replicated in vitro by mixing a series of studies are currently striving to define all of the factors
fibrillarin (FBL), a box C/D small nucleolar ribonucleoprotein necessary for this process in humans [16–19].
(snoRNP) methyltransferase enzyme that is enriched in the Ribosome assembly concludes in the cytoplasm, after the
DFC, and nucleophosmin (NPM1), a protein with a number of addition of the 5S rRNA which is transcribed from chromosome
functions that is enriched in the GC [12]. 1 by RNA polymerase III (RNAPIII) in humans. The pre‐SSU
At the FC/DFC interface, the rDNA is transcribed as a 47S and pre‐LSU are both exported through the nuclear pore com-
polycistronic precursor that contains 3 of the 4 the ribosomal plex, and export of each subunit requires common factors as
RNAs (rRNAs) which eventually are incorporated into the small well as subunit‐specific factors. Exportin 1 (XPO1, or Crm1 in
(18S) and large (5.8S and 28S) subunits of the ribosome. To yeast) is one of the best characterized proteins involved in both
form the mature rRNAs, the pre‐rRNA must be modified, large and small subunit export [20]. Additionally, the final steps
cleaved, and processed. Two types of RNA modifications are of pre‐rRNA processing occur in the cytoplasm, including the
primarily used: 2′‐O‐methylation and pseudouridylation processing of the 18SE to the mature 18S rRNA at site 3.
(reviewed in [13]). Generally, the modifications are placed at Overall, human ribosome biogenesis begins in the nucleolus but
functionally important regions of the ribosome, such as the encompasses the nucleus and cytoplasm as well, leaving many
tRNA‐binding sites or at the interface of the small and large steps in the process open to regulation.
subunits (reviewed in [13]). These modifications aid in folding
the pre‐rRNA into the correct secondary and tertiary structures.
Most of these modifications are made by snoRNPs which direct
their guide RNA to base pair with the corresponding site to be REGULATION OF RIBOSOME BIOGENESIS
modified and position the enzyme to perform the modification.
The majority of pre‐rRNA modifications are carried out by two Our understanding of the intricacies of the regulation of human
types of snoRNPs: box C/D, which perform the 2′‐O‐methylations, ribosome biogenesis is ever increasing. Regulation begins at the
and box H/ACA, which perform the pseudouridylations, both level of the rDNA with the organization and silencing of select
named for conserved snoRNA sequences. rDNA repeats. RNAPI transcription of the rDNA is also
­influenced by many cellular signals. Correct processing of the
pre‐rRNA is controlled by a number of molecules that signal to
FC modulate the transcription of r‐proteins, snoRNPs, and other
assembly and processing factors. Finally, nuclear export of the
DFC
pre‐ribosomal subunits can be monitored and regulated by
the cell. In all, ribosome biogenesis requires all three RNA poly-
merases and hundreds of other factors, so there are many ways
for the cell to control steps in this process. Just a few of these
mechanisms are described in detail below.
GC
mTOR and the regulation of ribosomal
Figure 15.1  The morphology of the nucleolus in mammalian cells. protein translation
On the left is the nucleolus of a cell actively engaged in ribosome bio-
One key regulator of ribosome biogenesis is the mechanistic tar-
genesis. On the right is the nucleolus of a cell undergoing inhibition of
rRNA transcription. Indicated are the three nucleolar compartments: get of rapamycin (mTOR) pathway. Essential for cell growth,
the fibrillar center (FC), the dense fibrillar component (DFC), and the the mTOR pathway responds to various cellular signals to
granular component (GC). See text for details. upregulate the transcription of r‐proteins and to ultimately
176 THE LIVER:  REGULATION OF RIBOSOME BIOGENESIS

A′ A0 1 3 2a 2 4

47S pre-rRNA 5′ 18S 5.8S 28S 3′


5′ETS ITS1 ITS2 3′ETS

A0 1 3 2a 2 4

45S 18S 5.8S 28S

site 2 cleavage site 1 cleavage

A0 1 3 2a 4 3 2a 2 4

30S 18S 32S 5.8S 28S 18S 5.8S 28S 41S

3 2a 3 2a 4

21S 18S 21S 18S 5.8S 28S 32S

18SE 18S 12S 5.8S 18SE 18S 12S 5.8S

18S 5.8S 28S 18S 5.8S 28S

Figure 15.2  Schematic depicting the two major pre‐rRNA processing pathways in human cells. The 47S pre‐rRNA is processed and cleaved
through multiple pathways to form the mature 18S, 5.8S, and 28S rRNAs that are incorporated into the synthesized ribosome. ETS and ITS
­represent external transcribed spacers and internal transcribed spacers, respectively.

increase global protein synthesis. There are two different mTOR plays an important role in ribosome biogenesis
mTOR‐containing cytoplasmic complexes, termed mTORC1 through control of the translation of mRNAs encoding
and mTORC2. The mTORC1 complex contains mTOR, RPTOR r‐proteins. In higher eukaryotes, mRNAs that code for r‐proteins
(regulatory associated protein of MTOR complex 1), AKT1S1 are denoted by a 5′ terminal oligopyrimidine (5′TOP) motif
(AKT substrate 1), DEPTOR (DEP domain containing mTOR‐ [24]. This motif begins with a C and contains, on average, 12.2
interacting protein), and MLST8 (mTOR‐associated protein, nucleotides in humans [25]. When cells are stimulated with
LST8 homolog). mTORC2 contains mTOR, RICTOR (RPTOR‐ nutrients, 5′TOP mRNAs are recruited to ribosomes for
independent companion of MTOR complex 2), PRR5 (proline‐ increased translation [26]. Rapamycin treatment prevents the
rich 5), MAPKAP1 (mitogen‐activated protein kinase‐associated translation of 5′TOP mRNAs, and mutation of the 5′TOP
protein 1), DEPTOR, and MLST8. Although mTORC1 and ­renders the mRNA’s translation insensitive to rapamycin treat-
mTORC2 have different complex members and substrates, ment [27–29]. More recent studies have shown that mTORC1
mTOR functions as a serine/threonine kinase in both complexes. affects 5′TOP translation through its phosphorylation of
Importantly for ribosome biogenesis, mTORC1 signals to RPS6 EIF4E‐binding proteins (4E‐BPs) [30]. Phosphorylation by
kinases 1 and 2 (S6K1 and S6K2). S6K1 phosphorylates several active mTORC1 allows EIF4E to bind EIF4G1 [30]. 5′TOP
substrates, ultimately promoting increased translation initiation mRNAs require this EIF4E–EIF4G1 interaction for their cap
and elongation, as well as increased ribosome biogenesis binding and translation, while other mRNAs do not, providing
through the upregulation of rDNA transcription (reviewed in an explanation for mTOR’s specific role in 5′TOP translation
[21]). mTORC1 also phosphorylates eukaryotic translation ini- after nutrient stimulation [30].
tiation factor 4B pseudogene 1 (4EBP1). 4EBP1 affects cap‐ Regarding feeding and nutrient stimulation, the liver relies
dependent translation, and its phosphorylation prevents its heavily on mTORC1 signaling. Nutrient uptake, as well as
binding to eukaryotic translation initiation factor 4E (EIF4E). uptake of amino acids such as leucine, stimulates ribosome
This leaves EIF4E free to bind eukaryotic initiation factor 4 ­biogenesis in the liver [31]. In the livers of rats administered
gamma proteins (EIF4Gs), which then bind 5′ caps of mRNAs leucine, mTOR signaling is upregulated and translation of
and recruit other factors required for translation (reviewed in 5′TOP‐containing mRNAs is increased [32]. This response can
[22]). Of note, some of mTORC1’s function is inhibited by the be inhibited using the drug rapamycin [32]. Translation of
immunosuppressant rapamycin [23]. ribosomal protein mRNAs is also upregulated during liver
­
15:  Ribosome Biogenesis and its Role in Cell Growth and Proliferation in the Liver 177

regeneration in rats, although the precise mechanism of action residue S44 and decreasing the amount of S199 phosphorylation,
is unknown [33, 34]. ultimately increasing TIF‐1A’s interaction with SL1 [44].
Additionally, in the liver, mTOR signaling acts as an impor- Additionally, mTOR may bind directly to the RNAPI
tant regulatory mechanism in the body’s response to ethanol. promoter [45].
Chen et  al. found decreased levels of the protein DEPTOR, The MAPK signaling pathway also has been shown to play a
a  known inhibitor of mTORC1 phosphorylation, in both a role in rDNA transcriptional regulation through multiple mech-
chronic‐plus‐binge ethanol‐fed mouse model of alcoholic stea- anisms. In the MAPK pathways, extracellular stimulation by
tosis and in patients with alcoholic liver disease [35]. Decreased factors such as epidermal growth factor (EGF) triggers a cas-
DEPTOR leads to increased levels of 4EBP1 and S6K phospho- cade of protein kinases which in turn promote factors necessary
rylation [35]. Phosphorylated S6K in turn phosphorylates sterol for growth and development. Like mTOR signaling, MAPK
regulator element‐binding transcription factor 1 (SREBF1), phosphorylation of two residues on TIF‐1A increases RNAPI
which translocates to the nucleus and upregulates the expres- initiation [46]. Additionally, UBF is phosphorylated via the
sion of proteins involved in fatty acid biosynthesis [36, 37]. MAPK signaling cascade. UBF phosphorylation may then
Phosphorylated mTORC1 can also phosphorylate lipin 1, pro- enhance UBF’s interaction with SL1 and/or disrupt formation of
moting the transcription of SREBF1 which contributes to the the enhanceosomes that are promoted by UBF and slow down
build up of lipids in alcoholic steatosis [38]. Overall, mTOR RNAPI [43, 47]. Both options increase RNAPI activity.
signaling highlights the importance of regulating ribosome Signaling through the retinoblastoma protein (pRB), a pocket
­biogenesis for proper liver health. family protein and tumor suppressor, can also affect ribosome
biogenesis via modulation of RNAPI transcription. This occurs
both directly through the RNAPI transcription machinery and
Regulation of RNAPI transcription indirectly through the interaction of pRB with other factors.
RNAPI transcriptional regulation is critical in the liver, since the pRB, as well as the pocket family protein p130, directly affects
liver responds so dramatically to nutrient availability. Upon rDNA transcription by disrupting the interaction between SL1
feeding, it has been shown that RNAPI activity quickly increases and UBF [48, 49]. Indirectly, pRB modulates rDNA transcrip-
[2]. Similarly, RNAPI activity increases during liver regenera- tion through its interactions with E2F1. When pRB is hyper-
tion, in addition to the increased r‐protein translation previously phosphorylated, it cannot bind to E2F transcription factors.
mentioned (cited many times, including in [39, 40]). These E2Fs are then free to activate transcription of several
The tandemly repeated rDNA is transcribed by RNAPI as a ­target genes that move the cell cycle successfully from G1 to S
large (47S) polycistronic precursor that contains the 18S, 5.8S, phase. Conversely, when the cell is stressed, pRB becomes
and 28S mature rRNAs. Only about half of the rDNA repeats hypophosphorylated, is able to bind and inhibit E2Fs, and the
are active at any given time, and RNAPI transcription is tightly cell cycle is stopped, ultimately feeding back on ribosome bio-
regulated to ensure generation of an adequate number of ribo- genesis (reviewed in [50]). Additionally, at low expression of
somes for cell growth or maintenance. RNAPI transcription ini- E2F1, E2F1 is able to bind directly to the rDNA promoter, acti-
tiation begins with the binding of upstream binding transcription vating RNAPI transcription [51]. At high E2F1 levels, p14ARF is
factor (UBTF, commonly referred to as UBF) to the upstream expressed, and p14ARF binds to E2F1 and inhibits its transcrip-
control element (UCE) (Figure 15.3). UBF binding then helps to tional activity so that rDNA transcription is decreased [52].
recruit selectivity factor 1 (SL1), a species‐specific complex In  turn, decreased rDNA transcription feeds back on E2F1
containing the TATA‐binding protein (TBP) and three TBP‐ expression to reduce it [53].
associated ­factors (TAFs) in humans [41]. The pre‐initiation
complex is formed when SL1 binds to the core promoter (CP) of
rRNA genes along with RNAPI. The interaction between SL1
Regulation of ribosome biogenesis by MYC
and RNAPI is mediated by TIF‐1A (Rrn3 in S. cerevisiae) [42]. MYC acts as a transcription factor to regulate expression of
Multiple signaling pathways regulate rDNA transcription. numerous proteins required for ribosome biogenesis. It has been
Among these is again mTOR signaling, which alters RNAPI studied regarding hepatocyte proliferation and growth, but
transcription through multiple mechanisms. These mechanisms MYC also plays a role in many other tissues and is especially
include S6K phosphorylation which in turn phosphorylates the significant in cancer [54]. While the MYC transcription factor
C‐terminal tail of UBF, aiding its recruitment of SL1 and the family contains multiple MYC proteins (including c‐MYC,
rest of the RNAPI transcription machinery [43]. mTOR also n‐MYC, and l‐MYC), c‐MYC is the best studied with regards to
works to activate TIF‐1A by promoting phosphorylation of ribosome biogenesis. c‐MYC modulates the transcription medi-
ated by all three RNA polymerases. To enhance RNAPI activity,
c‐MYC can regulate UBF and TIF‐1A expression [55, 56].
c‐MYC is also located in the nucleolus and can directly a­ ssociate
with SL1 to stimulate RNAPI transcription [57, 58]. Additionally,
SL1/TIF-IB PoI I
UBF TIF-IA A43 c‐MYC and n‐MYC regulate RNAPII transcription of r‐proteins
[59–61]. c‐MYC regulates transcription of the 5S rRNA as well
UCE CP by interacting with the RNAPIII transcription machinery [62].
MYC’s role in ribosome biogenesis is critical for proper liver
Figure 15.3  Schematic representation of the transcription initiation function. Overexpression of c‐MYC in hepatocytes results in
complex at the promoter of the rRNA gene. See text for details. enlarged nucleoli as well as increased expression of several
178 THE LIVER:  REGULATION OF RIBOSOME BIOGENESIS

proteins required for ribosome biogenesis [63]. Additionally, cells are still able to stabilize p53, leaving open the possibility
c‐MYC plays a role in several liver diseases [54]. For example, of an additional aspect of the nucleolar stress response in the
c‐MYC expression is upregulated in the livers of patients with liver during starvation [72].
alcoholic liver disease (ALD) and after ethanol feeding in a
mouse model that mimics the early stages of ALD [64].
Interestingly, combined c‐MYC overexpression and ethanol Regulation of ribosome biogenesis is
feeding results in decreased p53 levels [64], indicating a possi-
ble mechanism for the liver dysplasia observed in ALD patients.
coordinated with circadian rhythms
In mice, humans, and birds, it has been shown that liver size
fluctuates according to daily rhythms [73–76]. These daily
Nucleolar stress regulated through p53
rhythms, governed by the organism’s internal circadian clock,
Regulation of ribosome biogenesis is also controlled by the control a number of cellular cues that result in liver size changes.
tumor suppressor p53 in a process known as nucleolar stress For example, mice generally eat at night, during their most
(reviewed in [65, 66]). In a normally functioning cell, r‐proteins active phase. Throughout the night, therefore, mouse liver size
are engaged in the process of translation as structural compo- increases and, conversely, liver size decreases during the day
nents of ribosomes. In this scenario, MDM2 ubiquitinates the when mice are not eating (resting phase) [75]. Interestingly,
p53 protein, targeting p53 for degradation (Figure  15.4). findings by Sinturel et al. have shown that when an organism
However, when there are perturbations in ribosome biogenesis, eats is crucial for regulating this liver mass cycle, since mice fed
r‐proteins become disengaged. Two of these r‐proteins, uL18 during the day (not during their normal feeding time at night) do
(RPL5) and uL5 (RPL11), along with the 5S rRNA, form the 5S not undergo these changes [75].
ribonucleoprotein complex (5S RNP). The 5S RNP then At the biochemical level, liver size fluctuations correlate with
binds to MDM2, blocking MDM2’s ability to ubiquitinate p53, the changes in protein translation controlled by the number of
­ultimately resulting in stabilization of p53 levels (Figure 15.4) ribosomes present in the hepatocytes (i.e. more ribosomes leads
[67, 68]. p53 stabilization causes cellular senescence and to larger cell size and thus larger liver mass) [75]. Interestingly,
­apoptosis through regulating the transcription of several known the increased number of ribosomes is not due to the production
proteins [69]. of more ribosomes by increased RNAPI transcription, but rather
p53 also plays a key role in a number of liver diseases to the decreased ability to eliminate excess ribosomes. During
(reviewed in [70]). Notably, p53 is important in liver regenera- the active phase of the mice, more small subunit r‐proteins
tion, where it must first be stabilized after injury to remove the are  translated, likely through increased mTOR signaling as
affected tissue, and then downregulated to allow for prolifera- ­discussed above [75, 77–79]. During the resting phase, these
tion of new tissue [71]. p53 stabilization also occurs in the liv- r‐proteins are not made, resulting in excess pre‐18S rRNA.
ers  of mice during nutrient deprivation [72]. While scientists To reduce the number of ribosomes in the cell, pre‐18S rRNAs
have proposed a 5′ adenosine monophosphate kinase (AMPK)‐ are polyadenylated [80]. In the resting liver cells, this polyade-
dependent mechanism of p53 stabilization, AMPK‐depleted nylation then targets these RNAs for degradation by the

Normal growing cell

MDM2 Ub Cell growth and


proliferation
p53 (steady state
p53 levels)
Functional ribosomes

Nucleolar stress
uL18
uL18 p14ARF MDM2
5S rRNA uL5
5S rRNA p14ARF
uL5 Cell cycle arrest
p53 and apoptosis
(p53 levels
increased)
uL18
MDM2
Free ribosomal proteins 5S rRNA
uL5

Figure 15.4  The nucleolar stress response in human cells. During normal growth conditions, MDM2 ubiquitinates p53, marking it for degradation
and allowing cell growth and proliferation (top). Under conditions of nucleolar stress, r‐proteins are no longer engaged in functional ribosomes. The
5S RNP containing uL18, uL5, and the 5S rRNA binds MDM2, either with or without p14ARF, to prevent MDM2 from ubiquitinating p53. p53 levels
are therefore stabilized, triggering cell cycle arrest and apoptosis (bottom). Reproduced from [65] with permission of Portland Press Limited.
15:  Ribosome Biogenesis and its Role in Cell Growth and Proliferation in the Liver 179

catalytic component of the exosome, the 3′‐to‐5′ exonuclease acute myelogenous leukemia with increasing patient age
EXOSC10, and fewer ribosomes are produced overall [75, 80]. [90, 91, 93]. The median survival of a cohort of SDS patients
This strategy for controlling ribosome biogenesis may be espe- was recently found to be only 41 years [93].
cially useful in understanding the tissue‐specific intricacies of SDS patients often have liver abnormalities such as hepato-
the liver diseases discussed below. megaly, elevated transaminases and mildly elevated bile acid
levels, particularly early in childhood [94]. Most often the
transaminase levels normalize after infancy though the bile acid
levels often remain elevated. In some patients over the age of 30,
ROLE OF RIBOSOME BIOGENESIS liver microcysts, but not fatty liver or cirrhosis, can be visual-
IN LIVER DISEASE ized by magnetic resonance imaging. Taken together, these find-
ings suggest a connection between the liver and SDS. This is a
North American Indian childhood cirrhosis ripe topic for further follow‐up using new liver imaging tech-
niques. In addition, the development of the gut microbiome in
North American Indian childhood cirrhosis (NAIC, OMIM children may also play an important role [95].
604901) is a liver‐specific ribosomopathy that results from the
mutation of the ribosome biogenesis factor UTP4 (previously
Cirhin) [81, 82]. Identified in the Ojibway‐Cree First Nations Hepatitis B/C and hepatocellular carcinoma
children from Quebec, Canada, NAIC is characterized by
­transient neonatal jaundice that progresses to biliary cirrhosis For hundreds of years, pathologists have used the nucleolus to
[81, 83]. Thus far, the only effective treatment for this disease is prognosticate cancer without truly understanding the connec-
liver transplantation [83]. tion between them [96]. What is known is that many diverse
In this autosomal recessive disorder, all patients have a cancer cells have increased numbers and size of nucleoli, and
homozygous missense mutation on chromosome 16 (16q22) in that these findings correlate with a worse cancer prognosis
the UTP4 gene (R565W) [81, 84]. The UTP4 protein is a mem- (reviewed in [97]).
ber of the t‐Utp subcomplex, which is required for pre‐rRNA In hepatocellular carcinoma (HCC), increased nucleolar size
transcription and processing [85, 86]. Specifically, human UTP4 has been shown to predict the development of HCC in patients
is required for small subunit pre‐rRNA processing at sites A′, with chronic liver disease [98–100]. This correlation is espe-
A0, 1, and 2a (Figure 15.2) [87]. It is intriguing, therefore, that cially strong in patients infected with hepatitis B virus (HBV).
a defect in the function of this ubiquitous and essential protein Interestingly, the HBx oncoprotein of HBV has been known to
leads to liver‐specific pathology. directly affect nucleolar function by binding to the protein
Few models have been developed for studying NAIC. In a NPM1, which aids in bringing HBx to the nucleolus [101].
zebrafish model of NAIC, no pre‐rRNA processing defects were Once in the nucleolus, HBx–NPM1 works to increase rDNA
observed in zebrafish depleted of UTP4 using a morpholino tar- transcription, leading to increased cellular proliferation and
geting either the start site or the splice acceptor site between transformation [101].
exons 14 and 15 [88], despite the established role of UTP4 in The function of the hepatitis C virus (HCV) is also linked to
yeast and human pre‐rRNA processing [85–87]. However, mod- ribosome biogenesis. rDNA transcription is upregulated in
est increases in p53 levels were observed in the UTP4 zebrafish HCV‐infected cells [102, 103] via UBF, where UBF is phospho-
model, and the biliary defects seen were abrogated in a p53‐ rylated after HCV activation of cell cycle proteins cyclin D1
mutated background [88]. While NAIC is the only disorder of (CCND1) and cyclin‐dependent kinase 4 (CDK4). Additionally,
ribosome biogenesis, or ribosomopathy, known to affect liver NS5B, the RNA polymerase required for HCV replication, also
function specifically, this mechanism of p53 stabilization also interacts with the nucleolar phosphoprotein nucleolin (NCL)
occurs in other ribosomopathies [65]. It has also been reported [104]. There is thus a strong link between the nucleolus and
that the Utp4 homozygous knockout mouse is embryonic lethal, viruses responsible for chronic liver disease and HCC.
while the heterozygous mouse develops normally [89], although
these results have not been fully described. Further definition of
Fragile X syndrome
the role of UTP4 in liver pathogenesis is therefore needed.
Fragile X syndrome is caused by mutation in the FMR1 gene on
the X chromosome that results in a greater frequency of a CGG
Shwachman–Diamond syndrome repeated sequence, leading to transcriptional silencing of FMR1
Shwachman–Diamond syndrome (SDS, OMIM 260400) is a gene expression [105]. Loss of the FMR1 protein (FMRP)
ribosomopathy that usually presents in early childhood with causes a range of symptoms including intellectual disability and
bone marrow failure (neutropenia) and/or pancreatic insuffi- obesity [106]. The cell responds to FMRP loss by increasing
ciency manifested as failure to thrive [90, 91]. Ninety percent of mTOR and ERK signaling, both of which are crucial to ribo-
patients have mutations in the Shwachman–Bodian–Diamond some biogenesis and protein synthesis [107, 108]. Decreases in
syndrome gene (SBDS; SDO1 in yeast), though recently muta- eIF4E phosphorylation cause a concomitant increase in the
tions have also been found in DNAC21 and EFL1 [90, 92]. All translation of a protein important for neuronal function, matrix
of the proteins encoded by these genes participate in the late metalloproteinase 9 (MMP‐9). Thus, FMRP loss in fragile X
steps in cytoplasmic maturation of the LSU (60S) subunit of the causes an increase in MMP‐9, which in turn results in intellec-
ribosome. SDS can progress to myelodysplastic syndrome and tual disability.
180 THE LIVER:  REFERENCES

Although fragile X syndrome is not a liver‐specific disorder, 5. Parlato, R. and Kreiner, G. Nucleolar activity in neurodegenerative
recent studies have suggested that it may be treated with met- diseases:  a missing piece of the puzzle? J Mol Med (Berl),
­
2013;91(5):541–7.
formin, a drug that affects liver function [109, 110]. Metformin 6. McCann, K.L. and Baserga, S.J. Genetics. Mysterious ribosomopathies.
is prescribed to manage high blood sugar in patients with type 2 Science, 2013;341(6148):849–50.
diabetes [111]. Metformin was tested as a therapeutic agent for 7. Danilova, N. and Gazda, H.T. Ribosomopathies: how a common root can
fragile X syndrome because it inhibits the mTORC1 and MAPK/ cause a tree of pathologies. Dis Models Mech, 2015;8(9):1013–26.
8. Monty, K.J., Litt, M., Kay, E.R., and Dounce, A.L. Isolation and properties
ERK pathways that are activated in fragile X patients and that are
of liver cell nucleoli. J Biophys Biochem Cytol, 1956;2(2):127–45.
important for liver development [112]. Interestingly, in Frm1−/y 9. Farley, K., Surovtseva, Y., Merkel, J., and Baserga, S. Determinants of mam-
mice, metformin treatment did restore many of the cognitive malian nucleolar architecture. Chromosoma, 2015;124(3):323–31.
defects, including increasing social preference and decreasing 10. Shea, J.R. and Leblond, C.P. Number of nucleoli in various cell types of the
repetitive behaviors. Additionally, defects in insulin signaling mouse. J Morphol, 1966;119(4):425–33.
11. Krüger, T., Zentgraf, H., and Scheer, U. Intranucleolar sites of ribosome
and in circadian rhythms were also rescued by metformin treat-
­biogenesis defined by the localization of early binding ribosomal proteins.
ment in a Drosophila model of fragile X syndrome [113]. J Cell Biol, 2007;177(4):573.
Interestingly, metformin treatment did not rescue levels of 12. Feric, M., Vaidya, N., Harmon, T.S. et al. Coexisting liquid phases underlie
ERK phosphorylation in the livers of Frm1−/y mice, despite nucleolar subcompartments. Cell, 2016;165(7):1686–97.
rescuing these levels in the brain [109]. Additionally, it is pos- 13. Sloan, K.E., Warda, A.S., Sharma, S. et al. Tuning the ribosome: The influ-
ence of rRNA modification on eukaryotic ribosome biogenesis and function.
sible that metformin treatment is influencing other pathways RNA Biol, 2017;14(9):1138–52.
that are affected by FMRP loss, such as gluconeogenesis and 14. Henras, A.K., Plisson‐Chastang, C., O’Donohue, M.‐F., Chakraborty, A.,
the gut microbiome, and it is the rescue of these other path- and Gleizes, P.‐E. An overview of pre‐ribosomal RNA processing in eukary-
ways that contributes to metformin’s effect on the signs and otes. Wiley Interdiscip Rev RNA, 2015;6(2):225–42.
symptoms of fragile X syndrome [109]. In liver cancer cells, 15. Woolford Jr., J.L. and Baserga, S.J. Ribosome biogenesis in the yeast
Saccaromyces cerevisiae. Genetics [Internet], 2013;195:1–39.
the drug metformin has recently been suggested to have an 16. Farley‐Barnes, K.I., McCann, K.L., Ogawa, L.M. et al. Diverse regulators of
anticancer effect through the DEPTOR–mTOR pathway [114], human ribosome biogenesis discovered by changes in nucleolar number.
suggesting an unexplored mechanism of metformin action as Cell Rep, 2018;22(7):1923–34.
related to FMRP loss. 17. Badertscher, L., Wild, T., Montellese, C. et al. Genome‐wide RNAi screen-
ing identifies protein modules required for 40S subunit synthesis in human
cells. Cell Rep, 2015;13(12):2879–91.
18. Wild, T., Horvath, P., Wyler, E. et al. A protein inventory of human ribosome
biogenesis reveals an essential function of exportin 5 in 60S subunit export.
CONCLUSION PLoS Biol, 2010;8(10):e1000522.
19. Tafforeau, L., Zorbas, C., Langhendries, J.L. et al. The complexity of human
Several liver‐specific disorders are directly influenced by ribo- ribosome biogenesis revealed by systematic nucleolar screening of pre‐
rRNA processing factors. Mol Cell, 2013;51(4):539–51.
some biogenesis in humans. Understanding the cellular biology
20. Zemp, I. and Kutay, U. Nuclear export and cytoplasmic maturation of
of human ribosome biogenesis therefore has far‐reaching conse- ­ribosomal subunits. FEBS Lett, 2007;581(15):2783–93.
quences for treating liver disease. However, we have only a 21. Gentilella, A., Kozma, S.C., and Thomas, G. A liaison between mTOR
modest understanding of the complex process of ribosome bio- signaling, ribosome biogenesis and cancer. Biochim Biophys Acta,
­
genesis itself in humans, and we are only beginning to under- 2015;1849(7):812–20.
22. Hay, N. and Sonenberg, N. Upstream and downstream of mTOR. Genes Dev,
stand the layers of regulation needed to modulate ribosome 2004;18(16):1926–45.
biogenesis at the tissue level. Indeed, there are many questions 23. Thoreen, C.C., Kang, S.A., Chang, J.W. et  al. An ATP‐competitive mam-
remaining, including how a such a ubiquitous process as ribo- malian target of rapamycin inhibitor reveals rapamycin‐resistant functions of
some biogenesis may cause liver‐specific defects. Regardless, it mTORC1. J Biol Chem, 2009;284(12):8023–32.
is clear that the process of making ribosomes and the regulation 24. Yoshihama, M., Uechi, T., Asakawa, S. et al. The human ribosomal protein
genes: sequencing and comparative analysis of 73 genes. Genome Res,
of this process are important for liver metabolism and disease 2002;12(3):379–90.
pathogenesis. Therefore, targeting the nucleolus for therapeutic 25. Perry, R.P. The architecture of mammalian ribosomal protein promoters.
intervention may be a viable option in the future of treatment of BMC Evolutionary Biol, 2005;5(1):15.
liver‐associated pathologies. 26. Meyuhas, O. and Kahan, T. The race to decipher the top secrets of TOP
mRNAs. Biochim Biophys Acta, 2015;1849(7):801–11.
27. Jefferies, H.B., Reinhard, C., Kozma, S.C., and Thomas, G. Rapamycin
selectively represses translation of the “polypyrimidine tract” mRNA family.
Proc Natl Acad Sci U S A, 1994;91(10):4441–5.
REFERENCES 28. Terada, N., Patel, H.R., Takase, K. et  al. Rapamycin selectively inhibits
translation of mRNAs encoding elongation factors and ribosomal proteins.
1. Anand, P. and Gruppuso, P.A. Rapamycin inhibits liver growth during Proc Natl Acad Sci U S A, 1994;91(24):11477–81.
­refeeding in the rat via control of ribosomal protein translation but not cap‐ 29. Jefferies, H.B., Fumagalli, S., Dennis, P.B., Reinhard, C., Pearson, R.B., and
dependent translation initiation. J Nutr, 2006;136(1):27–33. Thomas, G. Rapamycin suppresses 5’TOP mRNA translation through
2. Conde, R.D. and Franze‐Fernandez, M.T. Increased transcription and ­inhibition of p70s6k. EMBO J, 1997;16(12):3693–704.
decreased degradation control and recovery of liver ribosomes after a period 30. Thoreen, C.C., Chantranupong, L., Keys, H.R. et al. A unifying model
of protein starvation. Biochem J, 1980;192(3):935–40. for mTORC1‐mediated regulation of mRNA translation. Nature,
3. Addis, T., Poo, L.J., and Lew, W. The rate of protein formation in the 2012;485:109.
organs  and tissues of the body: I. after casein refeeding. J Biol Chem, 31. Anthony, T.G., Anthony, J.C., Yoshizawa, F., Kimball, S.R., and Jefferson,
1936;116(1):343–52. L.S. Oral administration of leucine stimulates ribosomal protein mRNA
4. Derenzini, M., Montanaro, L., and Trere, D. Ribosome biogenesis and cancer. translation but not global rates of protein synthesis in the liver of rats. J Nutr,
Acta Histochem, 2017;119(3):190–7. 2001;131(4):1171–6.
15:  Ribosome Biogenesis and its Role in Cell Growth and Proliferation in the Liver 181

32. Reiter, A.K., Anthony, T.G., Anthony, J.C., Jefferson, L.S., and Kimball, S.R. 56. Poortinga, G., Hannan, K.M., Snelling, H. et al. MAD1 and c‐MYC regulate
The mTOR signaling pathway mediates control of ribosomal protein mRNA UBF and rDNA transcription during granulocyte differentiation. EMBO J,
translation in rat liver. Int J Biochem Cell Biol, 2004;36(11):2169–79. 2004;23(16):3325–35.
33. Aloni, R., Peleg, D., and Meyuhas, O. Selective translational control and 57. Arabi, A., Wu, S., Ridderstråle, K. et al. c‐Myc associates with ribosomal
nonspecific posttranscriptional regulation of ribosomal protein gene expres- DNA and activates RNA polymerase I transcription. Nat Cell Biol,
sion during development and regeneration of rat liver. Mol Cell Biol, 2005;7:303.
1992;12(5):2203–12. 58. Grandori, C., Gomez‐Roman, N., Felton‐Edkins, Z.A. et al. c‐Myc binds to
34. Yo‐ichi, N. and Kikuo, O. Stimulation of the synthesis of ribosomal pro- human ribosomal DNA and stimulates transcription of rRNA genes by RNA
teins  in regenerating rat liver with special reference to the increase in the polymerase I. Nat Cell Biol, 2005;7(3):311–18.
amounts of effective mRNAs for ribosomal proteins. Eur J Biochem, 59. Boon, K., Caron, H.N., van Asperen, R. et al. N‐myc enhances the expres-
1980;107(2):323–9. sion of a large set of genes functioning in ribosome biogenesis and protein
35. Chen, H., Shen, F., Sherban, A. et al. DEPTOR suppresses lipogenesis and synthesis. EMBO J, 2001;20(6):1383–93.
ameliorates hepatic steatosis and acute‐on‐chronic liver injury in alcoholic 60. Coller, H.A., Grandori, C., Tamayo, P. et al. Expression analysis with oligo-
liver disease. Hepatology, 2018. nucleotide microarrays reveals that MYC regulates genes involved in growth,
36. Porstmann, T., Santos, C.R., Griffiths, B. et al. SREBP activity is regulated cell cycle, signaling, and adhesion. Proc Natl Acad Sci U S A,
by mTORC1 and contributes to Akt‐dependent cell growth. Cell Metab, 2000;97(7):3260–5.
2008;8(3–3):224–36. 61. Menssen, A. and Hermeking, H. Characterization of the c‐MYC‐regulated
37. Düvel, K., Yecies, J.L., Menon, S. et al. Activation of a metabolic gene regu- transcriptome by SAGE: Identification and analysis of c‐MYC target genes.
latory network downstream of mTOR complex 1. Mol Cell, Proc Natl Acad Sci U S A, 2002;99(9):6274–9.
2010;39(2):171–83. 62. Gomez‐Roman, N., Grandori, C., Eisenman, R.N., and White, R.J. Direct
38. Peterson, T.R., Sengupta, S.S., Harris, T.E. et al. mTOR complex 1 regulates activation of RNA polymerase III transcription by c‐Myc. Nature,
lipin 1 localization to control the SREBP pathway. Cell, 2011;146(3):408–20. 2003;421:290.
39. Tsukada, K. and Lieberman, I. Synthesis of ribonucleic acid by liver nuclear 63. Kim, S., Li, Q., Dang, C.V., and Lee, L.A. Induction of ribosomal genes and
and nucleolar preparations after partial hepatectomy. J Biol Chem, hepatocyte hypertrophy by adenovirus‐mediated expression of c‐Myc in
1964;239(9):2952–6. vivo. Proc Natl Acad Sci U S A, 2000;97(21):11198–202.
40. Dabeva, M.D. and Dudov, K.P. Transcriptional control of ribosome produc- 64. Nevzorova, Y.A., Cubero, F.J., Hu, W. et al. Enhanced expression of c‐myc
tion in regenerating rat liver. Biochem J, 1982;208(1):101–8. in hepatocytes promotes initiation and progression of alcoholic liver disease.
41. Comai, L., Tanese, N., and Tjian, R. The TATA‐binding protein and associ- J Hepatol, 2016;64(3):628–40.
ated factors are integral components of the RNA polymerase I transcription 65. Farley, K.I. and Baserga, S.J. Probing the mechanisms underlying human
factor, SL1. Cell, 1992;68(5):965–76. diseases in making ribosomes. Biochem Soc Trans, 2016;44(4):1035–44.
42. Miller, G., Panov, K.I., Friedrich, J. et al. hRRN3 is essential in the SL1‐ 66. Woods, S.J., Hannan, K.M., Pearson, R.B., and Hannan, R.D. The nucleolus
mediated recruitment of RNA Polymerase I to rRNA gene promoters. EMBO as a fundamental regulator of the p53 response and a new target for cancer
J, 2001;20(6):1373–82. therapy. Biochim Biophys Acta, 2015;1849(7):821–9.
43. Hannan, K.M., Brandenburger, Y., Jenkins, A. et al. mTOR‐dependent regu- 67. Sloan Katherine, E., Bohnsack Markus, T., and Watkins Nicholas, J. The 5S
lation of ribosomal gene transcription requires S6K1 and is mediated by RNP couples p53 homeostasis to ribosome biogenesis and nucleolar stress.
phosphorylation of the carboxy‐terminal activation domain of the nucleolar Cell Rep, 2013;5(1):237–47.
transcription factor UBF. Mol Cell Biol, 2003;23(23):8862–77. 68. Donati, G., Peddigari, S., Mercer, C.A., and Thomas, G. 5S ribosomal RNA
44. Mayer, C., Zhao, J., Yuan, X., and Grummt, I. mTOR‐dependent activation is an essential component of a nascent ribosomal precursor complex that
of the transcription factor TIF‐IA links rRNA synthesis to nutrient availabil- regulates the Hdm2‐p53 checkpoint. Cell Rep, 2013;4(1):87–98.
ity. Genes Dev, 2004;18(4):423–34. 69. Fischer, M. Census and evaluation of p53 target genes. Oncogene,
45. Tsang, C.K., Liu, H., and Zheng, X.F.S. mTOR binds to the promoters of 2017;36:3943.
RNA polymerase I‐ and III‐transcribed genes. Cell Cycle, 2010;9(5):953–7. 70. Krstic, J., Galhuber, M., Schulz, T., Schupp, M., and Prokesch, A. p53 as a
46. Zhao, J., Yuan, X., Frödin, M., and Grummt, I. ERK‐dependent phosphoryla- dichotomous regulator of liver disease: the dose makes the medicine. Int J
tion of the transcription initiation factor TIF‐IA is required for RNA poly- Mol Sci, 2018;19(3):921.
merase I transcription and cell growth. Mol Cell, 2003;11(2):405–13. 71. Fan, X., Chen, P., Tan, H. et  al. Dynamic and coordinated regulation of
47. Stefanovsky, V., Langlois, F., Gagnon‐Kugler, T., Rothblum, L.I., and Moss, keap1‐NRF2‐ARE and p53/p21 signaling pathways is associated with aceta-
T. Growth factor signaling regulates elongation of RNA polymerase I tran- minophen injury responsive liver regeneration. Drug Metab Dispos,
scription in mammals via UBF phosphorylation and r‐chromatin remode- 2014;42(9):1532.
ling. Mol Cell, 2006;21(5):629–39. 72. Prokesch, A., Graef, F.A., Madl, T. et al. Liver p53 is stabilized upon starva-
48. Hannan, K.M., Hannan, R.D., Smith, S.D. et al. Rb and p130 regulate RNA tion and required for amino acid catabolism and gluconeogenesis. FASEB J,
polymerase I transcription: Rb disrupts the interaction between UBF and 2017;31(2):732–42.
SL‐1. Oncogene, 2000;19(43):4988–99. 73. Leung, N.W.Y., Farrant, P., and Peters, T.J. Liver volume measurement by
49. Cavanaugh, A.H., Hempel, W.M., Taylor, L.J. et al. Activity of RNA poly- ultrasonography in normal subjects and alcoholic patients. J Hepatol,
merase I transcription factor UBF blocked by Rb gene product. Nature, 1986;2(2):157–64.
1995;374(6518):177–80. 74. Wilson, W.O. and McFarland, L.Z. Diurnal changes in livers and digestive
50. Donati, G., Montanaro, L., and Derenzini, M. Ribosome biogenesis and con- systems of Coturnix as related to three photoperiodic regimens. Poultry Sci,
trol of cell proliferation: p53 is not alone. Cancer Res, 2012;72(7):1602. 1969;48(2):477–82.
51. Ayrault, O., Andrique, L., and Séité, P. Involvement of the transcriptional 75. Sinturel, F., Gerber, A., Mauvoisin, D. et al. Diurnal oscillations in liver mass
factor E2F1 in the regulation of the rRNA promoter. Exp Cell Res, and cell size accompany ribosome assembly cycles. Cell, 2017;169(4):651–
2006;312(7):1185–93. 63.e14.
52. Eymin, B., Karayan, L., Seite, P. et al. Human ARF binds E2F1 and inhibits 76. Fisher, H.I. and Bartlett, L.M. Diurnal cycles in liver weights in birds. The
its transcriptional activity. Oncogene, 2001;20(9):1033–41. Condor, 1957;59(6):364–72.
53. Donati, G., Brighenti, E., Vici, M. et al. Selective inhibition of rRNA tran- 77. Jouffe, C., Cretenet, G., Symul, L. et  al. The circadian clock coordinates
scription downregulates E2F‐1: a new p53‐independent mechanism linking ribosome biogenesis. PLoS Biol, 2013;11(1):e1001455.
cell growth to cell proliferation. J Cell Sci, 2011;124(17):3017. 78. Janich, P., Arpat, A.B., Castelo‐Szekely, V., Lopes, M., and Gatfield, D.
54. Zheng, K., Cubero, F.J., and Nevzorova, Y.A. c‐MYC – making liver sick: Ribosome profiling reveals the rhythmic liver translatome and circadian
role of c‐MYC in hepatic cell function, homeostasis and disease. Genes, clock regulation by upstream open reading frames. Genome Res,
2017;8(4):123. 2015;25(12):1848–59.
55. Poortinga, G., Wall, M., Sanij, E. et  al. c‐MYC coordinately regulates 79. Atger, F., Gobet, C., Marquis, J. et al. Circadian and feeding rhythms dif-
­ribosomal gene chromatin remodeling and Pol I availability during granulo- ferentially affect rhythmic mRNA transcription and translation in mouse
cyte differentiation. Nucl Acids Res, 2011;39(8):3267–81. liver. Proc Natl Acad Sci U S A, 2015;112(47):E6579–88.
182 THE LIVER:  REFERENCES

80. Shcherbik, N., Wang, M., Lapik, Y.R., Srivastava, L., and Pestov, D.G. 97. Derenzini, M., Montanaro, L., and Trere, D. What the nucleolus says to a
Polyadenylation and degradation of incomplete RNA polymerase I tran- tumour pathologist. Histopathology, 2009;54(6):753–62.
scripts in mammalian cells. EMBO Rep, 2010;11(2):106–11. 98. Trere, D., Borzio, M., Morabito, A. et al. Nucleolar hypertrophy correlates
81. Betard, C., Rasquin‐Weber, A., Brewer, C. et al. Localization of a recessive with hepatocellular carcinoma development in cirrhosis due to HBV infec-
gene for North American Indian childhood cirrhosis to chromosome tion. Hepatology, 2003;37(1):72–8.
region  16q22‐and identification of a shared haplotype. Am J Hum Genet, 99. Derenzini, M., Trerè, D., Oliveri, F. et al. Is high AgNOR quantity in hepat-
2000;67(1):222–8. ocytes associated with increased risk of hepatocellular carcinoma in
82. Weber, A.M., Tuchweber, B., Yousef, I. et al. Severe familial cholestasis in chronic liver disease? J Clin Pathol, 1993;46(8):727–9.
North American Indian children: a clinical model of microfilament dysfunc- 100. Borzio, M., Trere, D., Borzio, F. et  al. Hepatocyte proliferation rate is a
tion? Gastroenterology, 1981;81(4):653–62. powerful parameter for predicting hepatocellular carcinoma development
83. Drouin, E., Russo, P., Tuchweber, B., Mitchell, G., and Rasquin‐Weber, A. in liver cirrhosis. Mol Pathol, 1998;51(2):96–101.
North American Indian cirrhosis in children: a review of 30 cases. J Pediatr 101. Ahuja, R., Kapoor, N.R., and Kumar, V. The HBx oncoprotein of hepatitis
Gastroenterol Nutr, 2000;31(4):395–404. B virus engages nucleophosmin to promote rDNA transcription and cellu-
84. Chagnon, P., Michaud, J., Mitchell, G. et al. A missense mutation (R565W) lar proliferation. Biochim Biophys Acta, 2015;1853(8):1783–95.
in cirhin (FLJ14728) in North American Indian childhood cirrhosis. Am J 102. Raychaudhuri, S., Fontanes, V., Barat, B., and Dasgupta, A. Activation of
Hum Genet, 2002;71(6):1443–9. ribosomal RNA transcription by hepatitis C virus involves upstream
85. Gallagher, J.E., Dunbar, D.A., Granneman, S. et al. RNA polymerase I tran- ­binding factor phosphorylation via induction of cyclin D1. Cancer Res,
scription and pre‐rRNA processing are linked by specific SSU processome 2009;69(5):2057–64.
components. Genes Dev, 2004;18(20):2506–17. 103. Kao, C.F., Chen, S.Y., and Lee, Y.H. Activation of RNA polymerase I
86. Prieto, J.L. and McStay, B. Recruitment of factors linking transcription and transcription by hepatitis C virus core protein. J Biomed Sci,
­
processing of pre‐rRNA to NOR chromatin is UBF‐dependent and occurs 2004;11(1):72–94.
independent of transcription in human cells. Genes Dev, 2007; 104. Hirano, M., Kaneko, S., Yamashita, T. et  al. Direct interaction between
21(16):2041–54. nucleolin and hepatitis C virus NS5B. J Biol Chem, 2003;278(7):5109–15.
87. Freed, E.F., Prieto, J.L., McCann, K.L., McStay, B., and Baserga, S.J. 105. Kremer, E., Pritchard, M., Lynch, M. et al. Mapping of DNA instability at
NOL11, implicated in the pathogenesis of North American Indian childhood the fragile X to a trinucleotide repeat sequence p(CCG)n. Science,
cirrhosis, is required for pre‐rRNA transcription and processing. PLoS 1991;252(5013):1711–14.
Genet, 2012;8(8):e1002892. 106. Rajaratnam, A., Shergill, J., Salcedo‐Arellano, M. et al. Fragile X s­ yndrome
88. Wilkins, B.J., Lorent, K., Matthews, R.P., and Pack, M. p53‐mediated biliary and fragile X‐associated disorders. F1000Res, 2017;6:2112.
defects caused by knockdown of cirh1a, the zebrafish homolog of the gene 107. Gkogkas, C.G., Khoutorsky, A., Cao, R. et al. Pharmacogenetic inhibition
responsible for North American Indian childhood cirrhosis. PLoS One, of eIF4E‐dependent Mmp9 mRNA translation reverses fragile X s­ yndrome‐
2013;8(10):e77670. like phenotypes. Cell Rep. 2014;9(5):1742–55.
89. Yu, B., Mitchell, G.A., and Richter, A. Cirhin up‐regulates a canonical NF‐ 108. Hou, L., Antion, M.D., Hu, D. et al. Dynamic translational and proteasomal
kappaB element through strong interaction with Cirip/HIVEP1. Exp Cell regulation of fragile X mental retardation protein controls mGluR‐depend-
Res, 2009;315(18):3086–98. ent long‐term depression. Neuron, 2006;51(4):441–54.
90. Warren, A.J. Molecular basis of the human ribosomopathy Shwachman‐ 109. Gantois, I., Khoutorsky, A., Popic, J. et  al. Metformin ameliorates core
Diamond syndrome. Adv Biol Regul, 2018;67:109–27. deficits in a mouse model of fragile X syndrome. Nature Med,
91. Myers, K.C., Bolyard, A.A., Otto, B. et al. Variable clinical presentation of 2017;23(6):674–7.
Shwachman‐Diamond syndrome: update from the North American 110. Dy ABC, Tassone, F., Eldeeb, M., Salcedo‐Arellano, M.J., Tartaglia, N.,
Shwachman‐Diamond Syndrome Registry. J Pediatr, 2014;164(4):866–70. and Hagerman, R. Metformin as targeted treatment in fragile X syndrome.
92. Stepensky, P., Chacon‐Flores, M., Kim, K.H. et al. Mutations in EFL1, an Clin Genet, 2018;93(2):216–22.
SBDS partner, are associated with infantile pancytopenia, exocrine pancre- 111. Samocha‐Bonet, D., Debs, S., and Greenfield, J.R. Prevention and t­ reatment
atic insufficiency and skeletal anomalies in aShwachman‐Diamond like syn- of type 2 diabetes: a pathophysiological‐based approach. Trends Endocrinol
drome. J Med Genet, 2017;54(8):558–66. Metab, 2018;29(6):370–9.
93. Alter, B.P., Giri, N., Savage, S.A., and Rosenberg, P.S. Cancer in the National 112. Soares, H.P., Ni, Y., Kisfalvi, K., Sinnett‐Smith, J., and Rozengurt, E.
Cancer Institute inherited bone marrow failure syndrome cohort after fifteen Different patterns of Akt and ERK feedback activation in response to rapa-
years of follow‐up. Haematologica, 2018;103(1):30–9. mycin, active‐site mTOR inhibitors and metformin in pancreatic cancer
94. Toiviainen‐Salo, S., Durie, P.R., Numminen, K. et al. The natural history of cells. PLoS One, 2013;8(2):e57289.
Shwachman‐Diamond syndrome‐associated liver disease from childhood to 113. Monyak, R.E., Emerson, D., Schoenfeld, B.P. et  al. Insulin signaling
adulthood. J Pediatr, 2009;155(6):807–11 e2. ­misregulation underlies circadian and cognitive deficits in a Drosophila
95. Leung, D.H. and Yimlamai, D. The intestinal microbiome and paediatric fragile X model. Mol Psychiatry, 2017;22(8):1140–8.
liver disease. Lancet Gastroenterol Hepatol, 2017;2(6):446–55. 114. Obara, A., Fujita, Y., Abudukadier, A. et  al. DEPTOR‐related mTOR
96. Pianese, G. Beitrag zur Histologie und Aetiologie der Carcinoma. ­suppression is involved in metformin’s anti‐cancer action in human
Histologische und experimentelle Untersuchungen. Beitr Pathol Anat Allg liver cancer cells. Biochem Biophys Res Commun, 2015;460(4):
Pathol, 1896;142:1–193. 1047–52.
miRNAs and Hepatocellular
16 Carcinoma
Yusuke Yamamoto1, Isaku Kohama1, and Takahiro Ochiya1,2
1
Division of Molecular and Cellular Medicine, National Cancer Center Research Institute, Tsukiji,
Tokyo, Japan
2
Institute of Medical Science, Tokyo Medical University, Shinjuku, Tokyo, Japan

INTRODUCTION THE BASICS OF miRNA BIOLOGY


AND BIOGENESIS
Several types of small noncoding RNAs have been identified in
eukaryotic cells, including microRNAs (miRNAs), small inter- The sequences of miRNAs are highly conserved across species
fering RNAs (siRNAs), and Piwi‐interacting RNAs (piRNAs) [11]. Primary miRNA genes (pri‐miRNAs) are transcribed by
[1, 2]. Revolutionary advances in next‐generation sequencing the RNA polymerase II (Pol II) transcription units, and the
have made substantial contributions to the discovery of new transcripts are called pri‐miRNAs (Figure 16.1). The pri‐miRNAs
classes of small RNAs and the identification of a large number possess a cap and a poly‐A structure, as well as protein‐coding
of these nucleotides. miRNAs have a length of 18–24 n­ ucleotides, genes. The size of the pri‐miRNA usually exceeds 1 kilobase,
are endogenously expressed in almost all types of eukaryotic and it contains an RNA hairpin loop structure with a mature
cells, and primarily silence gene expression at the post‐ miRNA sequence [12]. In most cases, one pri‐miRNA contains
transcriptional level. In general, miRNAs are incorporated into one mature miRNA; however, some pri‐miRNAs include sev-
the RNA‐induced silencing complex (RISC) that functions as an eral mature miRNAs, such as the miR‐17–92 cluster and miR‐
RNA silencer and is located in the cytoplasm of cells [3]. Based 106a–363 cluster [13]. Some miRNA families, particularly the
on a computational prediction model, the expression of approxi- let‐7 family, are widely conserved in many species and include
mately 30% of all genes is influenced by miRNA‐induced gene more than 10 mature miRNAs in humans [14]. Some miRNAs
silencing [4]. Notably, miRNAs play important roles in regulat- are also located in the intron regions of protein‐coding genes,
ing nearly every physiological cellular process, and their dys- and intronic RNAs become pri‐miRNAs during the process of
regulation is strongly correlated with various types of disease, splicing; thus, the expression levels of these types of miRNAs
including cancer [3]. correlate with the parent gene expression [15].
The expression patterns and levels of miRNAs are tightly During miRNA biogenesis, pri‐miRNAs, which are transcribed
regulated in different types of cells during development; for by a Pol II RNA polymerase, contain hairpin loop structures
example, miR‐122 expression is limited to hepatocytes [5]. A (approximately 70 nucleotides) that are recognized by a
number of miRNAs have been shown to regulate cellular dif- DiGeorge syndrome critical region 8 (DGCR8) nuclear protein
ferentiation and proliferation under physiological conditions in the nucleus. DGCR8 interacts with Drosha [16], an RNase III
[6, 7]. Likewise, many miRNAs have been reported to play endonuclease, to form the microprocessor complex. In the com-
key roles in cancer initiation and progression, including inva- plex, Drosha cleaves the pri‐miRNAs at the base of stem‐loop
sion and metastasis processes in almost all types of cancers structures and creates 60–70 nucleotide stem‐loop intermedi-
[8–10]. The purpose of this chapter is to provide an overview ates, called a precursor miRNA (pre‐miRNA) [17]. The pre‐miRNAs
of the functions of miRNAs in hepatocellular carcinoma in the nucleus are transported to the cytoplasm by the nucleo-
(HCC) and their clinical applications, such as diagnostics cytoplasmic shuttler exportin‐5 (Exp‐5) [18]. In the cytoplasm, the
and therapeutics. hairpin structure of pre‐miRNA is cleaved by the RNase III

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
184 THE LIVER:  miRNAs IN HEPATOCELLULAR CARCINOMA

DNA

Transcribed by Pol RNA polymerase

Nucleus Pri-miRNA
(> 1 kb)
Processed by Drosha, DGCR8

Pre-miRNA
(60–70 nt)
Exported by Exp5
Cytoplasm

Mature-miRNA (22 nt) RISC

5’UTR 3’UTR

Protein coding sequence AAAAAAA

Translational inhibition
or mRNA degradation

Figure 16.1  The biogenesis and function pathways of miRNAs. Primary miRNA genes (pri‐miRNAs) are transcribed by the Pol II transcription
units, and the transcripts are called pri‐miRNAs with a cap and a poly‐A structure. In the process of miRNA biogenesis, pri‐miRNAs are recognized
by a DGCR8 nuclear protein in the nucleus. DGCR8 interacts with Drosha, an RNase III endonuclease, which cleaves the pri‐miRNAs at the base
of stem‐loop structures and creates a pre‐miRNA. The pre‐miRNAs in the nucleus are transported to the cytoplasm by Exp‐5. In the cytoplasm,
Dicer, an RNase III enzyme cleaves pre‐miRNA, yielding an imperfect miRNA duplex of approximately 22 nucleotides. The “guide” strand of the
duplex incorporates into RISC. Based on the “seed” sequence in miRNAs, the expression of target genes are inhibited by either translational
inhibition (partial complementarity) or mRNA degradation (perfect complementarity).

enzyme Dicer, yielding an imperfect miRNA duplex of approxi- (perfect complementarity) (Figure  16.1). A single miRNA is
mately 22 nucleotides [17]. The imperfect miRNA duplex con- considered to possess the capacity to control hundreds of target
tains a “guide” strand that is the antisense strand of the target genes and influence a wide range of gene pathways. These
gene sequence and a “passenger” strand. Basically, the “guide” unique characteristics indicate that miRNAs are potentially
strand of the duplex incorporates into the RNA‐induced silenc- critical modulators of nearly all physiological events and the
ing complex (RISC), although “passenger” strand of the duplex development of various diseases. Thus, one of the key issues in
might also function as an active miRNA in some cases. The miRNA biology is to identify the target genes. The initial pro-
RISC contains primary miRNA‐binding proteins, namely, a cess used to predict target genes is to observe the pairing of the
member of the argonaute protein family. In mouse and human, 3′ UTR of target genes and miRNA “seed” sequence located
four Ago (Ago I–IV) proteins are conserved [19]. Ago proteins from nucleotides 2 to 8 at the 5′ end of the miRNA. Several
possess a PAZ domain that is responsible for binding both sin- online bioinformatics tools have been developed to predict the
gle‐stranded and duplexed RNA; Ago II is a major protein in the miRNA target genes: Target Scan 7.1 (http://www.targetscan.
RISC with helicase and endonuclease activities [20]. In the pro- org/vert_72/), miRD (http://mirdb.org/), miRBase (http://www.
cess of RISC assembly, the “passenger” strand of the imperfect mirbase.org/help/targets.shtml), and miRWalk2.0 (http://zmf.
miRNA duplex that was cleaved by Dicer is removed and even- umm.uni‐heidelberg.de/apps/zmf/mirwalk2/). These websites
tually degraded. The “guide” strand is selectively incorporated predict both (i) the genes that are potentially targeted by a specific
into the RISC and acts as a gene silencer. When the “guide” miRNA and (ii) miRNAs that bind to a specific gene.
strand of the miRNA tightly binds to the Ago protein in the
RISC, the miRNA recognizes target genes through homologous
sequences at the 5′ end of the miRNA, which is designated the
“seed” sequence (generally, 6–8 nucleotides) [18]. Primarily, miRNAs IN HEPATOCELLULAR
miRNAs negatively regulate the expression target genes through CARCINOMA
an interaction between the “seed” sequence at their 5′ ends and
the 3′ untranslated region (3′ UTR) of their target genes. Based
on the complementarity of the “seed” sequence with the target
Hepatocellular carcinoma
gene sequence, the mature miRNA causes either translational The primary malignancies of the liver are mainly HCC, cholan-
inhibition (partial complementarity) or mRNA degradation giocarcinoma, and hepatic angiocarcinoma. More than 90% of
16: 
miRNAs AND HEPATOCELLULAR CARCINOMA 185

liver cancer cases are HCC, and it frequently develops in patients Although the earliest studies were mainly based on the
with hepatitis B virus (HBV) and hepatitis C virus (HCV) miRNA microarray platform, recent technological advances
infections and cirrhosis (80–90% of the HCC cases) [21]. Although in next‐generation sequencing have also contributed to
the treatment outcomes are continuously improving, HCC is the miRNA research in the field of cancer biology. Using both
third leading cause of cancer‐related death worldwide. Due to the comprehensive approaches, several miRNAs have been
higher prevalence rates of HBV and HCV, the incidence rates of identified as oncogenic and tumor‐suppressive miRNAs in
HCC are higher in Asia and Africa than elsewhere; the total HCC, and some of these miRNAs are also considered diag-
number of the HCC cases worldwide is expected to increase in nostic and prognostic biomarkers. As mentioned earlier,
the next few years. HCV is the most common cause of HCC in these miRNAs likely affect various pathways by inhibiting
Japan and Western countries, but in other Asian countries and hundreds of genes. Although many studies have been per-
Africa, HBV is the most common cause of HCC development. formed to investigate miRNA functions and their targets in
The main cause of hepatitis is a virus infection, but it is also other types of cancer, not specifically HCC, theoretically,
caused by alcoholism (alcoholic liver disease, ALD) and steato- the target genes identified for certain miRNAs in other can-
sis (non‐alcoholic steatohepatitis, NASH). HBV and HCV cers are potentially the same targets in HCC and thus the
infections cause chronic hepatitis, leading to abnormal cell pro- pathways affected by certain miRNAs would also be similar.
liferation and death. Hepatitis also gradually results in genetic Therefore, miRNA research in the field of HCC will likely
alterations in the liver. Many people who become infected with contribute not only to improving our understanding of
these viruses do not even know it. Some become asymptomatic miRNA functions in HCC but also to development of inno-
long‐term HBV or HCV carriers and develop liver fibrosis, vative and feasible therapeutic methods and novel diagnostic
which subsequently leads to liver cirrhosis within 20 years. In strategies for patients with HCC, particularly for the early
the last stage of the disease, the cirrhotic liver eventually devel- and accurate detection of HCC.
ops HCC. Approximately 80% of the HCC cases develop from
liver cirrhosis. In the process of HCC development from hepati-
tis via liver cirrhosis, hepatocytes repeatedly grow and undergo
Oncogenic roles of miRNAs in HCC
cell death, leading to the accumulation of genetic aberrations, A comprehensive analysis of miRNA profiling in HCC cases
such as point mutations, genomic deletions, and amplifications, provides clear evidence that miR‐21 functions as an oncogene
which are the main causes of HCC development [22]. during HCC development. Notably, miR‐21 is one of the miR-
When the patients have a sufficient hepatic function reserve, the NAs with the highest expression in primary HCCs, and its abun-
optimal treatment for HCC is partial surgical resection. Although dant expression is a hallmark of various types of cancers. As
the 5‐year survival rates after treatment have substantially improved miR‐21 expression is substantially increased in various types of
over the past few decades, the recurrence rates is still very high: cancers, it is designated as an “oncomiR.” One of the major
approximately 70%. Another HCC treatment is liver transplanta- functions of miR‐21 in HCC is to inhibit the expression of a
tion. Generally, patient who are eligible for liver transplantation tumor suppressor gene called phosphatase and tensin homolog
have multiple liver dysfunctions. Additionally, many HCC cases deleted from chromosome 10 (PTEN), and its expression is
are detected at an advance stage; thus, less than 40% of patients closely associated with the poor differentiation of tumor cells.
with HCC are eligible for surgery and transplantation. The PTEN gene has empirically been confirmed to be a direct
In general, radiological approaches, such as ultrasonography target of miR‐21. PTEN is a phosphatase that dephosphorylates
and computed tomography, are used for the diagnosis and sur- focal adhesion kinase (FAK). When FAK is phosphorylated, it is
veillance of HCC. However, these methods are not optimal to activated. The status of FAK activation correlates with aggres-
detect a small lesion in the liver. Other approaches for HCC sive tumor behavior in HCC [25]. Therefore, miR‐21‐mediated
screening are serological tests of tumor biomarkers, such as α‐ PTEN downregulation results in an increase in FAK phospho-
fetoprotein (AFP) and protein induced by vitamin K absence or rylation and an aggressive phenotype of HCC characterized by
antagonist‐II (PIVKA‐II). These methods are less invasive and invasion. The inhibition of miR‐21 decreases the growth rate of
can be readily applied for HCC screening; however, the low sen- HCC cells in soft agar and promotes apoptotic cell death [26].
sitivity and specificity of serological biomarkers are critical As expected, miR‐21 overexpression promotes the growth,
issues, particularly for the detection of early‐stage HCC [23, migration, and invasion of HCC cells [27]. The level of the
24]. Recently, expression profiling of miRNAs has been consid- miR‐21 transcript is increased by signal transducer and activator
ered as a novel biomarker to detect cancer. The first studies were of transcription 3 (STAT3), a major mediator of interleukin 6
performed using an miRNA microarray, which provided sub- (IL‐6) signaling. STAT3 is involved in tumor transformation by
stantial contributions to miRNA biology and cancer biology. In suppressing apoptotic signaling and directly binds to the pro-
these studies, miRNAs whose expression levels in cancer tis- moter region of the miR‐21 primary transcript to induce its
sues were decreased compared with patient‐matched normal tis- ­transcription [28].
sues, target oncogenes, and thus the miRNAs originally function Another well‐known oncogenic miRNA in HCC is the
as tumor suppressors. In contrast, miRNAs whose expression miR‐17–92 polycistronic cluster. This miRNA cluster contains
levels are increased in cancer target tumor suppressor genes and six miRNAs: miR‐17, miR‐18a, miR‐19a, miR‐20a, miR‐19b,
function as oncogenes in cancer development. Importantly, and miR‐92a, which are also considered “oncomiRs.”
these differential expression profiles of miRNAs are closely Overexpression of the miR‐17–92 cluster was confirmed in
associated with the clinical outcomes in patients with many 100% of human HCC cases, and some of these miRNAs, such
types of cancer, including HCC. as miR‐20a, were expressed in high levels in cirrhotic liver
186 THE LIVER:  miRNAs IN HEPATOCELLULAR CARCINOMA

tissues, suggesting that they might be important for the early tumorigenicity in vivo. In this study, TP53INP1 was identified
stage of HCC development [29]. The expression of the miR‐17– as one of the miR‐130b target genes. Thus, miR‐130b is one of
92 polycistronic cluster is positively regulated by the c‐Myc the key miRNAs regulating cancer stemness, including the
oncogene. The direct binding of c‐Myc to the promoter region self‐renewal capability and drug resistance, indicating that it is
in the miR‐17–92 polycistronic cluster has been reported [30]. also an oncogenic miRNA [35].
The function of the miR‐17–92 cluster in vivo was investigated The major issue in curing cancer is finding methods to pre-
using liver‐specific miR‐17–92 transgenic mice in combination vent cancer metastasis and recurrence after curative treat-
with a hepatic carcinogen (diethylnitrosamine) [31]. In this ment. Investigations of the molecular mechanisms of these
mouse model, liver‐specific miR‐17–92 transgenic mice devel- processes will provide new insights into clinical applications
oped a significantly greater number of HCC lesions than the for patients with HCC. In the field of miRNA biology, several
control mouse group treated with diethylnitrosamine. Similar miRNAs have been reported to function as upstream regula-
data were also obtained from in vitro experiments; this research tors of key molecules responsible for migration, invasion, and
group revealed that increased expression of the miR‐17–92 metastasis, which have important roles in determining liver
cluster clearly promoted the growth, colony formation, and cancer outcomes. Some oncogenic miRNAs have been
invasion of human HCC cell lines [31]. described as pro‐metastatic miRNAs. For example, overex-
Profiling of miRNA expression in HCC cell lines derived pressed oncogenic miR‐17–5p promotes the migration of
from chronic carriers of HBV and HCV and patients with HCC cells in vitro and in vivo by activating the p38 mitogen‐
nonviral‐associated HCC identified the differential expres- activated protein kinase (MAPK) pathway and increasing the
sion of several miRNAs: upregulation of miR‐222, miR‐221, phosphorylation of heat shock protein 27 (HSP27) [36]. Ding
and miR‐31, and downregulation of miR‐223, miR‐126, and et  al. [37] identified 22 miRNAs located at amplified or
miR‐122a. Upregulated miRNAs might function as onco- deleted genomic DNA regions in HCC. Among them,
genes, and in contrast, downregulated miRNAs might func- miR‐151, which is frequently amplified on chromosome
tion as tumor suppressors. Further analyses showed that the 8q24.3 and is coexpressed with its host gene FAK, is closely
deregulated expression patterns of miR‐223 and miR‐222 correlated with the intrahepatic metastasis of HCC.
clearly differentiated HCC cases from noncancerous liver tis- Overexpression of miR‐151 induced the migration and inva-
sues, regardless of the viral infection status [32]. In another sion of HCC cells in vitro and in vivo by inhibiting the expres-
study of HCV‐infected HCC cases, the miRNA profiling in a sion of RhoGDIA, a putative metastasis suppressor in HCC
set of 52 human primary liver tumors consisting of premalig- [37]. Another pro‐metastatic miRNA, miR‐143, has been
nant dysplastic liver nodules and HCCs using quantitative reported, and fibronectin type III domain‐containing 3B
real‐time polymerase chain reaction (qRT‐PCR), identified 10 (FNDC3B) is one of its direct target genes [38]. In this study,
upregulated and 19 downregulated miRNAs compared to nor- the intratumor administration of miR‐143 significantly
mal liver tissues [33]. A further analysis of these miRNAs enhanced HCC metastasis in an HCC cell‐xenografted mouse
confirmed that expression levels of miR‐122, miR‐100, and model. Additionally, a study using p21‐HBx transgenic mice
miR‐10a were significantly increased in HCV‐infected HCC showed that miR‐143 inhibition significantly blocked local
cases, indicating that they functioned as putative oncogenic liver metastasis and distant lung metastasis [38]. Oncogenic
miRNAs. Likewise, miRNA profiling of 89 HCC samples miR‐517 is also defined as a pro‐metastatic miRNA whose
using a ligation‐mediated amplification method identified overexpression causes metastatic dissemination in vivo [34].
subclasses of HCC (three main clusters: the wingless‐type A study profiling the expression of 156 miRNAs in HCC
MMTV integration site, interferon‐related, and proliferation) cases revealed that miR‐222 upregulation was commonly
through an unsupervised clustering analysis. In a subset of the observed in an HCC cohort. The suppression of miR‐222
proliferation subclass, poorly characterized miRNAs from expression led to reduced cell motility by negatively regulat-
chr19q13.42 were expressed at high levels. Among these ing AKT signaling. In this study, protein phosphatase 2A
miRNAs, enhanced expression of miR‐517a and miR‐520c subunit B (PPP2R2A) was identified as one of the miR‐222
promoted the growth, migration, and invasion of the HCC target genes using an in silico assay and luciferase reporter
cells in vitro. In particular, miR‐517 enhanced tumorigenesis assay of the PPP2R2A 3′UTR. Based on these data, miR‐222
and metastasis in an in vivo model. Thus, they are also consid- is also a pro‐metastatic miRNA [39].
ered oncogenic miRNAs [34]. The oncogenic miRNAs have also been reported to be key
As a new paradigm of cancer biology, many researchers factors in intrahepatic cholangiocarcinoma (ICC). For exam-
have focused on tumor‐initiating cells (TICs) or cancer stem ple, miR‐191 plays an important role in tumorigenesis. Next‐
cells (CSCs), which are considered to possess a higher self‐ generation sequencing technology with five pairs of ICC and
renewal capability and drug‐ and radio‐resistance. Ma et  al. matched‐to‐normal bile duct tissues revealed higher levels of
[35] identified a CD133‐positive population of HCC cells as miR‐191 expression in ICC than in adjacent normal bile duct
CSCs in HCC (approximately 1.3–13.6% of the cells in the tissues. Overexpression of miR‐191 promotes the growth,
tumor bulk) and also observed higher levels of miR‐130b invasion, and migration of ICC cells in vitro and in vivo.
expression in the CD133‐positive CSC population than in the Moreover, ten–eleven translocation 1 (TET1) was identified
CD133‐negative main population. Importantly, when miR‐ as a direct target gene of miR‐191, and reduced TET1 expres-
130b was overexpressed in the CD133‐negative HCC cells sion causes p53 promoter methylation, suppressing p53 tran-
from a lentivirus vector, the cells exhibited a higher resistance scription. The expression of miR‐191 correlates with poor
to chemotherapeutic agents and a significant increase in prognosis for patients with ICC [40].
16: 
miRNAs AND HEPATOCELLULAR CARCINOMA 187

Tumor‐suppressive effect of miRNAs on HCC at high levels in the mature liver and controlled the differentiation
of fetal hepatoblasts by directly targeting DNA methyltrans-
In the processes of cancer initiation and progression, both onco- ferase 1 (DNMT1), a major enzyme responsible for epigenetic
genes and tumor suppressor genes are essential components of silencing [47]. The authors also reported a beneficial effect of
the mechanism regulating the cell cycle. In miRNA biology, a miR‐148a on HCC suppression as a tumor suppressor miRNA.
number of reports identified specific miRNAs that negatively Regardless of the DNMT1 expression level, miR‐148a func-
regulate the cell cycle and function as tumor suppressors. Most tions as a tumor suppressor by controlling the expression of the
of tumor‐suppressive miRNAs target cyclin–cyclin‐dependent‐ c‐Met oncogene [47].
kinase (cyclin–CDK) complexes, a class of positive modulators The miR‐375 expression level was also downregulated in
of the cell cycle. A class of miRNAs were reported to function HCC tissues compared with adjacent nontumor tissues from
as major negative regulators of cell cycle progression; this group patients with HCC. Yes‐associated protein (YAP) is a potent
includes miR‐1, miR‐22, miR‐34a, miR‐122, miR‐375, and oncogenic driver, and miR‐375 overexpression decreases the
let‐7. The most interesting of these miRNAs is miR‐122, transcriptional activity of YAP, thus exerting a tumor‐suppres-
because it is abundantly expressed in the liver and comprises sive effect on HCC cells [48]. The research group also identified
over 70% of the total amount of miRNAs. stathmin 1 as one of downstream targets of miR‐223 using lucif-
The expression of miR‐122 is specifically and significantly erase reporter assay with the 3′ UTR; stathmin 1 expression
downregulated in human HCC cases compared with adjacent inversely correlates with miR‐223 expression in HCC cells [32].
normal liver tissues [41]. As shown in the study by Tsai et al., miR‐34a is one of the most well‐known tumor suppressor
miR‐122 functions as a tumor suppressor in HCC, and the miRNAs whose transcription is directly regulated by p53, the
authors experimentally identified 32 target genes related to cell guardian of the genome. A qRT‐PCR analysis of 83 HCC for-
movement, cell morphology, cell–cell signaling, and transcrip- malin‐fixed paraffin‐embedded (FFPE) tissue samples revealed
tion [41]. The miR‐122 expression level is downregulated in decreased miR‐34a expression in HCC samples compared to the
most HCC cases [42]. In this study, cyclin G1 was one of the patient‐matched liver tissues. As expected, ectopic expression
target genes of miR‐122; the expression levels of cyclin G1 of miR‐34a inhibits cell proliferation and induces apoptosis in
inversely correlated with miR‐122 expression. Obviously, cyc- HCC cells by influencing phospho‐ERK1/2 and phospho‐
lin G1 is a positive regulator of cell cycle progression, leading to STAT5 signaling. Thus, miR‐34a also functions as a tumor sup-
p53 downregulation and the induction of hepatocarcinogenesis, pressor in HCC [49].
and as miR‐122 inhibits cyclin G1, it functions as a tumor sup- Tumor suppressor miRNAs also function as anti‐metastatic
pressor in HCC. In contrast, the aberrant expression of miR‐122 miRNAs to some extent. These miRNAs mainly modulate the
has also been reported in patients with both HCC and HCV epithelial‐to‐mesenchymal transition (EMT). As a typical EMT‐
infection [33], presumably because miR‐122 expression in related miRNA, the transcription of the miR‐200 family is posi-
hepatocytes is essential for HCV RNA accumulation [43]. tively regulated by p53, which suppresses tumor progression
During liver development, miR‐122 expression is associated and metastasis. The upregulation of the miR‐200 family by p53
with four liver‐selective transcription factors, hepatocyte nuclear was confirmed by a profiling analysis of 92 primary HCCs and
factor 1α (HNF1α), HNF3β, HNF4α, and CCAAT/enhancer‐ 9 HCC cell lines. The miR‐200 family targets ZEB1 and ZEB2,
binding protein α (C/EBPα), and regulates the balance between which are master regulators of the EMT. Thus, p53 inhibits
the proliferation and differentiation of hepatocytes by directly EMT through miR‐200 family‐mediated ZEB1 and ZEB2
inhibiting CUTL1, a transcriptional repressor of genes specify- repression. Additionally, p53‐mediated activation of the
ing terminal differentiation in multiple cell lineages [44]. miR‐192 family suppresses ZEB2 expression. These miRNAs
Therefore, miR‐122 is thought to generally function as a tumor could function as anti‐metastatic miRNAs by repressing the
suppressor in HCC, but its roles in the specific circumstances EMT phenotype [50]. c‐Met is a transmembrane tyrosine kinase
remain unclear. receptor and an inducer of the EMT. Other tumor suppressor
Primary HCC exhibits a variety of genomic abnormalities, miRNAs, such as miR‐34a and miR‐23b, have also been
such as chromosomal instability, CpG hypermethylation, DNA reported to possess anti‐metastatic functions; miR‐34a sup-
rearrangements associated with HBV integration, DNA hypo- presses tumor invasion and migration by directly targeting
methylation, and, to a lesser extent, microsatellite instability c‐Met in HepG2 cells [51], and miR‐23b expression leads to
[45]. The hypermethylation of the promoter regions of tumor‐ urokinase‐type plasminogen activator (uPA) and c‐Met down-
suppressive miRNAs causes decreased expression and/or silenc- regulation and subsequently reduces cancer invasion and metas-
ing in cancer cells. The expression of miR‐1 is downregulated in tasis [52]. Notably, miR‐122 has been reported to function as an
primary HCC cases compared with patient‐matched liver tis- anti‐metastatic miRNA by directly inhibiting disintegrin and
sues, and miR‐1 is also a well‐known tumor suppressor miRNA. metalloproteinase domain‐containing protein 10 (ADAM10)
Initially, miR‐1 was identified as a methylated miRNA in HCC; [53] and ADAM17 [41]. ADAM family genes are closely asso-
thus, its expression is inhibited. Following treatment with 5‐ ciated with cancer metastasis; thus, miR‐122 functions as an
azacytidine (DNA hypomethylating agent) and/or trichostatin anti‐metastatic miRNA. Recently, Su et al. identified miR‐217,
A (histone deacetylase inhibitor), miR‐1 expression is restored whose expression level was decreased in highly invasive
in HCC cells. Overexpression of miR‐1 suppresses cell growth MHCC‐97H HCC cells and metastatic HCC tissues. Ectopic
and induces apoptotic cell death by directly targeting the FoxP1, expression of miR‐217 suppressed the invasion of MHCC‐97H
MET, and HDAC4 genes [46]. Likewise, miR‐148 was origi- cells by directly repressing E2F3 [54]. More recently, miR‐
nally identified as a hepatospecific miRNA that was expressed 501–3p has been described to be involved in the metastasis of
188 THE LIVER:  ROLES OF miRNAs IN HEPATITIS VIRUS INFECTION

HCC. Significantly lower expression of miR‐501–3p was the other hand, HBx regulates miRNA expression in the host
observed both in metastatic HCC cell lines and recurrent and cells. Notably, let‐7a has been identified to be negatively regu-
metastatic HCC tissue samples. Similar to other research reports lated by HBx; their expression levels are inversely correlated.
of anti‐metastatic miRNAs, ectopically enhanced expression of As let‐7a directly targets STAT3, HBx‐mediated downregula-
miR‐501‐3p suppresses the growth, migration, invasion and tion of let‐7a and upregulation of STAT3 support cell prolifera-
EMT of metastatic HCC cells. Lin‐7 homolog A (LIN7A) was tion, leading to hepatocarcinogenesis [62]. Additionally, an
identified as a direct target of miR‐501‐3p, and its inhibition miRNA microarray analysis revealed 10 miRNAs that were dif-
suppressed metastasis in HCC cells [55]. ferentially expressed between a stable HBV‐producing cell line
These tumor suppressor and/or anti‐metastatic miRNAs rep- (HepG2.2.15) and its control cell line (HepG2). High expres-
resent potentially useful prognostic biomarkers and treatment sion of miR‐501 has been observed in HBV‐producing cells.
strategies for HCC. The loss of miR‐501 expression suppresses HBV replication,
but does not affect the growth of HBV‐producing cells.
According to the results of a luciferase reporter assay, HBXIP,
an inhibitor of HBV replication, is a potential target of miR‐501.
Thus, miR‐501 would be an important miRNA regulating HBV
ROLES OF miRNAs IN HEPATITIS VIRUS replication in the host cells [63].
INFECTION miRNAs are involved in the physiology and function of the
immune system. Host innate antiviral immunity is the first line
Although miRNAs are involved in the process of HCC initiation of defense against hepatitis virus infection. The defense mecha-
and progression, they are also closely associated with infections nism is precisely controlled by a number of genes at multiple
with the hepatitis viruses HBV and HCV. HBV is a nuclear stages. The expression of miRNAs in the host cells was consist-
DNA virus, and no miRNAs have been identified in the HBV ently shown to play a key regulatory role in the process of HBV
genome, although one miRNA sequence in the HBV genome infection and its defense. Notably, miR‐155 is upregulated dur-
was computationally predicted [56]. In contrast, HBV influ- ing HBV infection and affects the host immune response;
ences the expression of miRNAs in the host cells to create a miR‐155 regulates the acute inflammatory response after the
favorable environment for HBV replication and immune eva- recognition of pathogens by toll‐like receptors. Overexpression
sion. HBV infection is basically characterized by two phases: of miR‐155 induces the expression of several interferon (IFN)‐
the acute and chronic phases. In the acute phase, HBV actively inducible antiviral genes and inhibits suppressor of cytokine
replicates while evading the host immune attack. In the chronic signaling 1 (SOCS1) expression, causing STAT1 and STAT3
phase, the HBV infection becomes dormant and the virus repli- phosphorylation in human hepatoma cells. Additionally,
cates stably and escapes from the immune system. In both infec- miR‐155 overexpression partially inhibits HBx gene expression
tion phases, miRNAs in the host cells play important roles in the in vitro [64]. Likewise, a study of 98 patients with HBV‐related
interaction between the host and virus. The most well‐studied HCC revealed that higher expression of miR‐200c blocked
miRNA involved in HBV infection is miR‐122, which is a liver‐ HBV‐mediated PD‐L1 expression by directly targeting the
specific miRNA that is abundantly expressed in hepatocytes. 3′UTR of PD‐L1 and resulted in a better prognosis. PD‐L1
Although miR‐122 is essential for the process of HCV infec- expression correlates negatively with miR‐200c expression in
tion, the loss of miR‐122 unexpectedly promotes the replication HBV‐related HCC cases [65].
of HBV [57, 58]. miR‐122 directly binds to the HBV genome In particular, the natural history of HBV infection in young
sequence, which is highly conserved, and suppresses the expres- children frequently exhibits a shift from an acute to chronic
sion of the viral genes. Moreover, HBV X protein (HBx), which infection as the virus becomes dormant in the host cells: infected
is essential for the transcription of the HBV genome, binds per- hepatocytes. When HBV forms a covalently closed circular
oxisome proliferator‐activated receptor‐gamma (PPARγ) and DNA (cccDNA) in the nucleus of the host cells, it stably sur-
suppresses the transcription of miR‐122, presumably inducing vives until the eventual reactivation of its life cycle, leading to
HBV replication [59]. the chronic infection phase [66]. Some miRNAs have been
The transfection of miRNA mimics showed that overexpres- reported to be required for the chronic HBV infection.
sion of miR‐1 increases HBV replication and upregulates HBV Methylation of the CpG islands in the cccDNA by DNA meth-
core promoter transcription, antigen expression, and progeny yltransferase 1 (DNMT1) prevents viral gene expression.
secretion. These effects of miR‐1 on HBV replication are not DNMT1 was also identified as a direct target of miR‐152, whose
directly mediated by targeting the HBV genome, as revealed by expression is frequently downregulated in HBV‐related HCCs
bioinformatics and luciferase reporter analyses. Thus, miR‐1 [67]. Thus, since the inhibition of miR‐152 causes global DNA
regulates the expression of several host genes to enhance HBV hypermethylation and increases the methylation levels of two
replication and reverse the cancer cell phenotype [60]. Moreover, tumor suppressor genes, glutathione S‐transferase pi 1 (GSTP1)
a screen with an miRNA library containing 2048 miRNAs iden- and E‐cadherin 1 (CDH1), miR‐152 functions as a tumor sup-
tified 39 miRNAs that repress HBV replication by examining pressor of the epigenetic aberrations associated with HBV‐
the intracellular and extracellular DNA and HBsAg levels. In related HCC [67]. In addition, a computational analysis of the
particular, miR‐204 substantially decreases both HBV DNA and HBV genome has identified seven sites that are potential targets
HBsAg levels in HCC cells. Rab22a was identified as one of the of human liver miRNAs. These miRNA target sites are located
targets of miR‐204, and the loss of Rab22a also suppresses in the cluster of a 995 bp segment within the viral polymerase
intracellular and extracellular HBV DNA expression [61]. On open reading frame (ORF) and the overlapping surface antigen
16: 
miRNAs AND HEPATOCELLULAR CARCINOMA 189

ORF and are conserved among the most common HBV sub- and thus the inhibition of the miRNA in HCC represents a
types [68]. The validation experiment using a luciferase reporter potential therapeutic strategy.
assay identified a direct interaction between miR‐125a‐5p and Kota and colleagues used miR‐26a as the target of HCC
the HBV sequence that clearly inhibited the reporter activity of therapy and reported the efficacy of an miRNA replacement in
the surface antigen [68]. HCC. The expression of miR‐26a is normally decreased in
Liver‐specific miRNAs also participate in the HCV infection HCC, and overexpression of miR‐26a induces cell cycle arrest
process. HCV is a single‐stranded RNA virus that infects hepat- by directly targeting cyclins D2 and E2. The systemic adminis-
ocytes and develops persistent infections. The major miRNA in tration of miR‐26a with an adeno‐associated virus (AAV) inhib-
the liver, miR‐122, is essential for HCV replication. The HCV its cancer growth and induces tumor‐specific apoptosis in a
RNA genome comprises a 5′ UTR, a long polyprotein ORF, and mouse model of HCC. The results of the delivery of miRNAs
a 3′ UTR. A genetic interaction between miR‐122 and the 5′ provided a new therapeutic strategy for HCC [71]. For the thera-
UTR of the viral RNA genome was identified in a bioinformat- peutics focusing on oncogenic miRNAs, one of the targets is
ics search, based on an analysis of the predicted miRNA‐bind- miR‐191. It has been reported as a therapeutic candidate miRNA
ing sites [57]. Notably, HCV RNA replicates in Huh‐7 cells for HCC therapy. Indeed, miR‐191 expression is increased by
expressing miR‐122, but not in the HepG2 cells that do not TCDD (2,3,7,8‐tetrachlorodibenzo‐p‐dioxin), a known liver
express miR‐122. When miR‐122 is inactivated with a 2′‐O‐ carcinogen, and it regulates a number of cancer‐related path-
methylated RNA oligonucleotide, the expression of the HCV ways. The administration of anti‐miR‐191 into the orthotopic
replicon and core protein are significantly suppressed [57]; thus, HCC xenograft inhibits the growth of tumor cells and reduces
miR‐122 represents an attractive target for the development of the tumor mass [72]. Another target is miR‐21, which is overex-
antiviral treatments. pressed in HCC and designated as an onco‐miR. Wagenaar and
colleagues developed potent and specific single‐stranded oligo-
nucleotide inhibitors of miR‐21 (anti‐miR‐21). Treatment with
anti‐miR‐21 significantly decreases cell viability in the majority
miRNAs AS THERAPEUTIC STRATEGIES of HCC cell lines by inducing apoptosis and necrosis. Similar to
FOR HCC the in vitro experiments, the effect of anti‐miR‐21 was con-
firmed in HCC tumor xenograft models [73]. Moreover, onco-
The biological significance of miRNAs in HCC initiation and genic miRNAs of the miR‐17 family were also targeted for
progression, as well as HBV and HCV viral replication, could pharmacological inhibition in an HCC model. Notably, miR‐17
provide new opportunities for developing the therapeutic tar- itself was blocked by a tough decoy inhibitor in HCC cell lines,
gets for HCC. When considering a therapeutic strategy for leading to a global derepression of direct targets of miR‐17. A
HCC, liver cirrhosis offers a better window for therapeutics lipid nanoparticle represented one of the most advanced plat-
because complete recovery of cirrhosis may prevent hepato- forms for the systemic delivery of an oligonucleotide to tumor
carcinogenesis. All three members of the miR‐29‐family are tissues in vivo and was used to encapsulate a potent anti‐miR‐17
significantly downregulated in the cirrhotic liver and are asso- family oligonucleotide, which was designated as RL01‐17(5).
ciated with TGF‐β‐mediated fibrosis [69]. Overexpression of RL01‐17(5) was systemically administered to orthotopic HCC
miR‐29b in mouse hepatic stellate cells (HSCs) decreases the xenografts, and significant tumor suppression was observed.
expression of α‐SMA, collagen I, and TIMP‐1. Overexpression Thus, these findings potentially represent proof‐of‐concept evi-
of miR‐29b in activated HSCs suppresses cell viability and dence for HCC therapies targeting the miR‐17 family [26].
colony formation, and causes cell cycle arrest in G1 phase by Therapeutic strategies based on miRNAs or anti‐miRNAs
downregulating cyclin D1 and p21cip1. Therefore, the admin- hold great promise due to their abilities to regulate a large num-
istration of miR‐29 might prevent hepatic fibrogenesis by tar- ber of genes and pathways. Prior to the therapeutic use of miR-
geting activated HSCs in the liver [70]. On the other hand, the NAs or anti‐miRNAs in the clinic, target identification and
HCV replication process is also targeted by miRNA‐based validation should be performed. Technological advances in the
therapeutics. Chronically infected chimpanzees were treated in vivo delivery systems, such as nanoparticles, virus vectors,
with a locked nucleic acid (LNA)‐modified oligonucleotide and use of exosomes, will enable efficient and safe miRNA‐ or
(SPC3649) complementary to miR‐122 and exhibited long‐ anti‐miRNA‐based gene therapy in HCC.
lasting suppression of HCV viremia, without side‐effects or
viral resistance in the chimpanzees. The prolonged effects of
SPC3649 are promising results for its use as an antiviral ther-
apy for HCV infection [43]. EXOSOMAL miRNAs IN HCC
Regarding miRNA therapeutics for HCC, a number of target
miRNAs have been proposed in both experimental and preclini- In 2007, an exemplary study by Valadi and colleagues reported
cal settings. Basically, two options for the target have been iden- that miRNAs and mRNAs were packaged in extracellular vesi-
tified: tumor suppressor miRNAs and oncogenic miRNAs. cles (EVs) [74]. These EVs are designated as exosomes,
When a target miRNA is a tumor suppressor, its expression level microvesicles, apoptotic bodies, and other extracellular parti-
is generally decreased in HCC tissues and increased in normal cles, according to their size, density, and secretion mechanism.
liver tissues. Thus, the administration of the miRNA into HCC Secreted EVs contain cytoplasmic components; they can form
tissues may exert therapeutic effects. In contrast, when a target at the plasma membrane by direct budding into the extracellular
miRNA is an oncogene, its expression level is increased in HCC environment and produce large (100–1000 nm), irregularly
190 THE LIVER:  EXOSOMAL miRNAs IN HCC

shaped microvesicles. In contrast, another type of particle, contrast, miR‐143, which functions as a tumor suppressor, is
exosomes, are thought to be released from multivesicular bodies expressed at high levels in normal prostate cells, and the EVs
(MVBs), whose size is 30–100 nm. EVs are secreted from cells from normal prostate cells contain miR‐143. EVs from normal
and transferred to another cell [74]. They are secreted from prostate cells are transferred to cancer cells to block cancer cell
almost all types of cell and play a key role in intercellular com- growth [78]. To date, a large number of studies have reported
munication by transporting RNA molecules and proteins that extracellular particles are an important source of miRNAs
(Figure 16.2). Thus, EVs are a novel tool for the exchange of in the circulation and suggest that EV‐derived miRNAs also
genetic information between cells, such as miRNAs. play key roles in carcinogenesis, including HCC initiation and
Notably, EVs have been detected in body fluids, including progression.
blood, saliva, urine, and breast milk. Although RNases are pre- According to Kogure and colleagues, HCC cells secrete EVs
sent in large amounts in body fluids such as the blood, circu- with distinct profiles of both RNAs and proteins compared with
lating extracellular RNAs are protected by the bilayer lipid their cells of origin. Similar to EVs derived from other cancer
membrane of EVs. Prior to this research, miRNAs were con- cells, the HCC cell‐derived EVs were confirmed to be internal-
sidered to function intracellularly, but they are even active and ized by other cells and modulate gene expression in recipient
function in the extracellular space via EVs. Indeed, cancer cells [79]. Liver‐specific miR‐122, a tumor suppressor miRNA,
cell‐derived EVs containing miRNAs are transferred into is also transferred via EVs between Huh7 and HepG2 human
endothelial cells and regulate angiogenesis [75]. Currently, hepatoma cells. In cell culture models, EV‐derived miR‐122
many studies have proposed that RNAs, including miRNAs, from Huh7 cells was internalized by miR‐122‐deficient HepG2
function as novel humoral factors in intercellular communica- cells and subsequently downregulated miR‐122 target mRNAs,
tion, and some have focused on the effect of EV‐derived miR- indicating that intercellular communication via EVs occurred
NAs in cancer biology [76]. between neighboring cells [80].
Cancer cells generally express oncogenic miRNAs at higher EVs secreted from HCC cells are also transferred to the sur-
levels and tumor‐suppressive miRNAs at lower levels, and EVs rounding fibroblasts. Cancer‐associated fibroblasts (CAFs) are
from the cancer cells basically reflect the miRNA expression well‐known to play a critical role in positively regulating the
profiles of the cells themselves. For example, p53‐mutated tumor microenvironment. The miRNAs sequencing of CAF‐
colon cancer cells secrete oncogenic miR‐1246 via EVs. EVs derived exosomes from patients with HCC revealed a signifi-
carrying miR‐1246 are delivered to the neighboring mac- cant loss of miR‐320a expression from CAF‐derived exosomes
rophages for reprogramming into a cancer‐promoting state, cre- [81]. An exogenous miRNA transfection experiment revealed
ating a favorable microenvironment for the cancer cells [77]. In that CAFs transferred miRNAs to HCC cells. Notably,

Fibroblast
Conversion to CAFs
Macrophage, etc.

Block cell
growth

Endothelial cells
1. Increased vascular
EVs containing permeability
HCC cells miRNAs 2. Angiogenesis

Figure 16.2  Intercellular communication in hepatocellular carcinoma via extracellular vesicles. HCC‐derived EVs modulate the properties of
microenvironmental cells such as immune cells, endothelial cells, fibroblasts, epithelial cells, and mesenchymal stem cells. Cancer cells use EVs to
build a favorable environment to form tumors. For example, HCC‐derived EVs containing miRNAs control CAF conversion in an endocrine manner
[82], increase vascular permeability in a paracrine manner [83], and induce angiogenesis [84]. In contrast, HCC cells also receive EVs from
surrounding cells. Macrophage‐derived EVs containing miRNAs block the proliferation of HCC cells. EVs could work as an intercellular
communication tool between neighboring cells in an autocrine manner [80].
16: 
miRNAs AND HEPATOCELLULAR CARCINOMA 191

miR‐320a functions as a tumor suppressor that inhibits HCC would be useful for establishing new diagnostic and/or prognos-
cell proliferation, migration, and metastasis by directly target- tic tools for HCC. Although the molecular landscapes of tumors
ing the PBX3 homeobox gene. Therefore, the transfer of normal were initially established using surgical or biopsy specimens,
stromal cell‐derived miR‐320a via EVs is probably important more recently, an analysis of circulating nucleic acids, known as
for the prevention of HCC development; however, CAF‐medi- a liquid biopsy strategy, has received increasing attention as a
ated HCC tumor progression is at least partially caused by the diagnostic tool for solid tumors. The liquid biopsy has several
loss of the tumor‐suppressive miR‐320a from exosomes [81]. advantages; for example, it is less invasive and can be applied to
Since exosomes function locally and systemically, EV‐derived any methods, such as the detection of cell‐free DNA (cfDNA)
miR‐1247–3p from highly metastatic HCC cells has been and miRNAs. A single biopsy provides spatially and temporally
described to regulate metastatic niche formation in the lung by limited information about the tumors; however, the liquid biopsy
converting normal fibroblasts to CAFs [82]. EV‐derived miR‐ provides the average genetic information and may be able to
1247–3p suppresses B4GALT3 expression, resulting in the reflect intratumor heterogeneity [87, 88].
activation of β1‐integrin–NF‐κB signaling in fibroblasts. Because the detection of tumor‐associated genomic DNA
Interestingly, serum miR‐1247‐3p levels in EVs correlate with mutations is useful for diagnostic decisions regarding treatment
lung metastasis in patients with HCC [82]. strategies, such as the selection of molecular target drugs, the
Furthermore, EV‐derived miR‐103 has been reported to cor- methods for detecting cfDNA were established for preclinical
relate with increased vascular permeability required for metas- use (e.g., CancerSEEK, which distinguishes eight cancer types,
tasis. Deep sequencing and qRT‐PCR validation identified a even at early stages) [89]. In addition to cfDNA, the liquid
correlation between a high serum miR‐103 level and a high biopsy approach is highly effective at detecting miRNAs, pro-
metastatic capacity in patients with HCC. When endothelial teins, and lipids packaged in EVs in the body fluids. An early
cells were retreated with EVs derived from hepatoma cells analysis of miRNAs detected in blood samples from patients
expressing miR‐103 at high levels, their permeability signifi- with HCC revealed that oncofetal miR‐500 is present at high
cantly increased, facilitating the transendothelial invasion of levels [90]. Therefore, the detection and/or quantification of
tumor cells [83]. Cancer cell‐derived miR‐103 transfer decreases serum miRNA levels using liquid biopsy represents a poten-
the integrity of endothelial junctions by directly targeting VE‐ tially feasible application for the diagnosis of HCC, assessments
cadherin (VE‐Cad), p120‐catenin (p120) and zonula occludens of prognosis, surveillance of recurrent tumors, and prediction of
1 [83]. In addition, an evaluation of serum miRNA levels drug responses.
revealed a high level of miR‐210‐3p in exosomes isolated from Deep sequencing followed by qRT‐PCR validation was
the sera of patients with HCC, and the miR‐210 level positively applied to identify serum biomarkers for HCC and identified
correlated with the microvessel density in HCC tissues [84]. By three miRNAs – miR‐25, miR‐375, and let‐7f – as biomarkers
directly targeting the SMAD4 and STAT6 genes, exosomal for HCC. The receiver operating characteristic (ROC) curve
miR‐210 enhanced angiogenesis, as confirmed by an in vitro analysis of these miRNAs yielded an area under the ROC curve
tubulogenesis assay using endothelial cells [84]. On the other (AUC) of 0.997, with a 97.9% sensitivity and 99.1% specificity
hand, macrophage‐derived EVs are taken up by HCC cells. Two in distinguishing HCC cases from controls [91]. A logistic
miRNAs, miR‐142 and miR‐223, are contained in EVs, and regression model was established with seven miRNAs (miR‐122,
transfer of these miRNAs contributes to the inhibition of the miR‐192, miR‐21, miR‐223, miR‐26a, miR‐27a, and miR‐801)
proliferation of HCC cells by inhibiting the expression of stath- by randomly separating samples into a training cohort and a
min‐1 and insulin‐like growth factor‐1 receptor (IGF1R). Thus, validation cohort to discover a plasma miRNA combination that
the transfer of EVs carrying specific miRNAs from immune discriminated HBV‐related HCC cases. The miRNA combina-
cells represents a potentially new defense system protecting tion discriminated the HCC cases from healthy individuals
against HCC initiation and progression [85]. (AUC: 0.941), patients with chronic hepatitis B (AUC: 0.842),
and patients with cirrhosis (AUC: 0.884) [92]. Significantly
higher plasma levels of a single miRNA, miR‐21, were observed
in patients with HCC than in patients with chronic hepatitis and
miRNAs AS DIAGNOSTIC BIOMARKERS healthy individuals. Interestingly, the ROC analysis of plasma
IN HCC miR‐21 levels displayed an AUC of 0.773, with a 61.1% sensi-
tivity and 83.3% specificity in distinguishing HCC from chronic
At the dawn of the use of miRNA profiling for cancer diagnos- hepatitis, and an AUC of 0.953 with an 87.3% sensitivity and
tics, whole transcriptomic analyses were applied to investigate 92.0% specificity in distinguishing HCC from healthy individu-
the molecular classification of the HCC tissues by determining als [93]. These results are better than a conventional plasma
the clinical and genetic properties of the tumor [34]. Within the HCC biomarker, AFP, suggesting that miR‐21 is a promising
last decade, a number of studies clearly indicated that miRNA miRNA biomarker [93].
profiling provides a specific miRNA expression fingerprint of In 2015, Lin and colleagues reported a comprehensive study
tissues stratified according to malignancy, risk factors, and of the serum miRNA profile for HCC diagnosis in five groups:
oncogene/tumor suppressor gene alterations, displaying the patients with HCC, patients with HBV‐related cirrhosis, patients
clinical and pathological features of HCC initiation and pro- with chronic hepatitis (HBV), inactive HBsAg carriers, and
gression; thus, miRNA profiles could be exploited as potential healthy individuals. Using six serum samples from patients with
cancer biomarkers [86]. Because the miRNA signature likely HCC and eight serum samples from patients with chronic hepa-
represents the HCC status, an understanding of miRNA profiles titis (control), 19 highly expressed miRNAs were identified
192 THE LIVER:  REFERENCES

with the miRNA microarray. Using a combination of four mod- biomarkers. Nevertheless, the published data are still inconsist-
els, such as a linear support vector machine, nonlinear support ent. Several explanations for this discrepancy are likely, such as
vector machine, linear discriminant analysis, and logistical the use of different detection platforms, normalization methods,
regression, the authors developed an miRNA classifier (seven sample preparation methods, and sample types (serum or
miRNAs: miR‐29a, miR‐29c, miR‐133a, miR‐143, miR‐145, plasma), as well as variability in the cohort size. Therefore,
miR‐192, and miR‐505) for detecting HCC [94]. This miRNA standardization is necessary to obtain a reliable liquid biopsy
classifier showed higher accuracy than the conventional AFP test for HCC. Furthermore, EVs have the potential to serve as a
method and detected small‐sized, early‐stage, and α‐fetopro- natural vector; thus, they could be applied to deliver specific
tein‐negative HCC cases in at‐risk patients. Moreover, the circu- miRNA molecules, proteins, or drugs to the disease site. In the
lating miRNA signature was applied to estimate the risk of HCC next decade, we will look ahead to the future of miRNA research
in patients with cirrhosis [95]. By comparing serum miRNA towards clinical applications of circulating miRNAs as diagnos-
profiles between 330 cirrhotic liver samples and 42 early‐stage tic, prognostic, and therapeutic tools.
HCC samples, the calculated score based on the expression lev-
els of five miRNAs in serum predicted patients with cirrhosis
who were at high and low risk of developing HCC. The follow‐
up study (median period: 752 days) showed a good prediction ACKNOWLEDGMENTS
accuracy (AUC: 0.725, P < 0.001) [95].
As described above, miRNA levels in blood samples are valu- This work was supported in part by the Japan Agency for
able and useful for the preclinical detection of HCC, providing Medical Research and Development (AMED): P‐CREATE;
patients with an opportunity to undergo curative resection and 17cm0106217h0002 and Development and New Energy
achieve a better prognosis [94]. Recently, diagnostic models and  Industrial Technology Development Organization;
based on serum miRNA levels have reported the successful dis- 16ae0101011h0003, a research grant from the Uehara Memorial
crimination of specific cancer types from other types of cancers Foundation and a research grant from the Naito Foundation.
[96] and the prediction of cancer metastasis [97], as well as the
differentiation of malignancy from benign tissues. Thus, the liq-
uid biopsy approach is a very promising method and a technol-
ogy that will likely be applied from the bench to the bedside in
REFERENCES
the near future.
1. Groszhans, H. and Filipowicz, W. Molecular biology: the expanding world
of small RNAs. Nature, 2008;451(7177):414.
2. Aravin, A.A., Hannon, G.J., and Brennecke, J. The piwi‐piRNA pathway
SUMMARY AND PERSPECTIVES provides an adaptive defense in the transposon arms race. Science,
2007;318(5851):761–64.
3. Bartel, D.P. MicroRNAs: target recognition and regulatory functions. Cell,
This chapter focused on the functions and potential clinical 2009;136:215–33.
applications, such as therapeutics and diagnostics, of miRNAs 4. Lim, L.P., Lau, N.C., Garrett‐Engele, P. et al. Microarray analysis shows that
in liver cancer. A brief overview of researchers’ progress in some microRNAs downregulate large numbers of target mRNAs. Nature,
improving our understanding of the function of miRNAs in hep- 2005;433(7027):769.
5. Chang, J., Nicolas, E., Marks, D. et al. miR‐122, a mammalian liver‐specific
atocarcinogenesis has also been provided. The research field of microRNA, is processed from hcr mRNA and may downregulate the high
miRNA biology in HCC has developed quickly and broadly affinity cationic amino acid transporter CAT‐1. RNA Biol, 2004;1(2):106–13.
from basic studies to clinical studies. The importance of miR- 6. Schratt, G.M., Tuebing, F., Nigh, E.A. et  al. A brain‐specific microRNA
NAs in basic biology for the promotion and inhibition of HCC regulates dendritic spine development. Nature, 2006;439(7074):283.
7. Chen, C.Z., Li, L., Lodish, H.F., and Bartel, D.P. MicroRNAs modulate
development is based on the unique and fascinating features of
hematopoietic lineage differentiation. Science, 2004;303(5654):83–86.
miRNAs, which influence thousands of genes and regulate a 8. Lu, J., Getz, G., Miska, E.A. et al. MicroRNA expression profiles classify
variety of pathways. For instance, a liver‐specific miRNA, human cancers. Nature, 2005;435:834–38.
miR‐122, constitutes approximately 70% of all the miRNA 9. Voorhoeve, P.M. A genetic screen implicates miRNA‐372 and miRNA‐373
molecules in the liver. As expected, miR‐122 is important for as oncogenes in testicular germ cell tumors. Cell, 2006;124:1169.
10. Mayr, C., Hemann, M.T., and Bartel, D.P. Disrupting the pairing between
liver development, and surprisingly is essential for the replica-
let‐7 and Hmga2 enhances oncogenic transformation. Science, 2007;
tion of HCV. In contrast, it functions as a tumor suppressor dur- 315:1576.
ing HCC development. Thus, one miRNA spatially and 11. Farh, K.K. Grimson, A., Jan, C. et al. The widespread impact of mammalian
temporally exerts multiple functions in different circumstances. microRNAs on mRNA repression and evolution. Science, 2005;310(5755):
Due to their abilities to regulate complex gene networks, miR- 1817–21.
12. Lee, Y., Jeon, K., Lee, J.T., Kim, S., and Kim, V.N. MicroRNA maturation:
NAs represent potential therapeutic targets in HCC, as evi- stepwise processing and subcellular localization. EMBO J, 2002;21:
denced by findings from studies overexpressing miR‐26 and 4663–70.
inhibiting miR‐17 [71, 73]. The recent and most critical break- 13. Truscott, M., Islam, A.B., and Frolov, M.V. Novel regulation and functional
through in the miRNA research field is the discovery of extra- interaction of polycistronic miRNAs. RNA, 2016;22:129–38.
cellular RNAs, such as EV‐packaged miRNAs. The detection of 14. Pasquinelli, A.E., Reinhart, B.J., Slack, F. et al. Conservation of the sequence
and temporal expression of let‐7 heterochronic regulatory RNA. Nature,
circulating miRNAs in the blood is performed using a microar- 2000;408(6808):86.
ray or deep sequencing, resulting in the diagnosis of HCC. As 15. Ruby, J.G., Jan, C.H., and Bartel, D.P. Intronic microRNA precursors that
shown, studies with a large cohort have identified HCC bypass Drosha processing. Nature, 2007;448(7149):83.
16: 
miRNAs AND HEPATOCELLULAR CARCINOMA 193

16. Gregory, R.I., Yan, K.P., Amuthan, G. et  al. The microprocessor complex 42. Gramantieri, L., Ferracin, M., Fornari, F. et al. Cyclin G1 is a target of miR‐
mediates the genesis of microRNAs. Nature, 2004;432(7014):235–40. 122a, a microRNA frequently down‐regulated in human hepatocellular car-
17. Lee, Y., Ahn, C., Han, J. et al. The nuclear RNaseIII Drosha initiates micro- cinoma. Cancer Res, 2007;67:6092–9.
RNA processing. Nature, 2003;425:415–19. 43. Lanford, R.E., Hildebrandt‐Eriksen, E.S., Petri, A. et al. Therapeutic silenc-
18. Meister, G. and Tuschl, T. Mechanisms of gene silencing by double‐stranded ing of microRNA‐122 in primates with chronic hepatitis C virus infection.
RNA. Nature, 2004;431:343. Science, 2010;327:198–201.
19. Chen, P.Y. and Meister, G. microRNA‐guided posttranscriptional gene regu- 44. Xu, H., He, J.H., Xiao, Z.D. et al. Liver‐enriched transcription factors regu-
lation. Biol Chem, 2005;386(12):1205–18. late microRNA‐122 that targets CUTL1 during liver development.
20. Song, J.J., Smith, S.K., Hannon, G.J., and Joshua‐Tor, L. Crystal structure of Hepatology, 2010;52:1431–42.
argonaute and its implications for RISC slicer activity. Science, 2004;305: 1434. 45. Herath, N.I., Leggett, B.A., and MacDonald, G.A. Review of genetic and
21. Siegel, R.L., Miller, K.D., and Jemal, A. Cancer statistics 2017. CA Cancer epigenetic alterations in hepatocarcinogenesis. J Gastroenterol Hepatol,
J Clin, 2017;67:7–30. 2006;21:15–21.
22. Forner, A., Llovet, J.M., and Bruix, J. Hepatocellular carcinoma. The Lancet, 46. Datta, J., Kutay, H., Nasser, M.W. et al. Methylation mediated silencing of
2012;379:1245–55. MicroRNA‐1 gene and its role in hepatocellular carcinogenesis. Cancer Res,
23. Oka, H., Tamori, A., Kuroki, T., Kobayashi, K., and Yamamoto, S. 2008;68:5049–58.
Prospective study of alpha‐fetoprotein in cirrhotic patients monitored for 47. Gailhouste, L., Gomez‐Santos, L., Hagiwara, K. et al. miR‐148a plays a piv-
development of hepatocellular carcinoma. Hepatology, 1994;19(1):61–6. otal role in the liver by promoting the hepatospecific phenotype and suppress-
24. Seo, S.I., Kim, H.S., Kim, W.J. et  al. Diagnostic value of PIVKA‐II and ing the invasiveness of transformed cells. Hepatology, 2013;58:1153–65.
alpha‐fetoprotein in hepatitis B virus‐associated hepatocellular carcinoma. 48. Liu, A.M., Poon, R.T., and Luk, J.M. MicroRNA‐375 targets Hippo‐signal-
World J Gastroenterol, 2015;21:3928–35. ing effector YAP in liver cancer and inhibits tumor properties. Biochem
25. Meng, F., Henson, R., Wehbe‐Janek, H. et  al. MicroRNA‐21 regulates Biophys Res Commun, 2010;394:623–7.
expression of the PTEN tumor suppressor gene in human hepatocellular can- 49. Dang, Y., Luo, D., Rong, M. et al. Underexpression of miR‐34a in hepatocel-
cer. Gastroenterology, 2007;133:647–58. lular carcinoma and its contribution towards enhancement of proliferating
26. Wagenaar, T.R., Zabludoff, S., Ahn, S.M. et  al. Anti‐miR‐21 suppresses inhibitory effects of agents targeting c‐MET. PLoS One, 2013;8:e61054.
hepatocellular carcinoma growth via broad transcriptional network deregula- 50. Kim, T., Veronese, A., Pichiorri, F. et al. p53 regulates epithelial–mesenchy-
tion. Mol Cancer Res, 2015;13:1009–21. mal transition through microRNAs targeting ZEB1 and ZEB2. J Exp Med,
27. Xu, G, Zhang, Y., Wei, J. et al. MicroRNA‐21 promotes hepatocellular carci- 2011;208:875–83.
noma HepG2 cell proliferation through repression of mitogen‐activated pro- 51. Li, N., Fu, H., Tie, Y. et  al. miR‐34a inhibits migration and invasion by
tein kinase‐kinase 3. BMC Cancer, 2013;13:469. down‐regulation of c‐Met expression in human hepatocellular carcinoma
28. Iliopoulos, D., Jaeger, S.A., Hirsch, H.A. et al. STAT3 activation of miR‐21 cells. Cancer Lett, 2009;275:44–53.
and miR‐181b‐1, via PTEN and CYLD, are part of the epigenetic switch 52. Salvi, A., Sabelli, C., Moncini, S. et al. MicroRNA‐23b mediates urokinase
linking inflammation to cancer. Mol Cell, 2010;39:493–506. and c‐met downmodulation and a decreased migration of human hepatocel-
29. Connolly, E., Melegari, M., Landgraf, P. et  al. Elevated expression of the lular carcinoma cells. FEBS J, 2009;276:2966–82.
miR‐17–92 polycistron and miR‐21 in hepadnavirus‐associated hepatocel- 53. Bai, S., Nasser, M.W., Wang, B. et al. MicroRNA‐122 inhibits tumorigenic
lular carcinoma contributes to the malignant phenotype. Am J Pathol, properties of hepatocellular carcinoma cells and sensitizes these cells to
2008;173:856–64. sorafenib. J Biol Chem, 2009;284:32015–27.
30. Hayashita, Y., Osada, H., Tatematsu, Y. et  al. A polycistronic microRNA 54. Su, J., Wang, Q., Liu, Y. et al. miR‐217 inhibits invasion of hepatocellular
cluster, miR‐17–92, is overexpressed in human lung cancers and enhances carcinoma cells through direct suppression of E2F3. Mol Cell Biochem,
cell proliferation. Cancer Res, 2005;65:9628–32. 2014;392:289–96.
31. Zhu, H., Han, C., and Wu, T. MiR‐17‐92 cluster promotes hepatocarcinogen- 55. Luo, C., Yin, D., Zhan, H. et al. MicroRNA‐501‐3p suppresses metastasis
esis. Carcinogenesis, 2015;36:1213–22. and progression of hepatocellular carcinoma through targeting LIN7A. Cell
32. Wong, Q.W., Lung, R.W., Law, P.T. et  al. MicroRNA‐223 is commonly Death Dis, 2018;9:535.
repressed in hepatocellular carcinoma and potentiates expression of 56. Jin, W.B., Wu, F.L., Kong, D. et al. HBV‐encoded microRNA candidate and
Stathmin1. Gastroenterology, 2008;135:257–69. its target. Comput Biol Chem, 2007;31:124–6.
33. Varnholt, H., Drebber, U., Schulze, F. et al. MicroRNA gene expression pro- 57. Jopling, C.L., Yi, M., Lancaster, A.M. et al. Modulation of hepatitis C virus
file of hepatitis C virus‐associated hepatocellular carcinoma. Hepatology, RNA abundance by a liver‐specific microRNA. Science, 2005;309:1577–81.
2008;47:1223–32. 58. Wang, S., Qiu, L., Yan, X. et  al. Loss of microRNA 122 expression in
34. Toffanin, S., Hoshida, Y., Lachenmayer, A. et al. MicroRNA‐based classifi- patients with hepatitis B enhances hepatitis B virus replication through cyc-
cation of hepatocellular carcinoma and oncogenic role of miR‐517a. lin G(1)‐modulated P53 activity. Hepatology, 2012;55:730–41.
Gastroenterology, 2011;140:1618–28. 59. Song, K., Han, C., Zhang, J. et al. Epigenetic regulation of MicroRNA‐122
35. Ma, S., Tang, K.H., Chan, Y.P. et  al. miR‐130b promotes CD133+ liver by peroxisome proliferator activated receptor‐gamma and hepatitis b virus X
tumor‐initiating cell growth and self‐renewal via tumor protein 53‐induced protein in hepatocellular carcinoma cells. Hepatology, 2013;58:1681–92.
nuclear protein 1. Cell Stem Cell, 2010;7: 694–707. 60. Zhang, X., Zhang, E., Ma, Z. et al. Modulation of hepatitis B virus replica-
36. Yang, F., Yin, Y., Wang, F. et al. miR‐17‐5p promotes migration of human tion and hepatocyte differentiation by microRNA‐1. Hepatology,
hepatocellular carcinoma cells through the p38 mitogen‐activated protein 2011;53:1476–85.
kinase‐heat shock protein 27 pathway. Hepatology, 2010;51:1614–23. 61. Naito, Y., Hamada‐Tsutsumi, S., Yamamoto, Y. et al. Screening of microR-
37. Ding, J., Huang, S., Wu, S. et al. Gain of miR‐151 on chromosome 8q24.3 NAs for a repressor of hepatitis B virus replication. Oncotarget,
facilitates tumour cell migration and spreading through downregulating 2018;9:29857–68.
RhoGDIA. Nat Cell Biol, 2010;12:390–9. 62. Wang, Y., Lu, Y., Toh, S.T. et al. Lethal‐7 is down‐regulated by the hepatitis
38. Zhang, X., Liu, S., Hu, T. et al. Up‐regulated microRNA‐143 transcribed by B virus x protein and targets signal transducer and activator of transcription
nuclear factor kappa B enhances hepatocarcinoma metastasis by repressing 3. J Hepatol, 2010;53:57–66.
fibronectin expression. Hepatology, 2009;50:490–9. 63. Jin, J., Tang, S., Xia, L. et al. MicroRNA‐501 promotes HBV replication by
39. Wong, Q.W., Ching, A.K., Chan, A.W. et al. MiR‐222 overexpression con- targeting HBXIP. Biochem Biophys Res Commun. 2013;430:1228–33.
fers cell migratory advantages in hepatocellular carcinoma through enhanc- 64. Su, C., Hou, Z., Zhang, C. et  al. Ectopic expression of microRNA‐155
ing AKT signaling. Clin Cancer Res, 2010;16:867–75. enhances innate antiviral immunity against HBV infection in human
40. Li, H., Zhou, Z.Q., Yang, Z.R. et  al. MicroRNA‐191 acts as a tumor pro- hepatoma cells. Virol J, 2011;8:354.
moter by modulating the TET1‐p53 pathway in intrahepatic cholangiocarci- 65. Sun, C., Lan, P., Han, Q. et al. Oncofetal gene SALL4 reactivation by hepa-
noma. Hepatology, 2017;66:136–51. titis B virus counteracts miR‐200c in PD‐L1‐induced T cell exhaustion. Nat
41. Tsai, W.C., Hsu, P.W., Lai, T.C. et al. MicroRNA‐122, a tumor suppressor Commun, 2018;9:1241.
microRNA that regulates intrahepatic metastasis of hepatocellular carci- 66. Ganem, D. and Prince, A.M. Hepatitis B virus infection – natural history and
noma. Hepatology, 2009;49:1571–82. clinical consequences. N Engl J Med, 2004;350:1118–29.
194 THE LIVER:  REFERENCES

67. Huang, J., Wang, Y., Guo, Y. et al. Down‐regulated microRNA‐152 induces 82. Fang, T., Lv, H., Lv, G. et al. Tumor‐derived exosomal miR‐1247–3p induces
aberrant DNA methylation in hepatitis B virus‐related hepatocellular carci- cancer‐associated fibroblast activation to foster lung metastasis of liver can-
noma by targeting DNA methyltransferase 1. Hepatology, 2010;52:60–70. cer. Nat Commun, 2018;9:191.
68. Potenza, N., Papa, U., Mosca, N. et al. Human microRNA hsa‐miR‐125a‐5p 83. Fang, J.H., Zhang, Z.J., Shang, L.R. et al. Hepatoma cell‐secreted exosomal
interferes with expression of hepatitis B virus surface antigen. Nucleic Acids microRNA‐103 increases vascular permeability and promotes metastasis by
Res, 2011;39:5157–63. targeting junction proteins. Hepatology, 2018;68:1459–75.
69. Roderburg, C., Urban, G.W., Bettermann, K. et  al. Micro‐RNA profiling 84. Lin, X.J., Fang, J.H., Yang, X.J. et al. Hepatocellular carcinoma cell‐secreted
reveals a role for miR‐29 in human and murine liver fibrosis. Hepatology, exosomal microRNA‐210 promotes angiogenesis in vitro and in vivo. Mol
2011;53:209–18. Ther Nucleic Acids, 2018;11:243–52.
70. Wang, J., Chu, E.S., Chen, H.Y. et al. MicroRNA‐29b prevents liver fibrosis 85. Aucher, A., Rudnicka, D., and Davis, D.M. MicroRNAs transfer from human
by attenuating hepatic stellate cell activation and inducing apoptosis through macrophages to hepato‐carcinoma cells and inhibit proliferation. J Immunol,
targeting PI3K/AKT pathway. Oncotarget, 2015;6:7325–38. 2013;191:6250–60.
71. Kota, J., Chivukula, R.R., O’Donnell, K.A. et  al. Therapeutic microRNA 86. Ladeiro, Y., Couchy, G., Balabaud, C. et al. MicroRNA profiling in hepato-
delivery suppresses tumorigenesis in a murine liver cancer model. Cell, cellular tumors is associated with clinical features and oncogene/tumor sup-
2009;137:1005–17. pressor gene mutations. Hepatology, 2008;47:1955–63.
72. Elyakim, E., Sitbon, E., Faerman, A. et al. hsa‐miR‐191 is a candidate onco- 87. Crowley, E., Di Nicolantonio, F., Loupakis, F., and Bardelli, A. Liquid
gene target for hepatocellular carcinoma therapy. Cancer Res, 2010;70: biopsy: monitoring cancer‐genetics in the blood. Nat Rev Clin Oncol,
8077–87. 2013;10:472–84.
73. Huang, X., Magnus, J., Kaimal, V. et al. Lipid nanoparticle‐mediated deliv- 88. Siravegna, G., Marsoni, S., Siena, S. et al. Integrating liquid biopsies into the
ery of anti‐miR‐17 family oligonucleotide suppresses hepatocellular carci- management of cancer. Nat Rev Clin Oncol, 2017;14:531–48.
noma growth. Mol Cancer Ther, 2017;16:905–13. 89. Cohen, J.D., Li, L., Wang, Y. et al. Detection and localization of surgically
74. Valadi, H., Ekström, K., Bossios, A. et  al. Exosome‐mediated transfer of resectable cancers with a multi‐analyte blood test. Science, 2018;359:
mRNAs and microRNAs is a novel mechanism of genetic exchange between 926–30.
cells. Nat Cell Biol, 2007;9:654–9. 90. Yamamoto, Y., Kosaka, N., Tanaka, M. et  al. (2009) MicroRNA‐500 as a
75. Kosaka, N., Iguchi, H., Hagiwara, K. et  al. Neutral sphingomyelinase 2 potential diagnostic marker for hepatocellular carcinoma. Biomarkers,
(nSMase2)‐dependent exosomal transfer of angiogenic microRNAs regulate 2009;14:529–38.
cancer cell metastasis. J Biol Chem, 2013;288:10849–59. 91. Li, L.M., Hu, Z.B., Zhou, Z.X. et  al. Serum microRNA profiles serve as
76. Kosaka, N., Yoshioka, Y., Fujita, Y. et al. Versatile roles of extracellular vesi- novel biomarkers for HBV infection and diagnosis of HBV‐positive hepato-
cles in cancer. J Clin Invest, 2016;126:1163–72. carcinoma. Cancer Res, 2010;70:9798–807.
77. Cooks, T., Pateras, I.S., Jenkins, L.M. et al. Mutant p53 cancers reprogram 92. Zhou, J., Yu, L., Gao, X. et al. Plasma microRNA panel to diagnose hepatitis
macrophages to tumor supporting macrophages via exosomal miR‐1246. Nat B virus‐related hepatocellular carcinoma. J Clin Oncol, 2011;29:4781–8.
Commun, 2018;9:771. 93. Tomimaru, Y., Eguchi, H., Nagano, H. et al. Circulating microRNA‐21 as a
78. Kosaka, N., Iguchi, H., Yoshioka, Y. et al. Competitive interactions of cancer novel biomarker for hepatocellular carcinoma. J Hepatol, 2012;56:167–75.
cells and normal cells via secretory microRNAs. J Biol Chem, 94. Lin, X.J., Chong, Y., Guo, Z.W. et al. A serum microRNA classifier for early
2012;287:1397–405. detection of hepatocellular carcinoma: a multicentre, retrospective, longitu-
79. Kogure, T., Lin, W.L., Yan, I.K. et  al. Intercellular nanovesicle‐mediated dinal biomarker identification study with a nested case‐control study. Lancet
microRNA transfer: a mechanism of environmental modulation of hepato- Oncol, 2015;16:804–15.
cellular cancer cell growth. Hepatology, 2011;54:1237–48. 95. Huang, Y.H., Liang, K.H., Chien, R.N. et al. A circulating microRNA signa-
80. Basu, S. and Bhattacharyya, S.N. Insulin‐like growth factor‐1 prevents ture capable of assessing the risk of hepatocellular carcinoma in cirrhotic
miR‐122 production in neighbouring cells to curtail its intercellular transfer patients. Sci Rep, 2017;7:523.
to ensure proliferation of human hepatoma cells. Nucleic Acids Res, 96. Yokoi, A., Matsuzaki, J., Yamamoto, Y. et al. Integrated extracellular micro-
2014;42:7170–85. RNA profiling for ovarian cancer screening. Nat Commun, 2018;9:4319.
81. Zhang, Z., Li, X., Sun, W. et al. (2017) Loss of exosomal miR‐320a from 97. Shiino, S., Matsuzaki, J., Shimomura, A. et al. Serum miRNA‐based predic-
cancer‐associated fibroblasts contributes to HCC proliferation and metasta- tion of axillary lymph node metastasis in breast cancer. Clin Cancer Res,
sis. Cancer Lett, 2017;397:33–42. 2019;25(6):1817–27.
Hepatocyte Apoptosis:
17 Mechanisms and Relevance
in Liver Diseases
Harmeet Malhi and Gregory J. Gores
College of Medicine, Division of Gastroenterology and Hepatology, Mayo Clinic, Rochester, MN, USA

INTRODUCTION signaling events and can be massive, which can result in organ
failure. Secondary signaling events include pathologic apoptosis‐
Apoptosis is a ubiquitous form of cell death occurring in human induced tissue inflammation, injury, and fibrosis. For example,
liver diseases (Figure  17.1). It has historically been defined in acute liver injury apoptosis is massive and correlates with
morphologically by the presence of cytoplasmic shrinkage outcome (i.e. liver transplantation or death) [4]. In chronic liver
(pyknosis), chromatin condensation, nuclear fragmentation injury apoptosis is continuous, modulates the inflammatory
(karyorhexis), the presence of plasma membrane blebbing, and response and promotes fibrogenesis, resulting in cirrhosis [5, 6].
the maintenance of an intact plasma membrane that retains its Hepatocyte apoptosis is evident in liver injury related to viral
integrity as the cell fragments into apoptotic bodies. Indeed, hepatitis, metabolic diseases, alcoholic steatohepatitis, autoim-
apoptotic bodies were first described in the liver in patients with mune hepatitis, and drug‐induced liver injury [7–9], emphasiz-
yellow fever where they were referred to as Councilman bodies. ing the shared pathogenic role of hepatocyte apoptosis in
A more current biochemical description of apoptosis is caspase‐ liver injury from multiple, varied, acute, and chronic insults.
dependent cell death [1]. Caspases are cysteine proteases that Apoptosis of other cellular compartments, such as sinusoidal
cleave proteins at sites next to aspartic acid residues. The above endothelial cells and stellate cells, also plays a role in liver
described apoptosis morphology is due to the cellular effects of injury. Here we discuss the signaling mediators and regulators
caspase activation. Caspase activation, and hence apoptosis, is a of hepatocyte apoptosis and the inclusion of injury stimulus‐
highly regulated form of cell death, with multiple checkpoints specific information within each mechanism.
and molecular mediators, activated via two distinct pathways,
the extrinsic pathway and the intrinsic pathway. The extrinsic
pathway is initiated via death receptor activation and the intrin-
sic pathway by intracellular perturbations that result in caspase THE EXTRINSIC PATHWAY
activation (Figure  17.2). In hepatocytes, both pathways con-
verge on mitochondria. Death receptors are cell surface transmembrane proteins that
Multiple intracellular molecules transmit apoptotic signals belong to the tumor necrosis factor/nerve growth factor (TNF/
and regulate the apoptotic signaling cascades, upstream and NGF) receptor superfamily, and are defined on the basis of
downstream of mitochondria [2]. Mitochondrial permeabiliza- ligand specificity (i.e. their affinity for tumor necrosis factor
tion is not only requisite but also sufficient for apoptosis; there- alpha (TNF‐α), Fas ligand (FasL), or tumor necrosis factor‐
fore, intracellular regulators downstream of mitochondrial related apoptosis‐inducing ligand (TRAIL) [10]. The extracel-
permeabilization such as caspase inhibitors cannot prevent cell lular N‐terminal domain binds their respective ligands; there is
death [3]. Unlike developmental apoptosis, which is carefully a membrane‐spanning region and then the intracellular C‐terminal
regulated in a spatiotemporal pattern and does not invoke domain, which contains a conserved sequence known as the
secondary events, pathologic apoptosis activates secondary death domain (DD). The ligand‐bound trimerized receptor

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
196 THE LIVER:  THE EXTRINSIC PATHWAY

(a) mitochondrial amplification categorizes hepatocytes as type


II cells, in contrast to type I cells in which caspase 8 or 10 can
directly activate caspase 3 and 7 without mitochondrial
involvement [12]. Caspase 8 proteolytically cleaves the proa-
poptotic BH3‐only protein of the Bcl‐2 family Bid to tBid
(truncated Bid), which leads to activation of Bax and Bak
(proapoptotic multidomain members of the Bcl‐2 family), and
pore formation in the outer mitochondrial membrane [13].
Multiple levels of signal transduction and amplification pre-
sent opportunities for regulation of death receptor‐mediated
apoptosis at many levels. Availability of cell surface receptor
and ligand is one level, for example the hepatocyte growth
factor (HGF) receptor Met associates with and regulates the
availability of Fas for binding its ligand [14]. Moreover, the
basal expression of death receptors in hepatocytes is low, and
known to increase in both acute and chronic liver diseases.
(b) Cellular caspase 8 (FLICE)‐like inhibitory protein (cFLIP)
can inhibit cytotoxic signaling by death receptors [15]. cFLIP
is an enzymatically inactive homolog of caspase 8 with con-
served structural homology in the DED that allows binding to
FADD. This binding precludes maximal cellular activation of
caspase 8. Pro‐ and antiapoptotic members of the Bcl‐2 fam-
ily regulate the extrinsic pathway by modulating the ability of
tBid to activate Bax and Bak (discussed later) [15].

Tumor necrosis factor‐α


TNF‐α is a circulating cytokine, primarily produced by cells of
the immune system, including Kupffer cells in the liver, although
it can also be produced by other cell types, such as hepatocytes.
Hepatocytes express both tumor necrosis factor receptor 1
(TNFR1), a 55 kDa protein, and tumor necrosis factor receptor
Figure 17.1  Hepatocyte apoptosis in non‐alcoholic steatohepatitis. 2 (TNFR2), a 75 kDa protein, but their functional significances
(a) Photomicrograph of a hematoxylin and eosin‐stained liver section differ [16]. TNFR1 is thought to mediate most of the biologic
from a dietary mouse model of obesity‐associated non‐alcoholic steato- effects of TNF‐α; it expresses a cytoplasmic death domain (DD)
hepatitis, demonstrating the presence of hepatocyte apoptosis. The and executes the apoptotic program by interacting with adaptor
black arrow points to an apoptotic hepatocyte. (b) The TUNEL stain
detects DNA strand breaks, a feature of cell death, in a corresponding
proteins [17] (Figure 17.3). Ligand‐activated TNFR1 generates
liver section from this mouse model of non‐alcoholic steatohepatitis. sequential signaling events referred to as complex I and
complex II. On binding TNF‐α, TNFR1 recruits the adaptor
protein tumor necrosis factor receptor‐associated death domain
complex brings together the DD, allowing recruitment of other (TRADD). Signaling then proceeds in two steps. The first step,
adaptor proteins to form a death‐inducing signaling complex or complex I, involves recruitment of tumor necrosis factor
(DISC). For death signaling, Fas‐associated protein with death receptor‐associated protein (TRAF‐2) and receptor‐interacting
domain (FADD) must be recruited to the DISC complex [10]. protein 1 (RIP1), leading to rapid activation of nuclear factor κB
FADD contains a death effector domain (DED) through which it (NFκB) [18]. NFκB transcriptionally activates expression of
binds inactive initiator caspases 8 and 10, in their procaspase prosurvival (e.g. Bcl‐xL, A1, XIAP, and cFLIP) and proinflam-
form. The procaspases form homodimers and undergo autopro- matory genes (e.g. interleukin 6). Complex I also activates the
teolytic cleavage with formation of active caspase 8 or 10 [11]. c‐Jun N‐terminal kinase (JNK) pathways. Complex I is internal-
This process of activation of initiator caspases is referred to as ized with dissociation of TRAF2, RIP1, and TRADD from the
the induced‐proximity model of caspase activation. Activation ligated receptor. TRADD then recruits FADD and procaspase 8
of initiator caspase 8 is essential for the transmission of apop- to initiate apoptotic signaling; this signaling pathway is referred
totic signals from death receptors to mitochondria. This occurs to as complex II. TRADD does not interact with TNFR2, nor
via the Bcl‐2 family protein Bid and is discussed below. does FADD directly interact with TNFR1. Therefore, TNF‐α/
In hepatocytes, mitochondrial permeabilization with TNFR1 signaling first leads to NFκB‐mediated transcriptional
amplification of the apoptotic cascade occurs in death recep- activation of prosurvival and proinflammatory genes prior to the
tor‐initiated apoptosis. This involves release of mitochondrial activation of apoptosis signaling. In cells resistant to NFκB, or
mediators of apoptosis which leads to the eventual activation in the presence of a transcriptional inhibitor such as actinomy-
of the effector caspases 3 and 7, with positive feedback cin D, which prevents the synthesis of prosurvival proteins, the
amplification of caspase 8 activation. The requirement of apoptotic effect of TNF‐α is unmasked.
17:  Hepatocyte Apoptosis: Mechanisms and Relevance in Liver Diseases 197

Figure 17.2  The extrinsic and intrinsic pathways of hepatocyte apoptosis. Mitochondrial permeabilization is required for hepatocyte apoptosis.
The extrinsic pathway is mediated by death receptors. Fas or TRAIL, upon ligation with their cognate receptors, activate events leading to mito-
chondrial permeabilization. The death‐inducing signaling complex is formed on the intracellular domain of ligated homotrimerized receptors in
conjunction with adaptor proteins, leading to caspase 8 activation, Bid cleavage, and activation of Bax and Bak. TNF‐α signaling pathway can
promote apoptosis by Bid‐induced lysosomal permeabilization. Intracellular perturbations such as ER stress, lysosomal permeabilization, or JNK
activate the intrinsic pathway of cell death. ER stress‐induced apoptosis is partly mediated by the transcription factor CHOP, which can upregulate
TRAIL‐R2 or Bim expression. JNK activation can be induced by TNF‐α, ER stress, or reactive oxygen species. These pathways are regulated by
the proapoptotic and antiapoptotic proteins of the Bcl‐2 family.

Necroptosis is a form of kinase‐regulated, caspase‐independ- express RIP3, and this may account for some of the incongruous
ent cell death best described following activation of TNFR1 by observations in mouse models. However, RIP1 has many roles,
TNF‐α, but characterized by morphology similar to necro- including kinase activity‐dependent and scaffolding function‐
sis – hence the name [19]. Necroptosis is known to occur only dependent roles that can promote apoptosis or inflammation.
under conditions of caspase 8 absence or inhibition and is medi- TNF‐α has pleiotropic effects in vivo, including hepatocyte
ated by the receptor‐interacting protein (RIP) family kinases, proliferation, liver inflammation, and modulation of hepatocyte
specifically RIP1 and RIP3. RIP1 has a DD and a caspase apoptosis. In a murine model of TNF‐α‐induced liver injury
recruitment domain which allows interactions at the DISC with (TNF‐α + d‐galactosamine), liver injury is Bax‐dependent [21].
the adaptor protein TRADD. That is why RIP1 is present in TNF‐α‐associated caspase 8 activation can also cause lysosomal
complex I. When caspase 8 is absent or inhibited, RIP1 is deu- permeabilization with release of intralysosomal cathepsin B
biquitinated and can interact with RIP3 to form the necrosome, into the cytosol, which causes mitochondrial dysfunction [22].
which also recruits mixed‐lineage kinase domain‐like pseudoki- Mice deficient in cathepsin B are protected from the injuri-
nase (MLKL), which is phosphorylated by RIP3 and translo- ous effects of TNF‐α [23]. c‐Jun N‐terminal kinase (JNK), a
cates to the plasma membrane where its trimerized form stress‐activated kinase, is activated by TNF‐α. Sustained activa-
mediates calcium influx and necroptosis [20]. When RIP1 is tion of JNK can lead to apoptosis by modulation of the Bcl‐2
polyubiquinated it promotes cell survival and inflammation. family of proteins. JNK can also transcriptionally activate death
The pathophysiologic contribution of necroptosis to the liver is receptor expression (i.e. TRAIL receptor 2/death receptor 5).
controversial due to the lack of RIP3 in normal hepatocytes. Furthermore, JNK can promote TNF‐α‐induced apoptotic sign-
Liver nonparenchymal cells and infiltrating immune cells aling at complex II by facilitating degradation of cFLIP, thus
198 THE LIVER:  THE EXTRINSIC PATHWAY

Figure 17.3  Complex I and complex II of tumor necrosis factor alpha signaling. Tumor necrosis factor receptor 1 (TNFR1), upon binding TNF‐α
on its extracellular domain, activates complex I and complex II. Complex I is formed by the adaptor proteins TNFR1‐associated death domain
protein (TRADD) and receptor‐interacting protein (RIP), which recognize and bind via their death domains (DD) and TNF receptor‐associated
factor (TRAF2) via its kinase domain or an intermediate domain. Complex I mediates the activation of nuclear factor κB (NFκB) and transient c‐Jun
N‐terminal kinase (JNK) activation. NFκB translocates to the nucleus transcriptionally, activating antiapoptotic and inflammatory genes, such as
cellular FLICE‐like inhibitory protein (cFLIP), Bcl‐xL, Mcl‐1, A1, and XIAP, which regulate apoptosis at multiple levels. Sustained JNK activation
requires the adaptor protein RIP and is mediated in part by oxidative stress. Complex II is formed by receptor dissociation of TRADD, RIP, and
TRAF2 and ligand‐independent recruitment of Fas‐associated death domain (FADD) via its DD. FADD contains a death effector domain (DED),
leading to recruitment and activation of procaspase 8. In select conditions in the absence or inhibition of caspase 8 signaling, receptor‐interacting
protein 1 (RIP1) interacts with RIP3, leading to the recruitment and phosphorylation of the mixed‐lineage kinase domain‐like pseudokinase
(MLKL), which leads to necroptosis by translocating to the cell membrane.

antagonizing an antiapoptotic TNF‐α‐induced NFκB target and inflammatory foci were decreased as compared to wild‐type
gene. Similarly, loss of cellular inhibitors of apoptosis proteins ethanol‐fed mice; TNFR2‐deficient mice developed liver
1 and 2, also antiapoptotic NFκB target genes, sensitizes carci- injury and apoptosis comparable to those in wild‐type controls
noma cells to TNF‐α‐mediated cytotoxicity [24]. TNF‐α can [27]. In ischemia reperfusion injury mice lacking TNFR1 and
lead to superoxide formation and caspase‐independent cell treated with pentoxyfylline, a pharmacologic TNF‐α inhibitor,
death by TRADD and RIP1‐mediated activation of Nox1 liver injury and apoptosis are significantly reduced [28]. Liver
NADPH oxidase leading to reactive oxygen species formation samples from patients with alcoholic steatohepatitis or non‐
[25]. This process is independent of FADD, and caspase 8 acti- alcoholic steatohepatitis demonstrate enhanced TNFR1
vation. Thus, a multitude of complex processes contribute to expression [29]. Serum levels of TNFR1 in patients with alco-
TNF‐α cytotoxicity. holic hepatitis are predictive of three‐month survival [30].
In experimental models of liver injury, a role for TNF‐α cell Thus, the TNF‐α cascade is activated in patients with many
death has been elucidated. Following partial hepatectomy, liver diseases, including fulminant hepatic failure, alcoholic
massive hepatocyte cell death occurs after completion of cell steatohepatitis, non‐alcoholic steatohepatitis, chronic hepatitis
cycle progression due to sustained TNF‐α signaling in mice C, and chronic hepatitis B [29, 31, 32]; it is indeed a hallmark
lacking tissue inhibitor of metalloproteinase 3 (Timp3), a of inflammatory changes in these conditions and likely con-
model characterized by abnormal chronically elevated TNF‐α tributes to hepatocyte apoptosis in vivo. Our understanding of
activity [26]. In TNFR1‐deficient ethanol‐fed mice, hepato- why the TNFR1‐initiated NFκB cell survival pathways fail in
cyte apoptosis, serum alanine aminotransferase levels (ALT), these diseases remains rudimentary.
17:  Hepatocyte Apoptosis: Mechanisms and Relevance in Liver Diseases 199

Fas affect HGF binding to its receptor Met. Pretreatment of cells


with HGF releases Fas from this complex, and enhances FasL
Fas (also known as Apo‐1, CD95) is ubiquitously expressed in binding and toxicity at lower concentrations of FasL. High con-
liver cell types. Hepatocytes are exquisitely sensitive to Fas‐ centrations of FasL are maximally toxic even in the absence of
induced apoptosis, and exogenously administered Fas agonistic HGF. Thus, the Met–Fas complex fine tunes and regulates the
antibody results in fulminant hepatic failure in mice [33]. Fas biologic availability of Fas in hepatocytes. In embryonic hepat-
signaling usually results in hepatocyte apoptosis, although there ocytes, Met prevents Fas‐induced cFLIP degradation, thus pre-
are reports of Fas‐induced proliferation of T cells and fibro- venting apoptosis.
blasts, chemokine secretion from macrophages, and Fas‐medi- In adult mice, genetic deficiency of Fas leads to hepatic
ated acceleration of liver regeneration after partial hepatectomy hyperplasia, in addition to enlargement of lymph nodes and
in mice [34]. Fas–Fas ligand (FasL) binding leads to receptor spleen [36]. The induction of fulminant hepatic failure in mice
oligomerization, bringing together the intracellular DD, recruit- by exogenous administration of Fas agonistic antibody is further
ment of FADD, and procaspase 8 or 10 at the DISC (Figure 17.4). regulated by the Bcl‐2 family of proteins. It can be abrogated by
This leads to activation and autoproteolytic activation of procas- overexpression of Bcl‐2 and enhanced by genetic inhibition of
pase 8 or 10, generation of tBid, activation of Bax and Bak, Bcl‐xL [37]. Genetic inhibition of Fas itself or Bid mitigates
mitochondrial permeabilization with eventual activation of cas- liver injury by Fas agonists [38]. Neutralization of Fas reduces
pase 3 and 7. Fas can be activated by soluble or circulating as warm ischemia/reperfusion‐related liver injury [39].
well as membrane‐bound FasL. FasL is expressed by cells of the Circulating levels of serum Fas are elevated in patients with
immune system, such as cytotoxic T lymphocytes (CTLs) and fulminant hepatic failure [40]. Levels of serum Fas vary by eti-
natural killer (NK) cells [35]. The liver is enriched in both these ology, and the highest levels occur in patients with drug‐induced
cell populations, therefore under constant “Fas attack.” However liver injury. Fas expression and apoptosis are enhanced in liver
Fas‐induced signaling is regulated at many levels. Cell surface samples from patients with chronic hepatitis C [41]. Circulating
expression of Fas, levels of FasL, and cFLIP inhibition of cas- levels of soluble Fas correlate with histologic activity, and along
pase 8 activation at the DISC are potential regulatory sites. Of with levels of caspase 3 activity, are predictive of response to
interest in hepatocytes is the sequestration of Fas by the hepato- therapy [42]. Similarly, in patients with chronic hepatitis B
cyte growth factor receptor (HGF) Met [14]. Met–Fas com- hepatocyte Fas levels and circulating levels of soluble Fas are
plexes prevent binding of FasL to Fas; however, Fas does not elevated [41, 43]. Fas expression is enhanced in liver samples

Figure 17.4  Fas and TRAIL receptor signaling: Fas and TRAIL receptors are activated by ligand binding, which leads to receptor oligomeriza-
tion, bringing together their conserved death domains (DD). The adaptor protein Fas‐associated death domain (FADD) binds to the trimerized
intracellular DD and via its death effector domain (DED) leads to activation of procaspase 8. Active caspase 8 leads to proteolytic cleavage of Bid
to tBid and downstream mitochondrial permeabilization via activation of Bax and Bak. Mitochondrial permeabilization leads to release of the
contents of the intermembrane space including cytochrome c, smac/DIABLO, Apaf‐1, and endonuclease G, culminating in the activation of caspase
3/7 and cleavage of cellular proteins.
200 THE LIVER:  THE INTRINSIC PATHWAY

from patients with non‐alcoholic fatty liver disease [7]. In disease due to a reduction in hepatocyte apoptosis [54, 55]. Free
experimental models of dietary and genetic fatty liver, steatotic fatty acids, which are elevated in the metabolic syndrome, tran-
livers are sensitized to exogenous Fas administration. Indeed, in scriptionally enhance TRAIL‐R2 expression in cell culture and
patients with non‐alcoholic fatty liver disease, the inhibition of render steatotic cells sensitive to TRAIL toxicity [51]. In acute
Fas by Met is diminished, providing another mechanism to hepatitis B‐induced liver failure in humans and experimental
explain the enhanced sensitivity to Fas‐induced hepatocyte adenoviral acute hepatitis in mice, TRAIL‐R2 expression is
apoptosis. Furthermore, free fatty acid treatment can increase enhanced, as is sensitivity to TRAIL. This occurs independently
Fas expression in vitro in cell culture models of hepatocyte stea- of Kupffer cells and NK cells, suggesting a hepatocyte‐gener-
tosis, sensitizing cells to Fas‐induced apoptosis. In the bile duct‐ ated paracrine loop for elimination of virally infected cells [56].
ligated mouse model of cholestatic liver injury, hepatocyte Circulating soluble TRAIL levels are elevated in patients with
apoptosis is mediated by Fas, and Fas‐induced apoptosis pro- chronic viral hepatitis B. Hepatitis B X antigen increases
motes hepatic fibrosis [44]. Toxic bile acids promote cell sur- TRAIL‐R1 expression in cell culture experiments, conferring
face expression of Fas, and can lead to ligand‐independent Fas sensitivity to TRAIL. In liver samples from patients with chronic
oligomerization and induction of hepatocyte apoptosis [45, 46]. hepatitis C, TRAIL‐R1 and TRAIL‐R2 expression and TRAIL‐
In bile salt‐mediated ligand‐independent hepatocyte apoptosis, induced apoptosis were enhanced [50]. Hepatitis C virus core
Fas phosphorylation is required for its translocation to the cell protein also selectively modulates cellular responsiveness to
surface; this can occur in a Yes kinase/epidermal growth factor TRAIL by promoting TRAIL‐induced Bid cleavage [57].
receptor‐dependent and JNK‐dependent manner [47].

Tumor necrosis factor‐related apoptosis‐


THE INTRINSIC PATHWAY
inducing ligand
The role of tumor necrosis factor‐related apoptosis‐inducing Intracellular stress leads to the activation of the intrinsic path-
ligand (TRAIL, also known as Apo‐2 ligand) and its receptors in way of apoptosis. Stress can be perceived and transduced by any
liver disease is an area with remarkable recent advances. TRAIL membrane‐defined organelle in the cell. For example, lys-
binds with several receptors [48]. TRAIL receptor 1 (TRAIL‐ osomes can mediate steatotic liver cell death, as can the endo-
R1/death receptor (DR) 4) and TRAIL receptor 2 (TRAIL‐R2/ plasmic reticulum (ER). DNA damage can lead to genotoxic
DR5/killer/TRICK2) are complete receptors and can induce stress and steatosis can activate JNK, also a mediator of the
apoptosis via caspase activation, similar to Fas [49]. This occurs intrinsic pathway of apoptosis. These processes converge on
via the adaptor protein FADD, recruitment of procaspase 8 and mitochondria and are transduced by the Bcl‐2 family of pro-
10 to the TRAIL receptor DISC, in a cFLIP‐regulated manner teins, and so are usually referred to as the Bcl‐2‐regulated or
(Figure  17.4). TRAIL receptor 3 (TRAIL‐R3/Apo‐3/TRAMP/ mitochondrial pathway of apoptosis. The Bcl‐2 family consists
WSL‐1/LARD, decoy receptor 1 (DcR1)) and TRAIL receptor of proapoptotic and antiapoptotic proteins, which are classified
4 (TRAIL‐R4, DR6, decoy receptor 2 (DcR2)) are incomplete into groups on the basis of the number of shared Bcl‐2 homol-
cell surface receptors and cannot stimulate apoptotic signaling. ogy (BH) domains. The proapoptotic proteins are structurally
Normal human hepatocytes, in situ and in vivo, are considered divided based on the number of shared BH domains into multi-
resistant to TRAIL‐induced apoptosis, though there are occa- domain (Bak and Bax, display BH1, 2, and 3 domains) and
sional reports of in vitro TRAIL‐induced hepatocyte apoptosis BH3‐only proteins (Bid, Noxa, Puma, Bim, Bmf, Bik, Hrk, and
[50]. This resistance to cell death may be secondary to cFLIP‐ Bad). The antiapoptotic proteins include Bcl‐2, Bcl‐xL, Bcl‐w,
induced inhibition of caspase 8 activation at the DISC or cell A1, Mcl‐1, and Boo, and share four BH domains with the excep-
surface expression/availability of TRAIL‐R1 or TRAIL‐R2. tion of Mcl‐1, which shares three BH domains with the rest of
However, diseased hepatocytes are sensitized to TRAIL‐induced the antiapoptotic Bcl‐2 family members. The liver expresses
apoptosis [51]. TRAIL also sensitizes to Fas‐induced hepato- Bcl‐xL and Mcl‐1; Bcl‐2 is not expressed by hepatocytes. Bax
cyte apoptosis by activating JNK and the proapoptotic BH3‐ and Bak are both abundantly expressed by hepatocytes. The
only protein Bim. antiapoptotic members of this family are located on the cyto-
TRAIL‐induced hepatocyte apoptosis has been demonstrated plasmic aspect of membrane‐bound organelles, primarily the
in cholestatic, viral, and metabolic liver diseases. Toxic bile mitochondria, though also on other organelles such as the ER.
acids transcriptionally regulate hepatocyte cell surface TRAIL‐ They protect cells from death by preventing spontaneous oli-
R2 expression in Fas‐deficient cells and inactivate cFLIP by gomerization of Bax and Bak, and may be necessary for survival
phosphorylation, thus dually sensitizing cells to TRAIL‐induced of certain cell types. Given this constitutive role in preventing
apoptosis [52]. In the bile duct‐ligated mouse model of choles- apoptosis, predictably, hepatocyte‐specific Bcl‐xL‐ or Mcl‐1‐
tasis, hepatocyte TRAIL‐R2 expression is enhanced and hepato- knockout mouse models demonstrate an increase in spontane-
cytes are sensitized to exogenously administered TRAIL [53]. ous hepatocyte apoptosis, liver injury, and fibrosis [58, 59]. Bax
By corollary, liver injury and hepatocyte apoptosis are signifi- and Bak are required for mitochondrial permeabilization, while
cantly reduced in TRAIL‐deficient mice following bile duct Bax is located in the cytosol and translocates to mitochondria
ligation. Steatosis is also associated with increased hepatocyte upon activation; Bak is a resident mitochondrial outer mem-
expression of TRAIL‐R2 and TRAIL‐R1, which imparts sensi- brane protein. The activation of Bax and Bak is regulated by
tivity to TRAIL toxicity, and TRAIL or TRAIL receptor dele- interactions between the antiapoptotic Bcl‐2 proteins and the
tion in mice improves dietary obesity‐associated fatty liver BH3‐only proapoptotic proteins. Several models have been
17:  Hepatocyte Apoptosis: Mechanisms and Relevance in Liver Diseases 201

proposed to explain the biochemical activation of Bax or Bak by apoptosis. SMAC inactivates post‐mitochondrial inhibitors of
proapoptotic BH3‐only proteins. Using Bim as an example, apoptosis proteins (IAP). Cytosolic cytochrome c, apoptotic
upon activation, Bim is released from the dynein motor com- protease‐activating factor 1 (Apaf) and ATP form a complex
plex, and can directly engage and activate Bax and Bak. called the apoptosome, leading to activation of procaspase 9 and
Alternatively, Bim can bind and negate the inhibitory effect of effector caspases 3 and 7 [64]. These effector caspases cleave
Bcl‐2 or Bcl‐xL, releasing Bax and Bak from inhibition by these over 500 substrates resulting in cellular demolition.
proteins (the derepression model). Cytokeratin 18 is a structural protein expressed in most epi-
Although the large number of BH3‐only proteins imparts thelial cells that is cleaved by caspase 3 at aspartate positions
redundancy, its primary effect is to impart stimulus specificity. 238 and 396. The fragment generated by this cleavage, cytoker-
For example, free fatty acids activate Bim and Puma [60]; Puma atin 18–aspartate 396 (CK18‐asp396) forms a neo‐epitope that
and Noxa are target genes of the tumor suppressor p53 [61]. is recognized by the M30 antibody. This neoepitope can be
Cytosolic Bid must be cleaved by active caspase 8 or 10 to detected in apoptotic tissues as well as serum by a commercially
generate truncated Bid (tBid), which translocates to the mito- available ELISA. Indeed circulating levels of CK18‐asp396 are
chondria to activate Bak or Bax. Bim is sequestered by the elevated in patients with liver injury and can correlate with
microtubule‐associated dynein motor complex in the cytosol, outcome [4]. Thus this biomarker presents a noninvasive, sim-
from which it is dissociated by proapoptotic stimuli. For exam- ple, and mechanistic tool to monitor progress and response to
ple, JNK‐mediated Bim phosphorylation promotes its mito- therapy in liver injury.
chondrial translocation. Bim is also regulated transcriptionally,
and known to be increased in ER stress‐induced apoptosis.
Cytosolic Bad is bound to the death inhibitory protein 14‐3‐3.
Lysosomes
Activating and inhibitory phosphorylation events have been Lysosomes are intracellular organelles with acid intravesicular
described for both Bim and Bad. pH that contain lysosomal proteases, known as cathepsins [65].
Cathepsin B and D, two of 11 known human cathepsins, are
stable and active at neutral pH. Methodical dissection of path-
Mitochondria ways that mediate intracellular death signals demonstrates that
In addition to the metabolic functions of mitochondria, hepato- lysosomes can be involved in the intrinsic pathway of cell death.
cytes require mitochondria to die. The mitochondrial intermem- Typically, lysosomal permeabilization, when it mediates apop-
brane space sequesters a number of proapoptotic proteins tosis, is selective and partial and is observed upstream of mito-
including cytochrome c, SMAC/DIABLO (second mitochon- chondrial permeabilization. Cathepsin B‐induced mitochondrial
drial activator of caspase/direct IAP‐binding protein with low permeabilization can occur via caspase 2 (in mice) and via pro-
pI), HtrA2/Omi, AIF (apoptosis‐inducing factor), and endonu- teolytic cleavage of Bid similar to death receptor‐induced acti-
clease G [2, 15]. Current models suggest that active Bax or Bak vation of Bid [66, 67]. Indeed Bid also links death receptors to
form pores in the outer mitochondrial membrane leading to lysosomal permeabilization; providing crosstalk between death
mitochondrial outer membrane permeabilization (MOMP) and receptors and their engagement of the lysosomal and mitochon-
release of these mediators into the cytosol [62]. In the absence drial pathways [66]. Bax activation by intracellular stress can
of Bak and Bax, cells are resistant to apoptotic stimuli; thus, also result in lysosomal permeabilization [68]. Cathepsin D lev-
Bak and Bax are critical effectors of MOMP. Though previously els were elevated in serum from patients with fulminant hepatic
thought to be a rapid and complete phenomenon wherein all failure as well as chronic hepatitis [69, 70]. Cathepsin B‐defi-
mitochondria within a cell would experience MOMP within cient mice are resistant to TNF‐α‐induced hepatocyte apoptosis
minutes, recent work has challenged this paradigm by describ- [23]. In models of cellular steatosis, cathepsin B inhibition pre-
ing partial MOMP in cells undergoing apoptotic stress. Since vents mitochondrial permeabilization and apoptosis. In cathep-
partial MOMP does not result in apoptosis, this type of cellular sin B‐deficient mice liver apoptosis, injury, and fibrosis are
stress has been named sublethal stress. Inactive Bak and Bax diminished following bile duct ligation [71]; liver apoptosis and
support mitochondrial fusion, whereas active Bak and Bax shift injury are abrogated in ischemia reperfusion injury as well [72].
mitochondrial dynamics toward fission. MOMP can also occur
secondary to permeability transition (PT) of the inner mitochon-
drial membrane (IMM) via the PT pore (PTP). The identity of
Endoplasmic reticulum
the PTP has received much attention and several candidate pro- The ER has an inbuilt mechanism to cope with excess or altered
teins have been evaluated, including adenine nucleotide trans- unfolded proteins that serves to correct the inciting imbalance.
porter and the voltage‐dependent anion channel. Current This process is termed the unfolded protein response (UPR).
evidence suggests that the PTP is formed from FOF1 ATP syn- The UPR can also be activated by stimuli that affect the function
thase dimers in the IMM [63], though these is some evidence to of the ER, such as calcium depletion, glycosylation inhibition
suggest that spastic paraplegia 7 could also form the PTP. (tunicamycin), ultraviolet radiation, and insulin resistance. The
Opening of the permeability transition pore leads to rapid fluxes ER stress response consists of a series of compensatory pro-
of ions and water, dissipation of the mitochondrial inner trans- cesses to correct both the excess and the stress of the unfolded
membrane potential, swelling of the mitochondria, and outer proteins. Global translation is attenuated to reduce the func-
mitochondrial membrane rupture leading to the release of the tional protein load of the ER. There is also selective translation
contents into the intermembrane space. MOMP releases inter- of UPR target genes aimed at protecting the ER [73, 74]. The
membrane contents into the cytosol and commits the cell to transducers of ER stress are membrane proteins that have an ER
202 THE LIVER:  THE INTRINSIC PATHWAY

luminal domain and a cytosolic domain. Inositol‐requiring pro- species, free fatty acids, and bile acids [88, 89]. JNK involve-
tein 1 alpha (IRE1α) and protein kinase RNA‐like ER kinase ment in apoptosis is temporally regulated and stimulus specific
(PERK) auto‐transphosphorylate when released from the ER [90]. The same inciting stimulus (e.g. TNF‐α) can induce bipha-
chaperone BiP/Grp78. IRE1α possesses endoribonucleolytic sic JNK activation mediated by distinct intracellular pathways.
activity leading to excision of an intron within X‐box binding Transient and early JNK activation promotes survival; and sus-
protein 1 (XBP1) mRNA to generate spliced XBP1 (sXBP1), a tained and late activation of JNK promotes apoptosis [91]. In
transcription factor that activates a subset of UPR target genes. the case of TNF‐α, production of reactive oxygen species medi-
IRE1α also recruits TRAF2, leading to JNK activation. PERK ates the delayed and sustained activation of JNK. Other stimuli,
phosphorylates and inactivates the eukaryotic translation initia- such as toxic free fatty acids, result in early and sustained JNK
tion factor 2α (eIF2α), resulting in global translation attenuation activation, culminating in apoptotic signaling [92]. JNK‐stimu-
with selective translation of activating transcription factor 4 lated proapoptotic signaling converges on mitochondria via the
(ATF4), which leads to transcription of C/EBP‐homologous activation of Bax and Bak. In the absence of Bax and Bak, JNK‐
protein (CHOP), and the ER chaperone BiP/Grp78. Activating induced cell death is mitigated. Furthermore, mitochondrial per-
transcription factor 6 (ATF6) is cleaved within the ER mem- meabilization and release of cytochrome c are abolished in cells
brane, generating an ATF6 fragment that translocates to the derived from mice lacking Jnk1 and 2 genes, in response to
nucleus and activates a subset of UPR target genes. It is not stimuli that cause intracellular stress [90]. JNK‐mediated
known if ATF6 also regulates apoptotic signaling. phosphorylation of pro‐ and antiapoptotic proteins upstream of
ER stress also activates a negative feedback regulatory loop mitochondria also regulates apoptotic sensitivity. JNK can
that terminates the UPR; however in the setting of sustained ER phosphorylate and activate the BH3‐only proteins; for example,
stress, proapoptotic signaling occurs [75]. Bax and Bak both Bim phosphorylation releases it from binding to the dynein
bind to the cytoplasmic domain of IRE1α, and in cells lacking motor complex and promotes apoptosis [93]. Sustained JNK
Bax and Bak, IRE1 stress‐generated JNK activation and XBP1 activation promotes caspase 8 formation at the DISC by activa-
splicing are reduced [76], thus linking the core apoptotic tion of the E3 ubiquitin ligase Itch, which ubiquinates and
machinery to ER stress response. Bax and Bak localize on the degrades cFLIP, promoting liver cell death [94]. JNK can phos-
ER membrane, in addition to mitochondrial membranes. In cells phorylate the antiapoptotic proteins Bcl‐2, Bxl‐xL, and Mcl‐1,
lacking both Bax and Bak, the ER is depleted of calcium and and the proapoptotic proteins Bmf and Bad.
unable to respond to certain death stimuli [77]. The proapop- JNK1 and JNK2 can both mediate liver injury in a stimulus‐
totic transcription factor CHOP can increase Bim expression specific manner. In a murine model of steatohepatitis induced
transcriptionally and by inhibiting its proteasomal degradation, by methionine‐ and choline‐deficient diet, JNK1 plays a pre-
leading to Bim‐dependent ER stress‐induced apoptosis [78]. dominant role. In high‐fat diet‐induced obesity and genetic
CHOP can also upregulate TRAIL‐R2 expression, sensitizing obesity in mice, JNK was activated; this was found to be pre-
cancer cells to TRAIL‐induced apoptosis [79]. dominantly JNK1, though JNK2 plays a role that is unmasked
The involvement of the ER stress‐induced apoptotic pathway in the absence of JNK1 [95]. In free fatty acid‐based cellular
in liver diseases is an area of emerging research. In the bile duct‐ models of hepatocyte steatosis, JNK2 is the predominant iso-
ligated mouse model of cholestasis, an early and transient form that mediates apoptosis [92]. Oleic acid, a minimally
induction of CHOP expression is observed [80]. Mice deficient toxic free fatty acid, also sensitizes steatotic hepatocytes to
in CHOP are protected from hepatocyte apoptosis, liver injury, TRAIL‐induced apoptosis by JNK‐dependent transcriptional
and liver fibrosis. In cell culture, the toxic bile acid glycocheno- upregulation of the death receptor TRAIL‐R2 [51]. This mech-
deoxycholic acid also induces ER stress and CHOP expression anism is shared by toxic bile acids, which also sensitize hepato-
in isolated rat hepatocytes [81]. In transgenic mice expressing cytes to TRAIL‐induced apoptosis by transcriptionally
hepatitis C viral core and E2 proteins, hepatocyte apoptosis is activating TRAIL‐R2 expression in a JNK‐dependent manner
associated with CHOP expression [82]. Cycloheximide, an [53, 96]. Liver injury induced by ischemia reperfusion is also
inhibitor of protein synthesis, induces ER stress, induction of mediated by JNK, and pharmacologic inhibition of JNK in
CHOP expression and apoptotic hepatocyte cell death in rat liv- donor livers improved graft survival and decreased apoptosis
ers [83]. In non‐alcoholic fatty liver disease, markers of ER after orthotopic liver transplantation [97]. In acetaminophen‐
stress were variably activated [84]. Toxic saturated fatty acids induced acute liver injury JNK activation was robust and sus-
also induce ER stress and apoptosis in liver cell lines [85]. In a tained, and led to Bax translocation to mitochondria and poor
mouse model of alcohol‐induced liver injury, CHOP‐deficient animal survival. Pharmacologic inhibition of JNK decreased
mice are protected from hepatocyte apoptosis, though able to liver injury and hepatocyte cell death and improved survival;
mount an ER stress response [86]. utilizing genetically deficient models of Jnk1 or Jnk2 it was
demonstrated that both mediate liver injury, though JNK2 was
predominant. JNK activation was observed in hepatocytes in
c‐Jun N‐terminal kinase
human liver samples from patients with acetaminophen‐
Given the role of JNK in multiple models of cell death, it war- induced acute liver failure. JNK inhibition was more effective
rants a separate discussion as a final common cell death media- in decreasing hepatocyte cell death than N‐acetylcysteine in a
tor. JNK1 and 2 are ubiquitously expressed, including liver, murine model of acetaminophen‐induced liver injury. In a
whereas JNK3 is not expressed in the liver [87]. JNK activation murine model of TNF‐α‐induced liver injury utilizing galac-
occurs downstream of kinase cascades that can be activated by tosamine and lipopolysaccharide, JNK2 mediated caspase 8
multiple stimuli including TNF‐α, IRE1α, reactive oxygen activation and mitochondrial permeabilization.
17:  Hepatocyte Apoptosis: Mechanisms and Relevance in Liver Diseases 203

THE CONSEQUENCES OF HEPATOCYTE In conclusion, hepatocyte apoptosis is a key mediator of liver


APOPTOSIS injury and inflammation in most forms of liver disease. Multiple
apoptotic pathways are activated by a given injurious stimulus
Apoptosis, inflammation, and injury are in some ways insepara- in a vulnerable hepatocyte. The predominant signaling pathway
ble, and it is difficult sometimes to dissect the primary event. that results in mitochondrial dysfunction in a given cell is diffi-
However, based on the inciting stimulus, apoptosis or inflam- cult to discern; however, multiple pathways could potentially
matory signaling may be the primary event, each stimulating the cooperate or oppose each other, to eventually result in mito-
other. The liver has a large population of Kupffer cells, NK chondrial permeabilization. Once mitochondrial permeabiliza-
cells, and NK T cells. These cells are a ready source of TNF‐α tion occurs, the hepatocyte is committed to cell death. Evidence
and other cytokines that mediate inflammation, such as Fas, of hepatocyte apoptosis can be demonstrated by serum markers
TRAIL, and TNF‐α that mediate hepatocyte apoptosis and and early studies demonstrate prognostic significance of apop-
transforming growth factor beta (TGF‐β) that activates stellate tosis markers. Lastly, therapeutic manipulation of apoptosis is
cells. Apoptotic hepatocytes can be engulfed by Kupffer cells, of benefit in preventing liver injury and fibrosis.
leading to generation of cytokines [6]. In keeping with this the
pharmacologic inhibition of apoptosis prevents Kupffer cell
activation. Also, in the bile duct‐ligated mouse, Kupffer cell
depletion decreases hepatocyte apoptosis, liver injury, and liver ACKNOWLEDGMENTS
inflammation. In addition, stressed hepatocytes increase expression
of NKG2D ligands, thus inviting NK‐ and NKT cell‐mediated Support for this work is provided by NIH grant DK 41876
destruction. (GJG), DK 111378 (HM), and the Mayo Foundation.
Fibrosis is the hallmark of ongoing liver injury. Hepatic stel-
late cells mediate hepatic fibrosis. In the normal liver, stellate
cells maintain a quiescent phenotype. On activation, they
undergo a metamorphosis to become myofibroblasts, secreting REFERENCES
collagen which leads to liver fibrosis. Stellate cells in vitro can
engulf apoptotic hepatocytes, leading to their activation, and 1. Malhi, H., Gores, G.J., and Lemasters, J.J. Apoptosis and necrosis in the
increased expression of TGF‐β, alpha smooth muscle actin, and liver: a tale of two deaths? Hepatology, 2006;43(2 Suppl 1):S31–44.
2. Green, D.R. and Kroemer, G. The pathophysiology of mitochondrial cell
collagen alpha 1 [98]. Similarly, in vivo hepatocyte apoptosis is
death. Science, 2004;305(5684):626–9.
a fibrogenic stimulus. Several experimental studies have dem- 3. Xiang, J., Chao, D.T., and Korsmeyer, S.J. BAX‐induced cell death may not
onstrated that the inhibition of hepatocyte apoptosis abrogates require interleukin 1 beta‐converting enzyme‐like proteases. Proc Natl Acad
liver fibrosis [5, 71, 99]. By corollary, apoptosis of activated Sci U S A, 1996;93(25):14559–63.
stellate cells should decrease liver fibrosis and dissociate ongo- 4. Rutherford, A.E., Hynan, L.S., Borges, C.B. et al. Serum apoptosis markers
in acute liver failure: a pilot study. Clin Gastroenterol Hepatol, 2007;
ing hepatocyte apoptosis from the ensuing fibrogenic response. 5(12):1477–83.
Indeed, activated stellate cells are sensitized to apoptotic signaling. 5. Canbay, A., Higuchi, H., Bronk, S.F. et al. Fas enhances fibrogenesis in the
This can be achieved by inhibition of NFκB, TRAIL‐mediated bile duct ligated mouse: a link between apoptosis and fibrosis. Gastroenterology,
stellate cell apoptosis, and NK cell‐mediated stellate cell apop- 2002;123(4):1323–30.
tosis. Indeed, the resolution phase of fibrosis requires apoptosis 6. Canbay, A., Friedman, S., and Gores, G.J. Apoptosis: the nexus of liver
injury and fibrosis. Hepatology, 2004;39(2):273–8.
of activated hepatic stellate cells. 7. Feldstein, A.E., Canbay, A., Angulo, P. et al. Hepatocyte apoptosis and fas
Lastly, the clinical applications of apoptosis are discussed in expression are prominent features of human nonalcoholic steatohepatitis.
the conclusion of this chapter. The cytokeratin 18‐derived M30 Gastroenterology, 2003;125(2):437–43.
neoantigen reflects epithelial cell apoptosis, is abundant in 8. Natori, S., Rust, C., Stadheim, L.M. et al. Hepatocyte apoptosis is a patho-
logic feature of human alcoholic hepatitis. J Hepatol, 2001;34(2):248–53.
hepatocytes, can easily be measured in serum by a commercially
9. Kohli, V., Selzner, M., Madden, J.F., Bentley, R.C., and Clavien, P.A.
available ELISA, and correlates with hepatocyte apoptosis in Endothelial cell and hepatocyte deaths occur by apoptosis after ischemia‐
diverse liver diseases. In a study with a small number of patients reperfusion injury in the rat liver. Transplantation, 1999;67(8):1099–105.
with chronic hepatitis C, pretreatment M30 levels were predic- 10. Guicciardi, M.E. and Gores, G.J. Life and death by death receptors. FASEB
tive of response to therapy, inferring from this that patients with J. 2009;23(6):1625–37.
11. Muzio, M., Stockwell, B.R., Stennicke, H.R., Salvesen, G.S., and Dixit,
an apoptotic response to virally infected hepatocytes are more
V.M. An induced proximity model for caspase‐8 activation. J Biol Chem.
likely to have a treatment response. In another study with chronic 1998;273(5):2926–30.
hepatitis C patients with normal transaminases, serum M30 lev- 12. Scaffidi, C., Fulda, S., Srinivasan, A. et al. Two CD95 (APO‐1/Fas) signaling
els correlated with fibrosis. In patients with non‐alcoholic fatty pathways. EMBO J, 1998;17(6):1675–87.
liver disease, serum M30 levels offer reliable discrimination of 13. Yin, X.M. Bid, a critical mediator for apoptosis induced by the activation of
Fas/TNF‐R1 death receptors in hepatocytes. J Mol Med (Berl), 2000
patients with steatohepatitis from simple steatosis, and increas- ;78(4):203–11.
ing levels are predictive of a higher likelihood of inflammation. 14. Wang, X., DeFrances, M.C., Dai, Y. et  al. A mechanism of cell survival:
Caspase inhibitors have demonstrated efficacy in preventing sequestration of Fas by the HGF receptor Met. Mol Cell, 2002;9(2):411–21.
hepatocyte apoptosis and injury in experimental models of liver 15. Danial, N.N. and Korsmeyer, S.J. Cell death: critical control points. Cell,
injury [99, 100]. In patients with chronic hepatitis C, orally 2004;116(2):205–19.
16. Yamada, Y., Webber, E.M., Kirillova, I., Peschon, J.J., and Fausto, N.
administered caspase inhibitor was found to be safe, and lowered Analysis of liver regeneration in mice lacking type 1 or type 2 tumor necrosis
transaminases. The caspase inhibitor Emricasan (IDUN‐6556) is factor receptor: requirement for type 1 but not type 2 receptor. Hepatology,
currently in clinical trials for non‐alcoholic steatohepatitis. 1998;28(4):959–70.
204 THE LIVER:  REFERENCES

17. Tartaglia, L.A., Ayres, T.M., Wong, G.H., and Goeddel, D.V. A novel domain 42. Toyoda, M., Kakizaki, S., Horiguchi, N. et al. Role of serum soluble Fas/
within the 55 kd TNF receptor signals cell death. Cell, 1993;74(5):845–53. soluble Fas ligand and TNF‐alpha on response to interferon‐alpha therapy in
18. Micheau, O. and Tschopp, J. Induction of TNF receptor I‐mediated apopto- chronic hepatitis C. Liver, 2000;20(4):305–11.
sis via two sequential signaling complexes. Cell, 2003;114(2):181–90. 43. Song, le H., Binh, V.Q., Duy, D.N. et al. Variations in the serum concentra-
19. Grootjans, S., Vanden Berghe, T., and Vandenabeele, P. Initiation and execu- tions of soluble Fas and soluble Fas ligand in Vietnamese patients infected
tion mechanisms of necroptosis: an overview. Cell Death Differ, with hepatitis B virus. J Med Virol, 2004;73(2):244–9.
2017;24(7):1184–95. 44. Miyoshi, H., Rust, C., Roberts, P.J., Burgart, L.J., and Gores, G.J. Hepatocyte
20. Cai, Z., Jitkaew, S., Zhao, J. et al. Plasma membrane translocation of trimer- apoptosis after bile duct ligation in the mouse involves Fas. Gastroenterology,
ized MLKL protein is required for TNF‐induced necroptosis. Nat Cell Biol, 1999;117(3):669–77.
2014;16(1):55–65. 45. Sodeman, T., Bronk, S.F., Roberts, P.J., Miyoshi, H., and Gores, G.J. Bile
21. Sass, G., Shembade, N.D., Haimerl, F. et  al. TNF pretreatment interferes salts mediate hepatocyte apoptosis by increasing cell surface trafficking of
with mitochondrial apoptosis in the mouse liver by A20‐mediated down‐ Fas. Am J Physiol Gastrointest Liver Physiol, 2000;278(6):G992–9.
regulation of Bax. J Immunol, 2007;179(10):7042–9. 46. Faubion, W.A., Guicciardi, M.E., Miyoshi, H. et al. Toxic bile salts induce
22. Guicciardi, M.E., Deussing, J., Miyoshi, H. et al. Cathepsin B contributes to rodent hepatocyte apoptosis via direct activation of Fas. J Clin Invest,
TNF‐alpha‐mediated hepatocyte apoptosis by promoting mitochondrial 1999;103(1):137–45.
release of cytochrome c. J Clin Invest, 2000;106(9):1127–37. 47. Reinehr, R., Becker, S., Wettstein, M., and Haussinger, D. Involvement of the
23. Guicciardi, M.E., Miyoshi, H., Bronk, S.F., and Gores, G.J. Cathepsin B Src family kinase yes in bile salt‐induced apoptosis. Gastroenterology,
knockout mice are resistant to tumor necrosis factor‐alpha‐mediated hepato- 2004;127(5):1540–57.
cyte apoptosis and liver injury: implications for therapeutic applications. Am 48. Kimberley, F.C. and Screaton, G.R. Following a TRAIL: update on a ligand
J Pathol, 2001;159(6):2045–54. and its five receptors. Cell Res, 2004;14(5):359–72.
24. Varfolomeev, E., Blankenship, J.W., Wayson, S.M. et  al. IAP antagonists 49. Schneider, P., Thome, M., Burns, K. et al. TRAIL receptors 1 (DR4) and 2
induce autoubiquitination of c‐IAPs, NF‐kappaB activation, and TNFalpha‐ (DR5) signal FADD‐dependent apoptosis and activate NF‐kappaB.
dependent apoptosis. Cell, 2007;131(4):669–81. Immunity, 1997;7(6):831–6.
25. Kim, Y.S., Morgan, M.J., Choksi, S., and Liu, Z.G. TNF‐induced activation 50. Volkmann, X., Fischer, U., Bahr, M.J. et  al. Increased hepatotoxicity of
of the Nox1 NADPH oxidase and its role in the induction of necrotic cell tumor necrosis factor‐related apoptosis‐inducing ligand in diseased human
death. Mol Cell, 2007;26(5):675–87. liver. Hepatology, 2007;46(5):1498–508.
26. Mohammed, F.F., Smookler, D.S., Taylor, S.E. et al. Abnormal TNF activity 51. Malhi, H., Barreyro, F.J., Isomoto, H., Bronk, S.F., and Gores, G.J. Free fatty
in Timp3‐/‐ mice leads to chronic hepatic inflammation and failure of liver acids sensitise hepatocytes to TRAIL mediated cytotoxicity. Gut,
regeneration. Nat Genet, 2004;36(9):969–77. 2007;56(8):1124–31.
27. Yin, M., Wheeler, M.D., Kono, H. et al. Essential role of tumor necrosis fac- 52. Higuchi, H., Yoon, J.H., Grambihler, A. et  al. Bile acids stimulate cFLIP
tor alpha in alcohol‐induced liver injury in mice. Gastroenterology, phosphorylation enhancing TRAIL‐mediated apoptosis. J Biol Chem,
1999;117(4):942–52. 2003;278(1):454–61.
28. Rudiger, H.A. and Clavien, P.A. Tumor necrosis factor alpha, but not Fas, 53. Higuchi, H., Bronk, S.F., Taniai, M., Canbay, A., and Gores, G.J. Cholestasis
mediates hepatocellular apoptosis in the murine ischemic liver. increases tumor necrosis factor‐related apoptotis‐inducing ligand (TRAIL)‐
Gastroenterology, 2002;122(1):202–10. R2/DR5 expression and sensitizes the liver to TRAIL‐mediated cytotoxicity.
29. Ribeiro, P.S., Cortez‐Pinto, H., Sola, S. et al. Hepatocyte apoptosis, expres- J Pharmacol Exp Ther, 2002;303(2):461–7.
sion of death receptors, and activation of NF‐kappaB in the liver of nonalco- 54. Idrissova, L., Malhi, H., Werneburg, N.W. et al. TRAIL receptor deletion in
holic and alcoholic steatohepatitis patients. Am J Gastroenterol, mice suppresses the inflammation of nutrient excess. J Hepatol,
2004;99(9):1708–17. 2015;62(5):1156–63.
30. Spahr, L., Giostra, E., Frossard, J.L. et al. Soluble TNF‐R1, but not tumor 55. Hirsova, P., Weng, P., Salim, W. et al. TRAIL deletion prevents liver inflam-
necrosis factor alpha, predicts the 3‐month mortality in patients with alco- mation but not adipose tissue inflammation during murine diet‐induced obe-
holic hepatitis. J Hepatol, 2004;41(2):229–34. sity. Hepatol Commun, 2017;1(7):648–62.
31. Torre, F., Rossol, S., Pelli, N. et al. Kinetics of soluble tumour necrosis factor 56. Mundt, B., Kuhnel, F., Zender, L. et al. Involvement of TRAIL and its recep-
(TNF)‐alpha receptors and cytokines in the early phase of treatment for chronic tors in viral hepatitis. FASEB J, 2003;17(1):94–6.
hepatitis C: comparison between interferon (IFN)‐alpha alone, IFN‐alpha plus 57. Chou, A.H., Tsai, H.F., Wu, Y.Y. et  al. Hepatitis C virus core protein
amantadine or plus ribavirin. Clin Exp Immunol, 2004; 136(3):507–12. modulates TRAIL‐mediated apoptosis by enhancing Bid cleavage and acti-
32. Fang, J.W., Shen, W.W., Meager, A., and Lau, J.Y. Activation of the tumor vation of mitochondria apoptosis signaling pathway. J Immunol, 2005;
necrosis factor‐alpha system in the liver in chronic hepatitis B virus infec- 174(4):2160–6.
tion. Am J Gastroenterol, 1996;91(4):748–53. 58. Takehara, T., Tatsumi, T., Suzuki, T. et al. Hepatocyte‐specific disruption of
33. Ogasawara, J., Watanabe‐Fukunaga, R., Adachi, M. et al. Lethal effect of the Bcl‐xL leads to continuous hepatocyte apoptosis and liver fibrotic responses.
anti‐Fas antibody in mice. Nature, 1993;364(6440):806–9. Gastroenterology, 2004;127(4):1189–97.
34. Desbarats, J. and Newell, M.K. Fas engagement accelerates liver regenera- 59. Weber, A., Boger, R., Vick, B. et  al. Hepatocyte‐specific deletion of the
tion after partial hepatectomy. Nat Med, 2000;6(8):920–3. antiapoptotic protein myeloid cell leukemia‐1 triggers proliferation and
35. Berke, G. The CTl’s kiss of death. Cell, 1995;81(1):9–12. hepatocarcinogenesis in mice. Hepatology, 2010;51(4):1226–36.
36. Adachi, M., Suematsu, S., Kondo, T. et al. Targeted mutation in the Fas gene 60. Barreyro, F.J., Kobayashi, S., Bronk, S.F. et al. Transcriptional regulation of
causes hyperplasia in peripheral lymphoid organs and liver. Nat Genet, Bim by Fox03A mediates hepatocyte lipoapoptosis. J Biol Chem,
1995;11(3):294–300. 2007;282(37):27141–54.
37. Lacronique, V., Mignon, A., Fabre, M. et  al. Bcl‐2 protects from lethal 61. Yu, J. and Zhang, L. The transcriptional targets of p53 in apoptosis control.
hepatic apoptosis induced by an anti‐Fas antibody in mice. Nat Med, Biochem Biophys Res Commun, 2005;331(3):851–8.
1996;2(1):80–6. 62. Kalkavan, H. and Green, D.R. MOMP, cell suicide as a BCL‐2 family busi-
38. Zhang, H., Cook, J., Nickel, J. et al. Reduction of liver Fas expression by an ness. Cell Death Differ, 2018;25(1):46–55.
antisense oligonucleotide protects mice from fulminant hepatitis. Nat 63. Giorgio, V., von Stockum, S., Antoniel, M. et al. Dimers of mitochondrial
Biotechnol, 2000;18(8):862–7. ATP synthase form the permeability transition pore. Proc Natl Acad Sci U S
39. Al‐Saeedi, M., Steinebrunner, N., Kudsi, H. et al. Neutralization of CD95 A, 2013;110(15):5887–92.
ligand protects the liver against ischemia‐reperfusion injury and prevents 64. Riedl, S.J. and Salvesen, G.S. The apoptosome: signalling platform of cell
acute liver failure. Cell Death Dis, 2018;9(2):132. death. Nat Rev Mol Cell Biol, 2007;8(5):405–13.
40. Ryo, K., Kamogawa, Y., Ikeda, I. et al. Significance of Fas antigen‐mediated 65. Guicciardi, M.E., Leist, M., and Gores, G.J. Lysosomes in cell death.
apoptosis in human fulminant hepatic failure. Am J Gastroenterol, Oncogene, 2004;23(16):2881–90.
2000;95(8):2047–55. 66. Guicciardi, M.E., Bronk, S.F., Werneburg, N.W., Yin, X.M., and Gores, G.J.
41. Kiyici, M., Gurel, S., Budak, F. et  al. Fas antigen (CD95) expression and Bid is upstream of lysosome‐mediated caspase 2 activation in tumor necro-
apoptosis in hepatocytes of patients with chronic viral hepatitis. Eur J sis factor alpha‐induced hepatocyte apoptosis. Gastroenterology, 2005;
Gastroenterol Hepatol, 2003;15(10):1079–84. 129(1):269–84.
17:  Hepatocyte Apoptosis: Mechanisms and Relevance in Liver Diseases 205

67. Stoka, V., Turk, B., Schendel, S.L. et  al. Lysosomal protease pathways to 84. Puri, P., Mirshahi, F., Cheung, O. et al. Activation and dysregulation of the
apoptosis. Cleavage of bid, not pro‐caspases, is the most likely route. J Biol unfolded protein response in nonalcoholic fatty liver disease.
Chem, 2001;276(5):3149–57. Gastroenterology, 2008;134(2):568–76.
68. Feldstein, A.E., Werneburg, N.W., Li, Z., Bronk, S.F., and Gores, G.J. 85. Wei, Y., Wang, D., Topczewski, F., and Pagliassotti, M.J. Saturated fatty
Bax inhibition protects against free fatty acid‐induced lysosomal perme- acids induce endoplasmic reticulum stress and apoptosis independently of
abilization. Am J Physiol Gastrointest Liver Physiol, 2006;290(6): ceramide in liver cells. Am J Physiol Endocrinol Metab, 2006;
G1339–46. 291(2):E275–81.
69. Gove, C.D., Wardle, E.N., and Williams, R. Circulating lysosomal enzymes 86. Ji, C., Mehrian‐Shai, R., Chan, C., Hsu, Y.H., and Kaplowitz, N. Role of
and acute hepatic necrosis. J Clin Pathol, 1981;34(1):13–16. CHOP in hepatic apoptosis in the murine model of intragastric ethanol
70. Kyaw, A., Aung, T., Htut, T., Myint, H., and Tin, K.M. Lysosomal enzyme feeding. Alcohol Clin Exp Res, 2005;29(8):1496–503.
activities in normals and in patients with chronic liver diseases. Clin Chim 87. Czaja, M.J. The future of GI and liver research: editorial perspectives. III.
Acta, 1983;131(3):317–23. JNK/AP‐1 regulation of hepatocyte death. Am J Physiol Gastrointest Liver
71. Canbay, A., Guicciardi, M.E., Higuchi, H. et  al. Cathepsin B inactivation Physiol, 2003;284(6):G875–9.
attenuates hepatic injury and fibrosis during cholestasis. J Clin Invest, 88. Urano, F., Wang, X., Bertolotti, A. et al. Coupling of stress in the ER to
2003;112(2):152–9. activation of JNK protein kinases by transmembrane protein kinase IRE1.
72. Baskin‐Bey, E.S., Canbay, A., Bronk, S.F. et  al. Cathepsin B inactivation Science, 2000;287(5453):664–6.
attenuates hepatocyte apoptosis and liver damage in steatotic livers after cold 89. Ueda, S., Masutani, H., Nakamura, H. et al. Redox control of cell death.
ischemia‐warm reperfusion injury. Am J Physiol Gastrointest Liver Physiol, Antioxid Redox Signal, 2002;4(3):405–14.
2005;288(2):G396–402. 90. Tournier, C., Hess, P., Yang, D.D. et  al. Requirement of JNK for stress‐
73. Ron, D. and Walter, P. Signal integration in the endoplasmic reticulum induced activation of the cytochrome c‐mediated death pathway. Science,
unfolded protein response. Nat Rev Mol Cell Biol, 2007;8(7):519–29. 2000;288(5467):870–4.
74. Malhi, H. and Kaufman, R.J. Endoplasmic reticulum stress in liver disease. 91. Ventura, J.J., Hubner, A., Zhang, C. et al. Chemical genetic analysis of the
J Hepatol, 2011;54(4):795–809. time course of signal transduction by JNK. Mol Cell, 2006;21(5):701–10.
75. Lin, J.H., Li, H., Yasumura, D. et al. IRE1 signaling affects cell fate during 92. Malhi, H., Bronk, S.F., Werneburg, N.W., and Gores, G.J. Free fatty acids
the unfolded protein response. Science, 2007;318(5852):944–9. induce JNK‐dependent hepatocyte lipoapoptosis. J Biol Chem,
76. Hetz, C., Bernasconi, P., Fisher, J. et al. Proapoptotic BAX and BAK modu- 2006;281(17):12093–101.
late the unfolded protein response by a direct interaction with IRE1alpha. 93. Lei, K. and Davis, R.J. JNK phosphorylation of Bim‐related members of
Science, 2006;312(5773):572–6. the Bc12 family induces Bax‐dependent apoptosis. Proc Natl Acad Sci U S
77. Scorrano, L., Oakes, S.A., Opferman, J.T. et al. BAX and BAK regulation of A, 2003;100(5):2432–7.
endoplasmic reticulum Ca2+: a control point for apoptosis. Science, 94. Chang, L., Kamata, H., Solinas, G. et al. The E3 ubiquitin ligase itch cou-
2003;300(5616):135–9. ples JNK activation to TNFalpha‐induced cell death by inducing c‐FLIP(L)
78. Puthalakath, H., O’Reilly, L.A., Gunn, P. et al. ER stress triggers apoptosis turnover. Cell, 2006;124(3):601–13.
by activating BH3‐only protein Bim. Cell, 2007;129(7):1337–49. 95. Hirosumi, J., Tuncman, G., Chang, L. et al. A central role for JNK in obe-
79. He, Q., Luo, X., Jin, W. et  al. Celecoxib and a novel COX‐2 inhibitor sity and insulin resistance. Nature, 2002;420(6913):333–6.
ON09310 upregulate death receptor 5 expression via GADD153/CHOP. 96. Higuchi, H., Bronk, S.F., Takikawa, Y. et al. The bile acid glycochenode-
Oncogene, 2008;27(18):2656–60. oxycholate induces trail‐receptor 2/DR5 expression and apoptosis. J Biol
80. Tamaki, N., Hatano, E., Taura, K. et al. CHOP deficiency attenuates choles- Chem, 2001;276(42):38610–18.
tasis‐induced liver fibrosis by reduction of hepatocyte injury. Am J Physiol 97. Uehara, T., Xi Peng, X., Bennett, B. et al. c‐Jun N‐terminal kinase mediates
Gastrointest Liver Physiol, 2008;294(2):G498–505. hepatic injury after rat liver transplantation. Transplantation, 2004;
81. Tsuchiya, S., Tsuji, M., Morio, Y., and Oguchi, K. Involvement of endoplas- 78(3):324–32.
mic reticulum in glycochenodeoxycholic acid‐induced apoptosis in rat 98. Canbay, A., Taimr, P., Torok, N. et  al. Apoptotic body engulfment by a
hepatocytes. Toxicol Lett, 2006;166(2):140–9. human stellate cell line is profibrogenic. Lab Invest, 2003;83(5):655–63.
82. Tumurbaatar, B., Sun, Y., Chan, T., and Sun, J. Cre‐estrogen receptor‐medi- 99. Canbay, A., Feldstein, A., Baskin‐Bey, E., Bronk, S.F., and Gores, G.J. The
ated hepatitis C virus structural protein expression in mice. J Virol Methods, caspase inhibitor IDN‐6556 attenuates hepatic injury and fibrosis in the bile
2007;146(1–2):5–13. duct ligated mouse. J Pharmacol Exp Ther, 2004;308(3):1191–6.
83. Ito, K., Kiyosawa, N., Kumagai, K. et al. Molecular mechanism investiga- 100. Natori, S., Higuchi, H., Contreras, P., and Gores, G.J. The caspase inhibitor
tion of cycloheximide‐induced hepatocyte apoptosis in rat livers by morpho- IDN‐6556 prevents caspase activation and apoptosis in sinusoidal endothelial
logical and microarray analysis. Toxicology, 2006;219(1–3):175–86. cells during liver preservation injury. Liver Transpl, 2003;9(3):278–84.
SECTION B:
THE HEPATOCYTE
Copper Metabolism
18 and the Liver
Cynthia Abou Zeid, Ling Yi, and Stephen G. Kaler
Section on Translational Neuroscience, Molecular Medicine Branch, Intramural Research Program, National
Institutes of Health, Bethesda, MD, USA

OVERVIEW THE LIVER: A CENTRAL ORGAN


OF COPPER METABOLISM
Among its several roles in the human body, the liver is a piv-
otal organ involved in metabolizing and excreting various Dietary copper absorption primarily occurs in the duodenum.
toxic substances into the bile. One such function is elimination The process begins with uptake of reduced copper via the apical
of trace metals, such as copper, that are necessary for several membrane of the enterocyte, through the copper transporter
physiological processes but are also toxic if present in excess. Ctr1 [10, 11]. DMT1 is a less‐specific divalent metal transporter
Copper is a micronutrient acquired from the diet and plays a that also imports iron, manganese, and nickel, and can play an
critical role in human growth and development. It is a divalent additional role in copper uptake [12]. Once inside the entero-
metal ion that can be found in both the cuprous (Cu+) and cyte, copper is shuttled by its chaperone ATOX1 to the copper‐
cupric (Cu2+) states. Its capacity to readily gain and donate transporting ATPase ATP7A. The latter is responsible for copper
electrons renders it important in various metabolic pathways. export via the enterocyte’s basolateral pole into the portal circu-
Copper is used as a cofactor for enzymes involved in the mito- lation (Figure  18.1). ATP7A has ubiquitous distribution and
chondrial electron transport chain, detoxification of reactive important roles in several cell types [13]. In hepatocytes, ATP7A
oxygen species, iron transport, connective tissue metabolism, has higher expression in the neonatal period with a reduction
melanin pigment production, synthesis of amidated neuropep- over time [14]. In the liver, the predominant copper‐transporting
tides, and catecholamine metabolism [1–5]. However, the ATPase is ATP7B, a protein closely related to ATP7A with simi-
chemical properties of copper also expose to the risk of oxida- larities in its structure and function [15]. The general role of
tive damage when homeostatic mechanisms are impaired. these pumps depends on copper levels in the cell. In physiologi-
Consequently, tissue copper levels are tightly regulated by sev- cal copper states, ATP7A and ATP7B localize to the trans‐Golgi
eral transporters and chaperone proteins involved in systemic network and are responsible for conveying copper into the
copper metabolism [3, 6]. The main route of systemic copper secretory pathway, for metalation of copper‐dependent enzymes.
regulation is considered to be biliary excretion. Emerging data In case of elevated intracellular levels of copper, these transport-
support regulation at the intestinal absorption level, but these ers traffic to the plasma membrane to export copper to the extra-
mechanisms are not completely elucidated [7–9]. In all cases, cellular milieu. The latter mechanism is how ATP7A works in
the liver is still counted as the major organ of copper metabo- intestinal absorption of copper, by sending the metal from the
lism, since it functions not only in the elimination of excessive enterocyte into the portal venous circulation [16, 17].
copper, but also in its distribution to several destinations fol- The copper pool in blood is then bound to serum proteins such
lowing intestinal uptake. as albumin and α2‐macroglobulin [18] and directed to the liver.

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
210 THE LIVER:  COPPER, THE LIVER, AND DISEASE

Figure 18.1  Copper metabolism in enterocytes and hepatocytes. (a) In enterocytes, copper (Cu+) uptake is mediated by CTR1, possibly in concert
with the metalloreductases STEAP1, STEAP2, and dCYTB. The roles of CTR2 (not shown) and DMT1 in this process are less certain. Within
enterocytes, GSH and MT function in copper sequestration and storage. The chaperones CCS, ATOX1, Cox17, Cox11, and Sco1 ferry copper to
specific proteins or organelles. With an increase in copper levels, ATP7A traffics to the basolateral surface and pumps copper into the blood. (b) In
hepatocytes, ATOX1 provides copper to ATP7B for metalation of ceruloplasmin (CP), and traffics to the apical membrane to pump copper into the
bile, the body’s major mechanism for copper removal. CCS has been proposed to deliver copper to XIAP, which may interact with COMMD1, a
protein mutated in hepatic copper toxicosis of Bedlington terriers and which may modulate ATP7B activity. This putative pathway is denoted by
dashed lines. Modified from [3] with permission of Springer Nature.

Copper is imported through the basolateral membrane of hepato- hepato‐lenticular degeneration [25]. It was not until three
cytes by Ctr1 and DMT1 (Figure 18.1) [3]. For release into the decades later that elevated copper levels and accumulation in
bloodstream, copper needs to be bound to proteins, the main spe- the brain and liver were identified, which led to the classifica-
cies in blood being ceruloplasmin, a secreted glycoprotein that tion of this disease as a disorder of copper metabolism [26].
acts as a ferroxidase and functions in systemic iron metabolism This biochemical phenotyping facilitated discovery of the
[19]. It is synthetized in the liver, and is initially devoid of copper putative role of ATP7B in Wilson disease, supported by the
(apoceruloplasmin) [20]. In the trans‐Golgi network of hepato- similarities with ATP7A [27].
cytes, copper is loaded onto apoceruloplasmin by ATP7B and Wilson disease is an autosomal recessive disorder caused by
subsequently secreted into the blood circulation. When liver cop- a dysfunction of the copper transporter ATP7B, and more than
per levels increase, ATP7B trafficks to the apical membrane of 600 disease‐causing mutations have been identified to date [27,
hepatocytes, from where it excretes excess copper into the biliary 28]. The distribution and frequency of specific mutations vary
canaliculi. Through biliary excretion, the liver is the major organ depending on the geographic region and population studied
responsible for copper level regulation by elimination of the [29–32]. The types of amino acid changes most frequently
metal. Renal excretion of copper also occurs, and becomes clini- found are missense mutations, small insertions or deletions, and
cally significant only in the presence of pathological mecha- splice site junction mutations [33]. Pathological variations in
nisms of copper overload from metabolic disorders [3], or in the sequence of this copper transporter cause copper overload
conditions resulting in biliary obstruction [21–24]. and injury in the hepatocytes. Excessive copper eventually spills
The liver is therefore the central organ in copper metabolism, into the blood circulation and accumulates in several extrahe-
which explains how dysfunction in one can impact the other. patic tissues, namely the brain, kidneys, and cornea, which
Disorders of copper metabolism manifest in the liver with clini- helps explain the signs and symptoms of this disease [34].
cal, laboratory, and pathological signs of hepatic dysfunction, as Wilson disease classically presents in individuals ranging in
discussed below. age from 3 to 50 years, with a variable combination of neuro-
logical, psychiatric, and hepatic manifestations [35, 36]. Clinical
signs and symptoms vary greatly among affected individuals,
even within a single family [30]. Liver disease classically occurs
COPPER, THE LIVER, AND DISEASE
in Wilson disease individuals who are symptomatic and diag-
nosed at an early age from childhood through early adulthood.
Copper metabolism disorders involving Neurological and psychiatric manifestations are more frequent
the liver in older individuals, where hepatic involvement may be less
pronounced [37]. In addition, very late onset (after age 70)
Wilson disease
and very early onset (nine months of age) of Wilson disease
Wilson disease was first described in 1912 by Samuel Alexander have been reported [38, 39]. Hepatic presentations can be both
Kinnier Wilson as a familial disorder associating liver cir- acute or chronic. Acute Wilsonian liver disease manifests as
rhosis with neurological manifestations, also referred to as a rapidly occurring jaundice from a hepatitis‐like injury or a
18:  Copper Metabolism and the Liver 211

Coombs‐negative hemolytic disease and progresses to fulmi- expression of metallothionein, an intracellular metal chelator
nant liver failure. The illness may also present as chronic liver that binds copper with an even higher affinity than zinc, and
disease, with or without cirrhosis. Especially since there are impairs its absorption. Pharmacotherapy using copper chelators
treatments available, Wilson disease should be considered in the is ideally continued as a lifelong treatment. However, side
differential diagnosis of any unexplained chronic liver disease effects such as dermatological changes, hypersensitivity
or new‐onset neuropsychiatric symptoms. ­reactions, autoimmune disease, bone marrow suppression, and
Psychiatric manifestations of Wilson disease include mood nephrotoxicity can be an obstacle for treatment adherence.
and personality disorders and sometimes cognitive deteriora- Resistant cases that do not respond to pharmacological therapy
tion. Neurological presentations are dominated by extrapyrami- and life‐threatening cases of Wilson disease may be treated with
dal symptoms including tremor, chorea, and choreoathetosis, liver transplantation. A liver transplant that replaces the major
problems in coordination and fine motor control, as well as dysfunctional organ is usually curative [41, 42].
rigidity and gait disturbance. From the ophthalmological stand- A number of new therapies for Wilson disease are currently
point, copper accumulation in the cornea may form a brown in preclinical stage studies. Targets are inhibition of ATP7B
pathognomonic circle called the Kayser–Fleischer ring visible mutant retention and degradation in the endoplasmic reticulum,
on slit lamp examination [30] (Figure  18.2). If clearly docu- promotion of copper excretion [43, 44], and introduction of
mented, this finding can confirm the diagnosis in a patient sus- engineered hepatic cells with the hope of repopulating the liver
pected of having Wilson disease. [45]. Viral gene therapy for Wilson disease aims to introduce
Confirmatory laboratory tests in Wilson disease include low working copies of ATP7B by delivering the normal ATP7B
levels of serum copper and ceruloplasmin that result from the complementary DNA (cDNA) to hepatocytes. This is accom-
inability of mutated ATP7B to load copper onto apoceruloplas- plished by incorporating the normal gene’s cDNA in a nonpath-
min, and increased 24‐hour urinary copper excretion. If liver ogenic virus that targets the liver [46–48]. Gene therapy showed
biopsy is obtained during diagnostic workup of unexplained promise in a mouse model with an advanced stage of Wilson
liver disease, hepatic copper levels are usually markedly ele- disease; liver enzymes and copper excretion were normalized
vated. A scoring system for the diagnosis of Wilson disease [49]. This approach has potential for future application in clini-
based on both clinical and laboratory criteria is available to cal trials.
guide clinicians faced with a possible case of this disorder [40].
Definitive confirmation is obtained by molecular analysis of the Huppke–Brendel syndrome
ATP7B gene.
Huppke–Brendel syndrome is a recently described autosomal
Treatment of Wilson disease should begin as soon as the
recessive copper metabolism disorder with a biochemical
diagnosis is made to prevent evolution to fulminant liver dis-
phenotype similar to that of Wilson disease, but with different
ease, cirrhosis, or irreversible neuropsychiatric damage. Therapy
etiology and clinical manifestations. Patients with Huppke–Brendel
in symptomatic patients is based on the administration of copper
syndrome are infants and children with low serum levels of cop-
chelators such as d‐penicillamine or trientine to prevent its
per and ceruloplasmin associated with syndromic features
accumulation in organs. Zinc salts can also be used as means to
including congenital cataracts, hearing loss, and severe develop-
inhibit copper absorption in the intestines. Oral zinc induces
mental delay [50]. On brain imaging, cerebellar atrophy, wid-
ened subarachnoid spaces, and hypomyelination are evident.
This disorder is lethal in early childhood, with death reported
between 22 months and 6 years of age in patients from the origi-
nal cohort. Initially, several known copper transporters and pro-
teins were investigated for their potential role in the pathogenesis
of this syndrome, with no success [51]. Homozygosity mapping
and deep sequencing later identified the genetic basis of the dis-
order, mutations in SLC33A1, a gene that encodes the acetyl‐
CoA transporter AT‐1 [52]. It is thought that AT‐1 is needed to
acetylate glycoproteins and gangliosides in the endoplasmic
reticulum and/or Golgi apparatus [53]. Copper metabolism
imbalance in Huppke–Brendel patients could be secondary to
the effects of a defective AT‐1 on trafficking of copper pumps
ATP7A and ATP7B, since both proteins normally undergo acet-
ylation of specific lysine residues (L. Yi and S.G. Kaler, unpub-
lished data).

Figure 18.2  Kayser–Fleischer ring in the cornea of an adult patient MEDNIK syndrome
with Wilson disease, resulting from copper deposition in Descemet’s MEDNIK is an acronym used to describe an autosomal recessive
membrane. Image courtesy of T.U. Hoogenraad, M.D., Department of
Neurology, University Hospital, Utrecht, The Netherlands. Reproduced
neurocutaneous syndrome that was first defined in French‐
from Kaler, Wilson disease, in Cecil’s Textbook of Medicine, 23rd edn. Canadian families sharing common ancestors [54]. MEDNIK
(eds. L. Goldman and D. Ausiello), Saunders, Philadelphia, 2008, ch. stands for mental retardation, enteropathy, deafness, neuropathy,
230, pp. 1593–5 with permission of Elsevier. ichthyosis, and keratodermia. In addition to these clinical signs
212 THE LIVER:  CONCLUSIONS

and symptoms, dysfunction in copper metabolism was later significant fat deposits accompanied by hepatocellular inflam-
identified in one of the index cases reported with this disease, and mation and necrosis (non‐alcoholic steatohepatitis or NASH).
subsequently confirmed in the original cohort [55]. Patients with The latter can potentially progress to fibrosis and cirrhosis, end‐
MEDNIK present a mixed biochemical phenotype with elements stage liver disease, and hepatocellular carcinoma [58]. NAFLD
of both copper deficiency and copper excess. Developmental is closely related to the metabolic syndrome, which encom-
delays and brain atrophy found in MEDNIK subjects are reminis- passes abdominal obesity, hypertension, dyslipidemia, and
cent of Menkes disease, a neurodegenerative disorder secondary impaired glucose tolerance or diabetes [59]. The pathophysiol-
to mutations of ATP7A [3]. At the same time, MEDNIK patients ogy of this disorder is complex and incompletely understood,
show features of copper overload and liver disease with hypocer- with a possible role of copper in the process. Interestingly, a
uloplasminemia, accumulation of copper in the liver, and intrahe- study of 124 patients with NAFLD showed reduced hepatic cop-
patic cholestasis [55]. These changes are similar to those seen in per concentrations compared to controls and other liver diseases.
Wilson disease caused by ATP7B mutations. This correlation was more pronounced in patients with advanced
The molecular basis of MEDNIK syndrome is mutation in the hepatic steatosis, non‐alcoholic steatohepatitis, and additional
AP1S1 gene that encodes the sigma 1A (σ1A) subunit of adaptor features of the metabolic syndrome [60]. Copper deficiency has
protein complex 1 (AP‐1), a complex that normally mediates previously been linked to atherogenic dyslipidemia, and a recent
intracellular trafficking of certain transmembrane proteins study of ATP7B showed additional data connecting copper and
between the trans‐Golgi network, endosomes, and plasma mem- lipid metabolism at the enterocyte level [7, 61, 62]. Copper‐defi-
brane of cells. The ATP7A and ATP7B copper ATPases are both cient diets in rats induced hepatic steatosis and insulin resistance,
transmembrane proteins with similar structures, and their func- which further supports the role of low copper in the pathogenesis
tion depends on their ability to translocate to different cellular of NAFLD [60–62]. Additionally, beta‐oxidation occurs in per-
compartments in response to changing cellular copper levels. In oxisomes and mitochondria, and the process is downregulated in
studies of ATP7A trafficking, the AP‐1 complex was implicated states of copper deficiency [63]. Copper is needed for normal
as a key player in this intracellular itinerary by interaction with a mitochondrial function, since it serves as a critical cofactor for
di‐leucine motif near the C‐terminus [56, 57]. This conclusion cytochrome c oxidase, the last enzyme complex in the mitochon-
was extrapolated to ATP7B, given the similarities between these drial electron transport chain. Copper‐deficient mice show mito-
two ATPases, including the di‐leucine motif to which AP‐1 binds. chondrial dysfunction [64], and similar alterations can be seen in
With a defective σ1A subunit of the AP‐1 complex, ATP7A/B humans with NAFLD. Finally, copper is also a cofactor of super-
appear unable to remain docked in the trans‐Golgi network under oxide dismutase 1 (SOD1), an important enzyme for controlling
normal or low copper states [57]. This defect in copper ATPase oxidative stress, which has been linked to the hepatic damage in
localization and trafficking may represent the link between NAFLD and NASH [4, 60].
AP1S1 mutations and the copper metabolism disturbances found
in patients with MEDNIK. It seems conceivable that ATP7B Hepatocellular carcinoma
function in the liver could be impacted more than ATP7A in other
The role of copper in cancer development and progression has
tissues, since two other isoforms of the σ1 subunit, σ1B and σ1C,
been widely studied, with several reports showing high serum
exist and may substitute for σ1A, depending on their expression
and tumor copper levels in multiple cancer types [65–67]. High
levels (L. Yi and S.G. Kaler, unpublished data). The cutaneous
copper levels promote angiogenesis and oxidative stress genera-
and other clinical features of MEDNIK presumably relate to
tion, processes which create a suitable environment for tumor
impaired function of other transmembrane protein cargos.
growth, and which could be counteracted by lowering copper
Zinc acetate therapy has been suggested and used to reduce
levels [67–70]. These realizations led to the investigation of
copper overload in the liver of MEDNIK patients [55]. Further
copper chelation as a potential anticancer therapy [71, 72], with
understanding of copper ATPase trafficking in MEDNIK subjects
applications in hepatocellular carcinoma (HCC).
could shed light on the precise pathophysiology of the disorder,
HCC is one of the cancers in which studies have documented
and allow the development of more precisely targeted therapies.
high tumor copper levels [73–75]. Serum copper and cerulo-
plasmin levels were also found to be elevated in patients with
Primary disorders of the liver and copper HCC, and higher levels correlated with worse prognosis [76].
Interestingly, other types of hepatic neoplasms, including cholan-
imbalance giocarcinoma and metastatic liver disease, did not show significant
Since the liver is the main regulatory organ of copper metabolism, copper level elevation, which helps distinguish them from HCC
the consequences of an imbalance in the metal’s levels are [77, 78]. The therapeutic strategy of lowering copper levels to treat
expected in hepatocytes, as previously described in diseases such cancer tested in vivo with HCC yielded promising results showing
as Wilson disease and MEDNIK. Conversely, evidence of copper suppression of angiogenesis and tumor growth [72, 79]. Additional
level dysregulation is also present in primary hepatic diseases. study of this therapeutic approach for other cancers is warranted.

Non‐alcoholic fatty liver disease


Non‐alcoholic fatty liver disease (NAFLD) is the accumulation CONCLUSIONS
of fat in the liver with structural disorganization, in the absence
of excessive alcohol consumption. The severity of steatosis and The liver is the main regulator of copper levels in the blood,
histological changes varies, ranging from benign steatosis to through its role in the distribution and excretion of this trace
18:  Copper Metabolism and the Liver 213

metal. Hepatocytes process copper by directing it to various path- 18. Liu, N., Lo, L.S., Askary, S.H. et al. Transcuprein is a macroglobulin regu-
ways, depending on the body’s needs and on intracellular levels. lated by copper and iron availability. J Nutr Biochem, 2007;18(9):597–608.
19. Hellman, N.E. and Gitlin, J.D. Ceruloplasmin metabolism and function.
Copper is shuttled by the transporter ATP7B into the secretory Annu Rev Nutr, 2002;22:439–58.
compartment of hepatic cells, where it is incorporated into the 20. Hellman, N.E., Kono, S., Mancini, G.M. et  al. Mechanisms of Copper
serum glycoprotein ceruloplasmin. In conditions of excessive Incorporation into Human Ceruloplasmin. J Biol Chem, 2002;277(48):
intracellular copper, the metal is excreted into the bile from the 46632–8.
21. Kuo, Y.M., Gitschier, J., and Packman, S. Developmental expression of the
canalicular (apical) pole of hepatocytes. Defects in ATP7B func-
mouse mottled and toxic milk genes suggests distinct functions for the
tion or trafficking lead to copper accumulation in the liver with Menkes and Wilson disease copper transporters. Hum Mol Genet,
consequent toxicity and hepatocellular injury. These mechanisms 1997;6(7):1043–9.
explain the liver manifestations seen in Wilson disease, Huppke– 22. Pyatskowit, J.W. and Prohaska, J.R. Copper deficient rats and mice both
Brendel syndrome, and MEDNIK syndrome. Copper dyshomeo- develop anemia but only rats have lower plasma and brain iron levels. Comp
Biochem Physiol C Toxicol Pharmacol, 2008;147(3):316–23.
stasis is also found in HCC and NAFLD, with implications for
23. Schilsky, M.L. and Thiele, D.J. Copper metabolism and the liver, in The
lipid metabolism, but the mechanisms underlying these condi- Liver (eds. I.M. Arias et  al.), Wiley‐Blackwell, Chichester, 2009, pp.
tions are incompletely understood. Further insights into copper 221–33.
metabolism at the cellular level, as well as carefully designed 24. Smallwood, R.A., Williams, H.A., Rosenoer, V.M., and Sherlock, S. Liver‐
clinical trials involving novel copper chelators or viral gene ther- copper levels in liver disease: studies using neutron activation analysis.
Lancet, 1968;292(7582):1310–13.
apy are needed to further advance understanding of these disease 25. Wilson, S.A.K. Progressive lenticular degeneration: a familial nervous disease
mechanisms and assess potential therapeutic remedies. associated with cirrhosis of the liver. Brain, 1912;34(4):295–507.
26. Cumings, J.N. The copper and iron content of brain and liver in the normal
and in hepato‐lenticular degeneration. Brain, 1948;71(4):410–15.
27. Bull, P.C., Thomas, G.R., Rommens, J.M., Forbes, J.R., and Cox, D.W. The
REFERENCES Wilson disease gene is a putative copper transporting P‐type ATPase similar
to the Menkes gene. Nat Genet, 1993;5(4):327–37.
1. Baker, Z.N., Cobine, P.A., and Leary, S.C. The mitochondrion: a central 28. Bennett, J. and Hahn, S.H. Clinical molecular diagnosis of Wilson disease.
architect of copper homeostasis. Metallomics, 2017;15;9(11):1501–12. Semin Liver Dis, 2011;31(3):233–8.
2. Bousquet‐Moore, D., Mains, R.E., and Eipper, B.A. PAM and copper  –  a 29. Ferenci, P. Regional distribution of mutations of the ATP7B gene in patients
gene/nutrient interaction critical to nervous system function. J Neurosci Res, with Wilson disease: impact on genetic testing. Hum Genet,
2010;88(12):2535–45. 2006;120(2):151–9.
3. Kaler, S.G. ATP7A‐related copper transport diseases‐emerging concepts and 30. Weiss, K.H. Wilson disease. GeneReviews [internet], 2016. https://www.
future trends. Nat Rev Neurol, 2011;7(1):15–29. ncbi.nlm.nih.gov/books/NBK1512/ (Accessed Oct 26, 2018).
4. McCord, J.M. and Fridovich, I. Superoxide dismutase an enzymic function 31. Gomes, A. and Dedoussis, G.V. Geographic distribution of ATP7B mutations
for erythrocuprein (hemocuprein). J Biol Chem, 1969;244(22):6049–55. in Wilson disease. Ann Hum Biol, 2016;43(1):1–8.
5. Rucker, R.B., Kosonen, T., Clegg, M.S. et  al. Copper, lysyl oxidase, and 32. Chang, I.J. and Hahn, S.H. The genetics of Wilson disease. Handb Clin
extracellular matrix protein cross‐linking. Am J Clin Nutr, 1998;67(5 Neurol, 2017;142:19–34.
Suppl):996S–1002S. 33. Bandmann, O., Weiss, K.H., and Kaler, S.G. Wilson disease and other neuro-
6. Lutsenko, S. Human copper homeostasis: a network of interconnected path- logical copper disorders. Lancet Neurol, 2015;14(1):103–13.
ways. Curr Opin Chem Biol, 2010;14(2):211–17. 34. Roberts, E.A. and Schilsky, M.L. A practice guideline on Wilson disease.
7. Pierson, H., Muchenditsi, A., Kim B.‐E. et al. The function of ATPase copper Hepatology, 2003;37(6):1475–92.
transporter ATP7B in intestine. Gastroenterology, 2018;154(1):168–180.e5. 35. Ferenci, P., Członkowska, A., Merle, U. et  al. Late‐onset Wilson disease.
8. Zerounian, N.R., Redekosky, C., Malpe, R., and Linder, M.C. Regulation of Gastroenterology, 2007;132(4):1294–8.
copper absorption by copper availability in the Caco‐2 cell intestinal model. 36. Kaler, S.G. Inborn errors of copper metabolism, in Handbook of Clinical
Am J Physiol Gastrointest Liver Physiol, 2003;284(5):G739–747. Neurology (eds. O. Dulac, M. Lassonde, and H.B. Sarnat), Elsevier, New
9. Berghe, V.D., Ve, P., and Klomp, L.W. New developments in the regulation of York, 2013, pp. 1745–54.
intestinal copper absorption. Nutr Rev, 2009;67(11):658–72. 37. Machado, A., Chien, H.F., Deguti, M.M. et al. Neurological manifestations
10. Nose, Y., Kim, B.‐E., and Thiele, D.J. Ctr1 drives intestinal copper absorp- in Wilson disease: report of 119 cases. Mov Disord, 2006;21(12):2192–6.
tion and is essential for growth, iron metabolism, and neonatal cardiac func- 38. Ala, A., Borjigin, J., Rochwarger, A., Schilsky, M. Wilson disease in septua-
tion. Cell Metab, 2006;4(3):235–44. genarian siblings: raising the bar for diagnosis. Hepatology, 2005;41(3):
11. Lee, J., Peña, M.M.O., Nose, Y., and Thiele, D.J. Biochemical characteriza- 668–70.
tion of the human copper transporter Ctr1. J Biol Chem, 2002;277(6): 39. Kim, J.W., Kim, J.H., Seo, J.K. et al. Genetically confirmed Wilson disease
4380–7. in a 9‐month old boy with elevations of aminotransferases. World J Hepatol,
12. Garrick, M.D., Singleton, S.T., Vargas, F. et al. DMT1: which metals does it 2013;5(3):156–9.
transport? Biol Res, 2006;39(1):79–85. 40. Ferenci, P., Caca, K., Loudianos, G. et al. Diagnosis and phenotypic classifi-
13. Lutsenko, S., Barnes, N.L., Bartee, M.Y., and Dmitriev, O.Y. Function and cation of Wilson disease. Liver Int, 2003;23(3):139–42.
regulation of human copper‐transporting ATPases. Physiol Rev, 2007;87(3): 41. Roberts, E.A. and Schilsky, M.L. Diagnosis and treatment of Wilson disease:
1011–46. an update. Hepatology, 2008;47(6):2089–111.
14. Kim, B.‐E., Turski, M.L., Nose, Y. et al. Cardiac copper deficiency activates 42. Schilsky, M.L. Liver transplantation for Wilson disease. Ann N Y Acad Sci,
a systemic signaling mechanism that communicates with the copper acquisi- 2014;1315:45–9.
tion and storage organs. Cell Metab, 2010;11(5):353–63. 43. Kaler, S.G. Microbial peptide de‐coppers mitochondria: implications for
15. Linz, R. and Lutsenko, S. Copper‐transporting ATPases ATP7A and ATP7B: Wilson disease. J Clin Invest, 2016;126(7):2412–4.
cousins, not twins. J Bioenerg Biomembr, 2007;39(5–6):403–7. 44. Lichtmannegger, J., Leitzinger, C., Wimmer, R. et  al. Methanobactin
16. Monty, J.‐F., Llanos, R.M., Mercer, J.F.B., and Kramer, D.R. Copper expo- reverses acute liver failure in a rat model of Wilson disease. J Clin Invest,
sure induces trafficking of the menkes protein in intestinal epithelium of 2016;126(7):2721–35.
ATP7A transgenic mice. J Nutr, 2005;135(12):2762–6. 45. Ranucci, G., Polishchuck, R., and Iorio, R. Wilson disease: prospective
17. Ravia, J.J., Stephen, R.M., Ghishan, F.K., and Collins, J.F. Menkes copper developments towards new therapies. World J Gastroenterol, 2017;23(30):
ATPase (Atp7a) is a novel metal‐responsive gene in rat duodenum, and 5451–6.
immunoreactive protein is present on brush‐border and basolateral mem- 46. Meng, Y., Miyoshi, I., Hirabayashi, M. et al. Restoration of copper metabolism
brane domains. J Biol Chem, 2005;280(43):36221–7. and rescue of hepatic abnormalities in LEC rats, an animal model of Wilson
214 THE LIVER:  REFERENCES

disease, by expression of human ATP7B gene. Biochim Biophys Acta, 62. al‐Othman, A.A., Rosenstein, F., and Lei, K.Y. Copper deficiency increases
2004;1690(3):208–19. in vivo hepatic synthesis of fatty acids, triacylglycerols, and phospholipids in
47. Roybal, J.L., Endo, M., Radu, A. et al. Early gestational gene transfer with rats. Proc Soc Exp Biol Med, 1993;204(1):97–103.
targeted ATP7B expression in the liver improves phenotype in a murine 63. Tosco, A., Fontanella, B., Danise, R. et al. Molecular bases of copper and
model of Wilson disease. Gene Ther, 2012;19(11):1085–94. iron deficiency‐associated dyslipidemia: a microarray analysis of the rat
48. Murillo, O., Luqui, D.M., Gazquez, C. et al. Long‐term metabolic correction intestinal transcriptome. Genes Nutr, 2010;5(1):1–8.
of Wilson disease in a murine model by gene therapy. J Hepatol, 64. Ibdah, J.A., Perlegas, P., Zhao, Y. et al. Mice heterozygous for a defect in
2016;64(2):419–26. mitochondrial trifunctional protein develop hepatic steatosis and insulin
49. Murillo, O., Moreno, D., Gazquez, C. et al. Improvement of gene therapy for resistance. Gastroenterology, 2005;128(5):1381–90.
Wilson disease. Mol Ther, 2016;24:S64. 65. Apelgot, S., Coppey, J., Fromentin, A. et al. Altered distribution of copper
50. Huppke, P., Brendel, C., Kalscheuer, V. et al. Mutations in SLC33A1 cause a (64Cu) in tumor‐bearing mice and rats. Anticancer Res, 1986;6(2):159–64.
lethal autosomal‐recessive disorder with congenital cataracts, hearing loss, and 66. Daniel, K.G., Harbach, R.H., Guida, W.C., and Dou, Q.P. Copper storage
low serum copper and ceruloplasmin. Am J Hum Genet, 2012;90(1):61–8. diseases: Menkes, Wilsons, and cancer. Front Biosci, 2004;9:2652–62.
51. Horváth, R., Freisinger, P., Rubio, R. et  al. Congenital cataract, muscular 67. Gupte, A. and Mumper, R.J. Elevated copper and oxidative stress in cancer
hypotonia, developmental delay and sensorineural hearing loss associated cells as a target for cancer treatment. Cancer Treat Rev, 2009;35(1):32–46.
with a defect in copper metabolism. J Inherit Metab Dis, 2005; 68. Brewer, G.J. Copper lowering therapy with tetrathiomolybdate as an antian-
28(4):479–92. giogenic strategy in cancer. Curr Cancer Drug Targets, 2005;5(3):195–202.
52. Kanamori, A., Nakayama, J., Fukuda, M.N. et  al. Expression cloning and 69. Lowndes, S.A. and Harris, A.L. The role of copper in tumour angiogenesis.
characterization of a cDNA encoding a novel membrane protein required for J Mammary Gland Biol Neoplasia, 2005;10(4):299–310.
the formation of O‐acetylated ganglioside: a putative acetyl‐CoA transporter. 70. Sproull, M., Brechbiel, M., and Camphausen, K. Antiangiogenic therapy
Proc Natl Acad Sci U S A, 1997;94(7):2897–902. through copper chelation. Expert Opin Ther Targets, 2003;7(3):405–9.
53. Hirabayashi, Y., Nomura, K.H., and Nomura, K. The acetyl‐CoA transporter 71. Goodman, V.L., Brewer, G.J., and Merajver, S.D. Copper deficiency as an
family SLC33. Mol Aspects Med, 2013;34(2–3):586–9. anti‐cancer strategy. Endocr Relat Cancer, 2004;11(2):255–63.
54. Montpetit, A., Côté, S., Brustein, E. et al. Disruption of AP1S1, causing a 72. Yoshii, J., Yoshiji, H., Kuriyama, S. et al. The copper‐chelating agent, trien-
novel neurocutaneous syndrome, perturbs development of the skin and spi- tine, suppresses tumor development and angiogenesis in the murine hepato-
nal cord. PLoS Genet, 2008;4(12):e1000296. cellular carcinoma cells. Int J Cancer, 2001;94(6):768–73.
55. Martinelli, D., Travaglini, L., Drouin, C.A. et  al. MEDNIK syndrome: a 73. Vecchio, F.M., Federico, F., and Dina, M.A. Copper and hepatocellular car-
novel defect of copper metabolism treatable by zinc acetate therapy. Brain, cinoma. Digestion, 1986;35(2):109–14.
2013;136(Pt 3):872–81. 74. Ebara, M., Fukuda, H., Hatano, R. et al. Relationship between copper, zinc
56. Holloway, Z.G., Velayos‐Baeza, A., Howell, G.J. et  al. Trafficking of the and metallothionein in hepatocellular carcinoma and its surrounding liver
Menkes copper transporter ATP7A is regulated by clathrin‐, AP‐2‐, AP‐1‐, parenchyma. J Hepatol, 2000;33(3):415–22.
and Rab22‐dependent steps. Mol Biol Cell, 2013 Jun 1;24(11):1735–48. 75. Ebara, M., Fukuda, H., Hatano, R. et al. Metal contents in the liver of patients
57. Yi, L. and Kaler, S.G. Direct interactions of adaptor protein complexes 1 and with chronic liver disease caused by hepatitis C virus. Reference to hepato-
2 with the copper transporter ATP7A mediate its anterograde and retrograde cellular carcinoma. Oncology, 2003;65(4):323–30.
trafficking. Hum Mol Genet, 2015;24(9):2411–25. 76. Casaril, M., Capra, F., Marchiori, L. et al. Serum copper and ceruloplasmin
58. Adams, L.A., Angulo, P., and Lindor, K.D. Nonalcoholic fatty liver disease. in early and in advanced hepatocellular carcinoma: diagnostic and prognos-
CMAJ, 2005;172(7):899–905. tic relevance. Tumori, 1989;75(5):498–502.
59. Greenfield, V., Cheung, O., and Sanyal, A.J. Recent advances in nonalcholic 77. Haratake, J., Horie, A., Takeda, S. et al. Tissue copper content in primary and
fatty liver disease. Curr Opin Gastroenterol, 2008;24(3):320–7. metastatic liver cancers. Acta Pathol Jpn, 1987;37(2):231–8.
60. Aigner, E., Strasser, M., Haufe, H. et al. A role for low hepatic copper con- 78. Gurusamy, K. and Davidson, B.R. Trace element concentration in metastatic
centrations in nonalcoholic fatty liver disease. Am J Gastroenterol, liver disease: a systematic review. J Trace Elem Med Biol,
2010;105(9):1978–85. 2007;21(3):169–77.
61. al‐Othman, A.A., Rosenstein, F., Lei, K.Y. Copper deficiency alters plasma 79. Yoshiji, H., Kuriyama, S., Yoshii, J. et al. The copper‐chelating agent, trientine,
pool size, percent composition and concentration of lipoprotein components attenuates liver enzyme‐altered preneoplastic lesions in rats by angiogenesis
in rats. J Nutr, 1992;122(6):1199–204. suppression. Oncol Rep, 2003;10(5):1369–73.
The Central Role of the Liver
19 in Iron Storage and Regulation
of Systemic Iron Homeostasis
Tracey A. Rouault1, Victor R. Gordeuk2, and Gregory J. Anderson3
1
Eunice Kennedy Shriver National Institute of Child Health and Human Development, National Institutes of
Health, Bethesda, MD, USA
2
University of Illinois at Chicago, Chicago, IL, USA
3
QIMR Berghofer Medical Research Institute, Brisbane, Queensland, Australia

THE LIVER AS A MAJOR IRON Alternatively, ferritin multimers can undergo degradation in
REPOSITORY lysosomes to release iron. This process is guided by interac-
tion of ferritin with NCOA4, which binds to ferritin and
In healthy adult humans, the liver is estimated to store from ­delivers it to the lysosome (Figure 19.1) [9].
0.07 to 0.4 g of iron. A combined analysis of previous studies
using quantitative phlebotomy indicated that total body
median storage iron was approximately 0.8 g among 39 nor-
mal men predominantly in the third and fourth decades of life, SERUM TRANSFERRIN IS THE MAJOR
and 0.4 g among 20 normal women of similar age [1–5]. SOURCE OF IRON FOR TISSUES
Based on the assumption that one‐third of iron stores are nor-
mally in the liver, this would translate to a normal median Transferrin (TF) is a high‐affinity binding protein for ferric iron
hepatic iron content of 0.27 g for men and 0.13 g for women. that circulates in the plasma with a concentration of 9–28 nmol
Iron is sequestered within the iron storage protein ferritin, a L−1. It is synthesized mainly by hepatocytes and is secreted into
large 24‐subunit protein composed of ferritin L (light) and H the circulation. The plasma usually contains an excess of TF,
(heavy) chains (Figure 19.1). The ferritin subunits combine to such that less than one‐third of the iron‐binding sites on TF are
form a spherical protein shell that contains numerous chan- occupied by iron. TF contains two lobes, each of which contains
nels. On the inside of the spherical protein, iron is oxidized a ferric (Fe3+) iron‐binding site [11]. Cells in tissues throughout
from the ferrous (Fe2+) form to the ferric (Fe3+) form by the the body can obtain iron from circulating TF by increasing
ferroxidase activity of the ferritin H chain, and insoluble ferric expression of transferrin receptors (TFRs). TFR1 is the major
iron deposits grow from their initial deposition sites on side‐ receptor for TF and binds diferric TF with high affinity (107–109
chains of ferritin L subunits [6, 7]. As ferritin iron uptake and M−1) at physiological pH. Monoferric and apoTF are bound with
oxidation continue, thousands of ferric iron atoms accumulate much lower affinity (106 and <105 M−1 respectively) [12]. Upon
within the ferritin sphere, which has the capacity to store up to binding diferric TF, a TF–TFR complex forms and then inter-
4500 iron atoms. As the iron contained within ferritin is not nalizes within endosomes. As endosomes undergo acidification,
very bioavailable, ferritin sequesters iron and reduces the iron is released from the TFR–TF complex. The released iron is
availability of cytosolic iron to interact with oxygen and gen- still in the oxidized ferric form (Fe3+) and must undergo reduc-
erate harmful reactive oxygen species. Ferritin also serves as a tion by a reductase of the STEAP family of proteins [13] prior
source of iron for the metabolic needs of the cell. Iron can be to being transported from within the endosome to the cytosol by
released from intact ferritin that has been mono‐ubiquitinated a divalent metal‐ion transporter (DMT) known as DMT1 [14].
and will be ultimately degraded by the proteasome [8]. When iron enters the cytosol, it can be incorporated into many

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
216 THE LIVER:  NON‐TRANSFERRIN‐BOUND IRON UPTAKE AND HEPATIC IRON OVERLOAD

TFR2 EXPRESSION IS REGULATED BY


DIFFERENT MECHANISMS THAN TFR1
Aside from the fact that TFR2 is mainly expressed in hepatocytes,
whereas TFR1 is expressed highly in erythroid progenitors and
to varying degrees in most other tissues, the major difference
between the two receptors is a difference in the mechanisms by
which their expression is regulated. TFR1 is regulated at the
transcriptional level by hypoxia [23, 24], but it is also potently
regulated by a post‐transcriptional regulatory system that con-
sists of iron‐regulatory proteins (IRPs), which bind to RNA
stem‐loops known as iron responsive elements (IREs) in tran-
scripts such as ferritin and TFR1 (Figure 19.2) [25]. In cells that
have cytosolic iron deficiency, IRPs bind to IREs, and IRP bind-
ing prevents translation of transcripts such as those encoding the
ferritin H and L chains, which have IREs located within their 5′
UTRs (Figure 19.3a). Conversely, IRP binding to IREs in the 3′
UTR of TFR1 stabilizes the transcript, leading to an increase in
mRNA levels, and a concomitant increase in TFR1 protein lev-
Figure 19.1  Ferritin is a 24 subunit protein in which subunits of H or
L chains coassemble to create a protein spherical structure with a cen- els (Figure 19.3b). In most cells that are iron‐loaded, IRPs do
tral cavity. In this depiction, generated from coordinates of a reported not bind to IREs, and ferritin synthesis increases, whereas TFR1
ferritin crystal structure [10], three separate subunits on the anterior of expression decreases. Thus, iron loading in most cells leads to
the spherical protein are in dark gray, to show the arrangement of indi- an increase in iron sequestration and to a decrease in TFR1‐
vidual subunits. Pores in the structure allow entry of iron, which accu- mediated iron uptake (reviewed in [25, 26]).
mulates as a ferric precipitate within the hollow core. Each ferritin Hepatocytes differ from other cells because they express high
multimer can sequester up to 4500 iron atoms.
levels of TFR2, and TFR2 mRNA levels remain high with iron
loading [27]. In fact, binding of diferric TF leads to increased
enzymes and proteins that require iron for function. Iron can TFR2 stability by changing TFR2 trafficking [28]. One reason
also be transported to subcellular locations such as mitochon- why hepatocytes are an important iron repository may thus be
dria, which use an iron importer known as mitoferrin to facili- because TFR2‐dependent iron uptake continues even when
tate mitochondrial iron uptake [15]. expression of TFR1 is diminished by iron overload in other cells
throughout the body. The propensity to maintain iron uptake
through non‐IRP‐regulated TFR expression may also be valua-
ble for the hepatocyte in its role as a systemic iron regulator, as
THE ROLE OF TFR2, A SECOND discussed later. Although TFR2 is implicated in a signaling
TRANSFERRIN RECEPTOR HIGHLY pathway that responds to iron status (see later), the liver also has
a strong capacity to take up non‐transferrin‐bound iron (NTBI).
EXPRESSED BY HEPATOCYTES
Whereas in most tissues TFR1 is the first component of the
major cellular iron uptake pathway, the liver expresses large
amounts of a second, homologous TFR, TFR2 [16]. In fact, NON‐TRANSFERRIN‐BOUND IRON
TFR2 appears to be considerably more abundant in the liver than UPTAKE AND HEPATIC IRON OVERLOAD
TFR1 [17]. Despite their homology, recent data suggest that
there are differences in the way the two TFRs bind TF [18]. The In addition to the uptake of TF‐bound iron, the liver can efficiently
finding that TFR2 has a 25‐fold lower affinity for diferric TF take up iron that is not sequestered by TF (i.e. NTBI). NTBI
than TFR1 [16, 19] led to the supposition that TFR2 was not becomes detectable in the plasma mainly when TF iron‐binding
important for iron uptake. However, the concentration of diferric sites become fully saturated, as occurs in various diseases of
TF in plasma is estimated at about 8–12 μM [20], a concentration systemic iron overload. The plasma half‐life of NTBI is less
that would likely be sufficient to saturate both TFRs, suggesting than 30 seconds, whereas TF‐bound iron has a half‐life of
that both might play a role in hepatic iron uptake. Despite this, approximately 50 minutes [29, 30]. Thus, the isolated perfused
studies under defined conditions in hepatoma cells demonstrate liver can remove 60–75% of NTBI from plasma on first pass,
that the TFR1‐mediated uptake of TF‐bound iron is by far the but less than 2% of TF‐bound iron [29, 31, 32]. In addition, if
dominant iron uptake pathway [21]. In addition, the livers of plasma TF is first saturated with iron, the clearance of a subse-
patients with mutations in TFR2 or mice lacking TFR2 readily quent dose of radioactive iron, given either orally or intrave-
load with iron, indicating that TFR2 is dispensible for hepatic nously, is very rapid, with the liver being the major site of iron
iron upake [22]. As will be described in more detail below, the deposition [30]. Perhaps the best demonstration of the capacity
major role of TFR2 in the liver is as a regulator of hepcidin, and of the liver to acquire NTBI comes from studies of congenital
hence it plays a critical role in body iron homeostasis. hypotransferrinemia, where little TF is synthesized in either a
19:  The Central Role of the Liver in Iron Storage and Regulation of Systemic Iron Homeostasis 217

A consensus motif for the IRE


G

GU can also be A
A G C
C X U/C/A
G
N-N′
N-N′
N-N′ 5 base pair stem
N-N′
N-N′ C
Unpaired C
nucleoside − N-N′
usually cytosine N-N′ Lower stem
N-N′ (at least 3 base pairs)
5′
5′-N-N′-3′

Consensus IRE
3′

Functional IREs in 5′UTRs Functional IREs in 3′UTRs


Ferritin Transferrin receptor
Mitochondrial aconitase Divalent metal transporter
Succinate dehydrogenase D. melanogaster (non-consensus)
eALASynthase CDC14A
Ferroportin-basolateral iron exporter
HIF2 alpha

Figure 19.2  Iron responsive elements (IREs) are RNA stem‐loops found in the mRNAs encoding numerous proteins important in iron homeosta-
sis. IREs consist of a base‐paired stem, an unpaired cytosine that separates the upper and lower stems, and a six‐membered loop with the sequence
CAGUGX, in which X can be any residue except G and there is a base pair between A2 and G5 of the loop. Reproduced from [25] with permission
of Elsevier.

mouse model [33] or human patients [34]. In this condition, the and this activity can be supplied by the prion protein Pr(C) [44].
liver and some other organs develop massive iron loading at a Early studies on the kinetics of NTBI uptake indicated that the
very young age. Such studies show that the role of TF is not process was increased under iron loading conditions and it was
simply to deliver iron to tissues, but to allow the regulated deliv- suggested that this plays a protective role by removing poten-
ery of the metal. TF has such a high binding affinity for iron tially toxic “free” iron from the circulation [45]. Consistent with
under normal physiological conditions (affinity constant 1023 this finding is the more recent observation that ZIP14 levels are
M−1) [35] that the amount of circulating NTBI is extremely upregulated by iron loading [46].
small and is essentially undetectable. Where NTBI does become
important is in various pathological states, particularly iron‐
loading disorders in which the capacity of plasma TF to bind
iron is exceeded. In HFE‐associated hemochromatosis, for THE RELEASE OF IRON FROM HEPATIC
example, NTBI can reach 10–15 μM [36]. It has proved difficult CELLS
to precisely define the chemical nature of NTBI, but most of it
may be chelated by small organic acids such as citrate and Although a great deal is known about the uptake of iron, the
amino acids, and some can also be bound by proteins such as mechanisms by which the liver can release stored iron in times
albumin [37]. of need are less well understood. The capacity of the liver to
Until recently, the uptake pathway of NTBI had been a relinquish stored iron is particularly well illustrated during
­significant unresolved problem in iron homeostasis. Early stud- phlebotomy therapy for primary iron‐loading diseases such as
ies suggested that NTBI uptake may be mediated by DMT1 [38, HFE‐associated hemochromatosis [47]. In such diseases, the
39] but when DMT1 is disrupted, the liver is still able to accu- liver may contain 20 g or more of stored iron, but all of this can
mulate NTBI, indicating that other uptake pathways were be mobilized for hemoglobin synthesis following phlebotomy.
involved [40, 41]. Another candidate is ZRT‐ and IRT‐like pro- While in HFE‐associated hemochromatosis most of the iron
tein 14 (ZIP14) or SLC39A14 [42]. This plasma membrane storage in the liver occurs in hepatocytes (see later), in transfu-
transporter is expressed in the liver and was shown to mediate sional iron overload the resident macrophages (the Kupffer
NTBI uptake in hepatoma cells [43]. Subsequent studies with cells) also can store much of the iron, and iron can be mobilized
ZIP14‐knockout mice have shown that ZIP14 provides the main from Kupffer cells as well as hepatocytes by phlebotomy. In
pathway for the uptake of NTBI by the liver and pancreas [43]. addition, Kupffer cells play an important role in the recycling of
Indeed, the same study showed that deletion of ZIP14 in mice hemoglobin‐derived iron from senescent red blood cells, and
can completely prevent hepatic iron loading in mouse models of this iron must also be released into the circulation [48].
hemochromatosis. Interestingly, the reduction of ferric iron The protein responsible for iron release from most body cells
appears to be required before it can be utilized by hepatic ZIP14 is ferroportin (FPN) [49–51], a highly conserved ferrous iron
218 THE LIVER:  THE RELEASE OF IRON FROM HEPATIC CELLS

(a)
Ferritin mRNA One IRE in 5′UTR

5′ 3′
AUG

−Fe +Fe
IRP

40s AAAAAA

IRE occupied by IRP,


inhibiting translation
initiation

IRE unoccupied, allowing


60s polysome formation and
increased ferritin synthesis
(b)
TFR mRNA Five IREs in 3′UTR

AAAAAA
Protein 3′
5′
coding

−Fe

IRP +Fe

Endonuclease
AAAAAA cleavage site
Protein
coding AAAAAA
Protein
coding
One or more IREs occupied
by IRP, protecting mRNA
from rate-determining IRE unoccupied, rendering
step mRNA degradation mRNA susceptible to an
endonuclease

Figure 19.3  (a) Iron‐regulatory proteins (IRP) 1 and 2 repress translation of transcripts that contain an iron responsive element (IRE) in the 5′
UTR of the transcript by preventing binding of translation factors and ribosomes. (b) Binding of IRPs to IREs in the 3′ UTR protects the transferrin
receptor (TFR1) transcript from degradation, and likely also protects other transcripts that contain a 3′ IRE by a similar mechanism.

export protein that contains 9 or 10 transmembrane domains. supplied by circulating ceruloplasmin (CP) [54]. CP is the major
While FPN is most highly expressed in cell types that efflux copper‐containing protein in the plasma, but it is a protein of
large amounts of iron, such as intestinal enterocytes and mac- iron metabolism rather than copper metabolism. In the absence
rophages [52], it also performs an iron efflux role in all cells of CP (e.g. in rare cases of aceruloplasminemia), the plasma
studied to date, including hepatocytes [53]. FPN mRNA con- iron level is reduced and iron accumulates in the tissues, whereas
tains an IRE in its 5′ UTR, like the ferritin genes, and its expres- copper homeostasis is unaffected [55, 56]. Recent data have
sion in the liver increases with iron loading, suggesting that it is indicated that the membrane‐bound CP homolog hephaestin
under iron‐dependent translational control through the IRE/IRP (HEPH) can also contribute to hepatic iron efflux [57]. How CP
system [51]. Since FPN levels increase with iron loading, it is contributes to iron efflux is unclear, but there is some evidence
feasible that this increase represents a protective mechanism to that a glycophosphatidylinositol (GPI)‐linked form of CP on
help the liver limit its iron accumulation. glioma cells can stabilize FPN on the cell surface by oxidizing
Efficient export of iron from tissues requires an iron oxidase transported ferrous iron [58], and HEPH has been reported to
in addition to FPN. For most tissues the oxidase activity is stabilize FPN in hippocampal neurons [59]. More plasma
19:  The Central Role of the Liver in Iron Storage and Regulation of Systemic Iron Homeostasis 219

membrane FPN means increased cellular iron release. Whether degraded in lysosomes [61–63], which results in reduced export
this is a general mechanism of CP/HEPH action that also applies of iron because FPN is the only known iron exporter [66, 67].
to the liver remains to be determined. The liver is the principal This is the proposed mechanism by which hepcidin reduces iron
site of synthesis of CP and thus it plays an important role sys- efflux from macrophages, duodenal enterocytes, hepatocytes,
temically in facilitating iron efflux. and other target cells.
Although hepcidin‐dependent degradation of FPN is accepted
as a major mechanism by which intestinal iron uptake is
decreased by hepcidin overexpression, other levels of regulation
HEPATOCYTES SYNTHESIZE also exist. For instance, expression of the mRNAs encoding
AND SECRETE THE REGULATORY FPN, DMT1, and DCYTB is increased in enterocytes under
PEPTIDE HORMONE HEPCIDIN iron‐deficient conditions [68] and this cannot be readily
explained through the direct actions of hepcidin. These changes
In systemic iron regulation, secretion of the peptide hormone may be secondary to reduced iron levels in the enterocytes,
hepcidin has an important role in determining how much iron is which could influence DMT1 levels via the IRE/IRP system and
released from macrophages, and likely other body cells, and the expression of FPN and DCYTB through the hypoxia‐induc-
how much iron is absorbed through the duodenum [60]. ible factor HIF-2 [69]. This could also explain why the effects of
Hepcidin is related to other peptide hormones known as hepcidin on macrophages occur before the duodenal mucosa is
defensins that function in defense against bacterial infections. It affected [70]. Changes in the subcellular localization of FPN
is synthesized as an 84 amino acid precursor, which undergoes and DMT1 can also occur in enterocytes with changes in iron
cleavage and is released into the circulation as an active 25 levels, with higher levels on the plasma membrane under iron‐
amino acid peptide hormone [61]. Forms of hepcidin that are deficient conditions, adding further to the complexity [71, 72].
truncated at the N‐terminus and contain 22 or 20 amino acids
are also present in serum, but these forms are not active [61].
(Figure 19.4 shows the hepcidin sequence and the positions of
disulfide bonds.) TRANSCRIPTIONAL REGULATION
As a circulating peptide, hepcidin can interact with tissues OF HEPCIDIN EXPRESSION IS THE MAJOR
and cells throughout the body. Hepcidin binds to the iron
exporter FPN on target cell membranes, where it forms a com-
DETERMINANT OF SYSTEMIC IRON
plex that internalizes and undergoes ubiquitination and lysoso- HOMEOSTASIS
mal degradation [62, 63]. This in turn reduces the flow of iron
into the circulation. It is not yet clear if this mechanism of action Hepcidin is the central regulator of systemic iron homeostasis as
can explain all of the effects of hepcidin (see later). it coordinates release of iron to serum from macrophages, intes-
The transcription of hepcidin is low when there is systemic tinal enterocytes, and other body cells. Whereas it was once
iron depletion, and high when there is systemic iron overload. believed that sensing of systemic iron homeostasis occurred in
Levels are also reduced when erythropoiesis is stimulated, and the cells of the duodenal mucosa, numerous studies now indi-
increased under inflammatory conditions. The mechanisms by cate that the most important site of sensing is within hepatocytes
which hepatocytes transduce information about systemic iron [65, 73], with signaling information converging at the hepcidin
status into transcriptional activity of the hepcidin promoter have promoter. Over the last 15 years a great deal has been learned
been extensively investigated [65], but the regulatory system about the mechanisms regulating the expression of the HAMP
remains incompletely understood. gene, which encodes hepcidin. The BMP–SMAD signaling
In general, peptide hormones such as hepcidin effect changes pathway has proved to be essential for both basal and regulated
by binding to specific receptors in target cell populations. transcription of HAMP [74], but a number of other pathways are
Binding of a ligand to a receptor often activates a process of also involved. These explain how hepcidin expression responds
signal transduction, a cascade of intracellular events that to changes in iron levels, inflammation, erythropoiesis, and sev-
changes the transcription of various genes in the target cell pop- eral other stimuli. These are summarized in Figure 19.5. A great
ulation. Hepcidin signaling does not conform to this general deal of our knowledge of hepcidin regulation has come from the
pathway, because hepcidin binds to the iron exporter FPN rather analysis of human disorders of iron homeostasis, and particu-
than to a member of a previously recognized receptor family. larly the various forms of iron‐loading disease known as heredi-
Upon binding hepcidin, FPN is internalized and subsequently tary hemochromatosis. These conditions are briefly summarized
here, but are considered in more detail is subsequent sections.
The essential role of the BMP–SMAD pathway in hepci-
1 5 10 15 20 25 din regulation was identified in family studies of patients
who develop severe juvenile‐onset hemochromatosis due to
DTHFPICIFCCGCCHRSKCGMCCKT mutations in the hemojuvelin gene (HJV) [75]. HJV is a
member of the bone morphogenetic protein receptor (BMPR)
Figure 19.4  Sequence of the peptide hormone hepcidin. The peptide
contains eight cysteines and is believed to contain four disulfide bonds.
family and the signaling pathways of these receptors have
Arcs indicate the proposed cysteine pairs that form the disulfide bonds been studied extensively in other physiological settings. In
that lead to formation of a hairpin in the hepcidin structure. Reproduced general, BMP receptors function by activating SMAD pro-
from [64] with permission of Elsevier. teins, a family of transcriptional activators/repressors that
220 THE LIVER:  TRANSCRIPTIONAL REGULATION OF HEPCIDIN EXPRESSION IS THE MAJOR DETERMINANT

Figure 19.5  The hepcidin signaling pathway in hepatocytes in response to iron and inflammation. Changes in the expression of hepcidin in
response to altered body iron requirements occur through at least two pathways. One is through the iron‐dependent regulation of BMP6 and BMP2
in hepatic nonparenchymal cells, which subsequently signal through the BMPR/SMAD pathway on hepatocytes to alter HAMP transcription. The
other is through the influence of diferric TF, which likely modulates the interactions between HFE and TFRs 1 and 2. How this in turn alters HAMP
expression is not known, but it is likely to involve effects on BMP/SMAD signaling. Interleukin 6 (IL‐6) signals through a separate surface receptor.
Transcription factors activated by this pathway activate hepcidin transcription by binding to a STAT3‐binding site. Ultimately, hepcidin (HAMP)
transcription increases in hepatocytes that sense high systemic iron or inflammatory signals.

function in numerous pathways to connect receptor‐binding recent data have suggested that that the system is far more
events to changes in nuclear transcription of target genes complicated and that the enzyme can cleave multiple compo-
[74]. Notably, animals that lack hepatocyte SMAD4, a pro- nents of the hepcidin regulatory pathway in addition to HJV,
tein that combines with other members of the SMAD family including multiple BMPRs (ALK2, ALK3, ACTRIIA,
to regulate transcription of target genes, develop significant BMPR2) as well as HFE and TFR2 [84]. Consistent with
iron overload associated with a profound reduction in hepci- this, TMPRSS6 has been shown to suppress hepcidin expres-
din expression [76]. Multiple studies have now shown that a sion independently of HJV [84].
functioning BMP–SMAD pathway is essential for HAMP The first gene identified as a cause of hereditary hemochro-
transcription, but the pathway is also important for the iron‐ matosis when mutated, HFE [85], is an upstream regulator
dependent regulation of hepcidin. The levels of two BMPR of  hepcidin expression. The mechanisms by which HFE
ligands, BMP2 and BMP6, have been shown to be increased ­influences HAMP expression are incompletely understood,
when body iron levels rise and these in turn can signal although early studies suggested that it could modulate B­ MP–
through the BMP–SMAD pathway to increase HAMP tran- SMAD signaling [86, 87]. Patients with HFE‐related
scription [77–79]. hemochromatosis have reduced hepcidin expression and con-
Another link between the BMP–SMAD pathway and hep- sequently increased dietary iron absorption, hence the iron
cidin expression is provided by the transmembrane serine loading [88]. Much rarer than mutations in HFE are those in
protease TMPRSS6. TMPRSS6 acts to suppress BMP– the TFR2 gene, but they also lead to reduced hepcidin expres-
SMAD signaling, so when TMPRSS6 is defective in either sion and hemochromatosis [89], and TFR2 also appears to
mice or humans [80, 81] this suppression is relieved and hep- affect BMP–SMAD signaling [90]. How HFE and TFR2 reg-
cidin levels are constitutively high. This in turn leads to sys- ulate hepcidin expression is only partially understood. HFE is
temic iron deficiency due to reduced dietary iron absorption, able to interact with TFR1 and HFE and TF can compete for
and in humans the condition is known as iron‐refractory iron‐ binding to TFR1 [91, 92]. An intact HFE/TFR1 complex
deficiency anemia [81]. It was originally thought that the appears to be able to reduce hepcidin expression [93], so it
main substrate of TMPRSS6 was HJV [82, 83], but more has been proposed that when diferric TF levels are high, the
19:  The Central Role of the Liver in Iron Storage and Regulation of Systemic Iron Homeostasis 221

complex dissociates and hepcidin levels are increased [65]. MUTATIONS IN GENES INVOLVED
Other studies have shown that TFR2 binds HFE when both IN HEPATOCYTE SENSING OF SYSTEMIC
proteins are overexpressed [94], and HFE binding appears to
increase the affinity of TFR2 for diferric TF [95]. Thus,
IRON LEVELS CAUSE
TFR2‐dependent TF binding may play an important role in HEMOCHROMATOSIS
hepatocyte iron sensing and systemic iron homeostasis.
Precisely how HFE and TFR2 might work together to alter “Hemochromatosis” is the term for a group of inherited diseases
hepcidin expression is unclear. One possibility is that under characterized initially by increased serum iron levels, and later
low iron conditions HFE binds to TFR1, but when the level of by parenchymal iron overload in the liver, heart, and endocrine
diferric TF rises, it displaces HFE from TFR1 and HFE then organs [113]. Macrophages and the duodenal mucosa play
binds to TFR2 and alters hepcidin transcription, possibly by ­particularly important roles in the disturbed iron homeostasis
modulating BMP/SMAD signaling [94, 96]. Interestingly, associated with hemochromatosis. Macrophages are normally a
when both HFE and TFR2 are knocked out together in mice, major repository of iron in the body, because they metabolize
the phenotype is more severe than either of the single knock- senescent red cells and store any iron derived from heme catab-
outs, suggesting the two proteins might have at least some olism that is not released to plasma transferrin. Whereas the
independent effects [97]. Recent data suggest that HFE can bone marrow macrophages of non‐hemochromatosis patients
also stabilize some BMP receptors [98], adding a further layer are usually rich in stored iron, as assessed by Prussian blue iron
of complexity to this regulatory system. stain, the macrophages of hemochromatosis patients have
Plasma levels of diferric TF have been strongly implicated reduced ability to retain iron retrieved from erythrophagocyto-
in the hepcidin signaling pathway [99]. The degree of satura- sis, and consequently hepatic and splenic macrophages contain
tion of plasma TF saturation integrates the removal and little stainable iron. Moreover, despite high serum iron levels,
return of iron to the circulation from many tissue sources into duodenal iron absorption remains relatively high in hemochro-
one common value. Thus, the TF saturation may be the prin- matosis patients. As the disease progresses, increased serum
cipal clinically measurable indicator of iron status that con- iron levels cause parenchymal iron overload in the liver, heart,
veys information about bodily iron status to multiple tissues and endocrine organs.
[65, 100–102]. Such a value could feed into the HFE/TFR2‐ There are now five different genes whose products are recog-
mediated pathway of hepcidin regulation, but could also nized to cause forms of genetic hemochromatosis. These are
affect the expression of BMPs 2 and 6. It is likely that multi- HFE, TFR2, HJV, FPN (in part), and HAMP, and it is likely that
ple pathways are involved. Recently, levels of diferric TF the list of disease genes will grow as research continues
have also been implicated in mediating (at least in part) the (Figure 19.6). All of these disorders are characterized by abnor-
suppression of HAMP expression in response to stimulated mally reduced hepcidin production in conjunction with elevated
erythropoiesis [103]. plasma transferrin saturation and increased parenchymal rather
In addition to responding to signaling through the BMP– than macrophage iron stores. Fortunately, impaired function of
SMAD pathway, hepcidin expression is strongly induced by these genes does not adversely affect hematopoiesis. As a result,
inflammation (Figure 19.5) [104]. This response of hepcidin iron balance can be restored by regular phlebotomy procedures
is considered to be a major determinant (though likely not the to remove red cells rich in hemoglobin iron; these red cells are
only one) that underlies the anemia associated with chronic readily replaced by increased erythropoiesis.
inflammation [105, 106]. LPS, as well as the proinflammatory
cytokines interleukin 6 (IL‐6), IL‐1, IL‐22, and interferon‐α
are all able to activate hepcidin expression through a pathway
that involves JAK–STAT signaling and a binding site for the
transcription factor STAT3 in the HAMP promoter [100, 107–
THE DISCOVERY OF HFE, THE FIRST
109]. However, the stimulation of HAMP expression in IDENTIFIED CAUSE OF GENETIC
response to inflammation also requires a functioning BMP/ HEMOCHROMATOSIS
SMAD pathway as IL‐6 cannot enhance hepcidin expression
in Smad4‐knockout mice [76] or in the absence of the type I Although early studies demonstrated that the commonest form of
BMPR ALK3 [110]. genetic hemochromatosis was caused by a mutation on chromo-
A final point should be made on the cellular location of the some 6 based on linkage to the HLA region [114], the affected
hepcidin regulatory machinery. Most work has focused on the gene, HFE, was not identified until 1996, in part because the dis-
hepatocyte as this in the principal site of hepcidin synthesis. ease gene resided in a region of the chromosome that was not
HFE, TFR2, and HJV are also more strongly expressed in hepat- prone to the informative recombination events that facilitate
ocytes than in other types of hepatic cells (such as Kupffer cells, positional cloning [85]. Initially, a single mutation in the coding
endothelial cells and stellate cells). However, another key regu- region of HFE that results in substitution of a tyrosine for a
latory component, BMP6, is synthesized predominantly by cysteine at codon 282 (C282Y) was recognized, and a PCR test
endothelial cells, and to a lesser extent by Kupffer cells and for the C282Y mutation confirmed that it was widespread in pop-
hepatic stellate cells [111, 112]. This indicates that there is ulations of Northern European descent [115]. HFE did not belong
important intrahepatic crosstalk between different cell types in to a known gene family, and initial insight into its potential func-
the liver, adding another layer of complexity to the hepcidin tion came from the observation that it could bind to TFR1 [116]
regulatory network. and interfere with the binding of diferric transferrin [116a].
222 THE LIVER:  TFR2 AND HEMOCHROMATOSIS

Figure 19.6  A model for systemic iron homeostasis. Serum signals reflecting iron stores, erythropoiesis, and inflammation are transduced by
receptors, and information converges on the hepcidin promoter, which determines how much hepcidin is synthesized. Hepcidin binds to ferroportin
on the membrane surface of macrophages, enterocytes, and other target cells, leading to reduced ferroportin expression, and thereby enhancing iron
flow into the plasma. Concomitantly, more iron is sequestered in macrophages and other body cells and duodenal iron absorption is reduced.

However, an animal model indicated that HFE does not exert its may be the most cost‐effective approach to discovering patients
effects on phenotype directly through binding to TFR1 [93], and with HFE‐related hemochromatosis, and serial serum ferritin
subsequent studies showed that HFE functions predominantly as measurements can be used to monitor the iron load [117, 122].
a regulator of hepcidin expression [47]. As noted above, it may do
this in conjunction with TFR2, which is highly expressed in the
liver where systemic iron sensing occurs [28, 95].
The presence of homozygosity for the C282Y mutation in TFR2 AND HEMOCHROMATOSIS
HFE predisposes to iron loading, but does not mandate it. Many
C282Y homozygotes have little evidence of iron overload and Initially discovered as a second TFR with 45% homology to
few symptoms [117]. One large study concluded that overt TFR1, TFR2 differs in that it is mainly expressed in the liver,
hemochromatosis occurred in 28% of male homozygotes for the with some expression in the brain, heart, and spleen, but little
C282Y mutation, but only 1% of female homozygotes [118]. expression in other tissues [16]. Soon after its initial discovery,
Both environmental and genetic factors can play a role in ame- mutations in TFR2 were found to cause hemochromatosis simi-
liorating the iron‐loading phenotype. Environmental factors, lar to HFE hemochromatosis in both the histological pattern of
such as low dietary iron intake, physiological and pathological iron overload (parenchymal loading with sparing of macrophages)
blood loss, and blood donation, are likely to be most important and in disease severity. TFR2‐associated hemochromatosis is
in limiting iron accumulation. However, there are also genetic more severe than HFE‐related disease, but not as severe as
modifiers that can influence whether HFE mutations will cause juvenile hemochromatosis. The disease usually presents in
disease. Examples of these include mutations in the HAMP adults [123, 124], although it sometimes presents in juveniles
[119], DCYTB [120], and GNPAT [121] genes. [125]. Recapitulation of the essential features of the disease in a
Numerous tests have been used to diagnose hemochromato- mouse model confirmed that TFR2 was another cause of
sis, including TF saturation, serum ferritin level, liver biopsy classical hemochromatosis [126]. Interestingly, TFR2 levels
with evidence of parenchymal iron overload and cirrhosis, and increase with binding of TF, due to decreased degradation of
genetic testing for the C282Y mutation. However, a single TFR2 in lysosomes, and binding of TF results in more recycling
serum ferritin determination, with a follow‐up mutation test, of TFR2 [28].
19:  The Central Role of the Liver in Iron Storage and Regulation of Systemic Iron Homeostasis 223

CAUSES OF JUVENILE MUTATIONS OF THE IRON EXPORTER


HEMOCHROMATOSIS FERROPORTIN CAUSE A GROUP
OF DISTINCT IRON OVERLOAD
Among hemochromatosis patients, the severity of disease can
SYNDROMES
vary substantially. For many years, investigators recognized a
distinct type of hemochromatosis in which severe iron overload
Mutations in the gene encoding the iron export protein FPN
and organ damage occurred before patients reached the age of
(SLC40A1) can also lead to iron loading [135]. Patients with
30 (i.e. juvenile hemochromatosis). This unusually severe clini-
mutations in SLC40A1 fall into two classes, with distinct pheno-
cal presentation correlates with mutations in genes that encode
types depending on the particular mutation they possess [136].
proteins that are central to body iron homeostasis: the HAMP
One type of mutation is autosomal dominant and the resulting
gene that encodes hepcidin and the HJV gene that encodes
syndrome is referred to by some experts as ferroportin disease.
hemojuvelin.
This type of mutation leads to reduced levels of FPN on the
plasma membrane or reduced iron transport capacity. Reduced
Hepcidin mutations cause severe early‐onset iron export means that less iron enters the plasma and that iron
hemochromatosis accumulates in tissue macrophages, as internal iron recycling is
disrupted. Thus the characteristic of this type of ferroportin-
As described earlier, hepcidin is now acknowledged to be the linked disease is reduced plasma iron in the face of reticuloen-
most important regulator of systemic iron homeostasis through dothelial cell iron accumulation. In the liver, iron accumulation
its capacity to regulate iron release from macrophages, duode- is seen predominantly in Kupffer cells, which is distinct from
nal enterocytes, and other cells, and it is not surprising that the periportal hepatocyte‐dominated distribution observed in
mutations in hepcidin itself cause a very severe iron‐loading the early stages of HFE‐related hemochromatosis. The reduced
phenotype [127]. plasma iron can also mean that a mild anemia tends to develop,
especially with phlebotomy therapy. This in turn can stimulate
iron absorption and increase the body iron load. In ferroportin
Hemojuvelin mutations cause severe
disease, iron absorption is less efficient due to functional FPN
early‐onset hemochromatosis deficiency, but there still appears to be a net accretion of iron by
Using positional cloning approaches to investigate families the body. The second type of ferroportin mutation is autosomal
with juvenile hemochromatosis, a gene not previously recessive and is referred to by some experts as ferroportin‐asso-
­recognized as important in iron homeostasis was discovered ciated hemochromatosis. The condition is characterized by mis-
to harbor mutations that cause disease [75, 128]. The gene sense mutations in SLC401 that generate a FPN protein that has
product was given the name “hemojuvelin” because of its reduced responsiveness to hepcidin. In this situation, the iron
causal role in juvenile hemochromatosis, in which patients loading is quite similar to that associated with disruption of
typically develop hypogonadotrophic hypogonadism, hepatic HFE, TFR2, hemojuvelin, and hepcidin, with elevated transfer-
fibrosis or cirrhosis, and cardiomyopathy before age 30. By rin saturation and periportal deposition of iron in the hepato-
sequence homology, HJV belongs in the family of repulsive cytes of the liver. The failure of FPN to respond appropriately to
guidance molecules (RGMs), proteins that bind to pairs of hepcidin means that the body’s normal feedback mechanism for
type I and type II serine/threonine kinase receptors that are limiting iron loading is dysfunctional.
important in transducing signals from proteins in the
transforming growth factor β (TGF‐β) superfamily. Upon
­
ligand binding, these type I and type II receptor pairs activate
the SMAD proteins, which translocate to the nucleus to mod-
SUPPRESSION OF HEPCIDIN EXPRESSION
ulate transcription of various target genes after they have BY ERYTHROFERRONE IS IMPORTANT
undergone receptor‐dependent phosphorylation. BMPs are IN IRON‐LOADING ANEMIAS
signaling molecules in the TGF‐β superfamily that interact
with a specific set of type I and type II receptor pairs known Patients with anemias characterized by ineffective erythropoiesis,
as BMP receptors (reviewed in [129]). For some receptors in such as thalassemia major and intermedia syndromes, develop
the BMP signaling pathway, RGMs bind to the receptor– systemic iron overload in the absence of blood transfusions due
ligand complex and potentiate signaling [130]. It appears that to enhanced intestinal iron absorption [137]. The magnitude of
HJV, which is required for the increase of hepcidin transcrip- iron overload is independent of the degree of anemia [138] and
tion that occurs in response to iron loading [131], is a co‐ the iron loading is predominantly parenchymal, similar to
receptor that facilitates BMP signaling in hepatocytes [132]. hereditary hemochromatosis due to mutations in HFE, TFR2,
HJV is linked to membranes through a GPI membrane link- HJV, and HAMP. These observations suggested that there is a
age, and it specifically facilitates signaling of BMP2 and common pathophysiologic pathway for the iron loading in ane-
BMP4 ligands [133]. Mouse models in which the hemojuvelin mias characterized by ineffective erythropoiesis and the iron
gene is deleted (Hjv−/−) develop parenchymal iron overload typ- loading in hereditary hemochromatosis [139]. Indeed, there is
ical of hemochromatosis, but they develop relatively little now ample evidence that deficiency of hepcidin with respect to
organ damage, unlike humans [131, 134], an observation that the body’s iron burden underlies the iron loading seen in both
remains unexplained. hereditary hemochromatosis and anemias characterized by
224 THE LIVER:  ACKNOWLEDGMENTS

ineffective erythropoiesis [140]. Investigators recently discov- THE LIVER IS IMPORTANT IN CLEARING
ered that erythroferrone, a factor secreted by erythroid progeni- REACTIVE HEME FROM CIRCULATION
tors in the bone marrow, is an important mediator of the
suppression of hepatic hepcidin expression that is observed with Hepatocytes are the main site of synthesis of hemopexin, an
increased erythropoiesis, especially ineffective erythropoiesis abundant serum protein that binds free heme with high affinity
[141, 142]. As noted earlier, BMP signaling through phospho- [150]. The structure of hemopexin enables it to sequester heme
rylation of SMAD proteins is a central pathway to regulate hep- in a nonreactive form [151] and heme–hemopexin complexes
cidin transcription [132, 143]. Erythropoietin robustly induces are then metabolized by cells that have hemopexin receptors,
bone marrow erythroferrone mRNA, and erythroferrone appears including hepatocytes, macrophages, neurons, and syncytio-
to act through hepatocyte SMAD1 and SMAD5 signaling to trophoblasts [152]. Increased hepatic uptake of hemopexin trig-
suppress hepcidin production [144] but precisely how erythro- gers an increase in hepatic heme oxygenase‐1 expression, which
ferrone interacts with hepatocytes to modulate SMAD signaling enables hepatocytes to fully metabolize the newly absorbed
is unknown. Other recent evidence indicates that a reduction in heme. The hemopexin system is a major defense against tissue
the level of diferric TF can contribute to the reduction in HAMP injury by free heme, which is important in the pathogenesis of
expression associated with stimulated erythropoiesis [103]. ischemia–reperfusion and traumatic injuries. Mice engineered
to lack hemopexin are more prone to developing renal failure
ELEVATED HEPCIDIN EXPRESSION after hemolysis than wild‐type mice, indicating that hemopexin
is important for protecting animals from the toxic effects of free
EXPLAINS THE ANEMIA OF CHRONIC
heme in the circulation [153].
DISEASE Haptoglobin is a tetrameric protein expressed by the liver that
combines with free plasma hemoglobin and protects tissues,
The anemia of chronic disease, better termed the anemia of particularly the kidney, from hemoglobin‐dependent damage
chronic inflammation, is characterized by decreased production [154]. Hemoglobin–haptoglobin complexes are cleared from
of erythrocytes by the bone marrow without underlying nutrient the circulation by liver and spleen macrophages, which express
or hormonal deficiency and with no intrinsic abnormality of the the hemoglobin–haptoglobin receptor CD163 [155]. The liver
bone marrow. This hypoproliferative anemia is marked by the protects the other organs against damage from hemolysis and
presence of one of several chronic underlying inflammatory dis- trauma by synthesizing and secreting hemopexin and haptoglobin,
orders, which can be of infectious, malignant, inflammatory, or and also by taking up heme–hemopexin and hemoglobin–
traumatic origin. Typically the anemia is mild or moderate in haptoglobin complexes, and catabolizing the heme it acquires
character, the mean corpuscular volume is mildly reduced, the through these pathways using heme oxygenase.
serum concentrations of iron and transferrin are decreased, and
the serum ferritin concentration is in the upper end of the normal
range or frankly elevated. Duodenal iron absorption is reduced
and storage of iron in macrophages is increased. The changes in
iron metabolism in the anemia of chronic inflammation are likely
HEPATIC IRON OVERLOAD AND
attributable to increased secretion of hepcidin by hepatocytes. THE RELATIONSHIP OF IRON
Several experiments indicate that exogenous delivery of hep- OVERLOAD TO ALCOHOLIC
cidin or overexpression of hepcidin can cause chronic anemia. LIVER DISEASE
The injection of hepcidin in mice rapidly leads to a prolonged
hypoferremia [145]. Mice that overexpress hepcidin under the A third or more of alcoholics develop increased amounts of
control of a liver‐specific promoter depend on parenteral iron hepatic iron compared to controls [156, 157]. In rural Africa,
injections after birth, suggesting that hepcidin is a negative reg- there is a strong association among the consumption of a tradi-
ulator of iron transport in the small intestine and that it induces tional fermented beverage with high ionic iron concentration,
iron retention in macrophages that normally recycle iron from increased body iron stores, and liver toxicity and cirrhosis
senescent erythrocytes [146]. Chronic overexpression of human [158–162]. In experimental rats, dietary iron supplementation
hepcidin in mice leads to hypoferremia, increased hepatic iron, exacerbates alcohol‐induced hepatocyte damage and promotes
and anemia [147]. Patients with large hepatic adenomas that liver fibrogenesis [163]. Similarly, the presence of elevated
produce inappropriately high levels of hepcidin mRNA have iron stores accentuates the hepatic toxicity of alcohol in the
iron‐refractory anemia similar to that observed in anemia of setting of HFE hemochromatosis in Caucasians [164]. The
chronic disease, and this anemia resolves spontaneously after increased iron absorption associated with alcohol ingestion
adenoma resection or liver transplantation [148]. may be related to suppression of hepcidin production by
In addition to hepatocytes, monocytes also produce low lev- hepatocytes [165–167].
els of hepcidin and such autocrine production may contribute to
macrophage iron loading and the anemia of chronic inflamma-
tion [149]. Changes in iron metabolism probably do not fully
explain the anemia of chronic disease. Additional factors may ACKNOWLEDGMENTS
include suppression of erythroid precursors by inflammatory
cytokines and some limitation to the production of or response We thank Helge Uhrigshardt for creating the ferritin structural
to erythropoietin. representation.
19:  The Central Role of the Liver in Iron Storage and Regulation of Systemic Iron Homeostasis 225

REFERENCES 28. Johnson, M.B., Chen, J., Murchison, N., Green, F.A., and Enns, C.A.
Transferrin receptor 2: evidence for ligand‐induced stabilization and redirec-
tion to a recycling pathway. Mol Biol Cell, 2007;18(3):743–54.
1. Pritchard, J.A, and Mason, R.A. Iron stores of normal adults and replenish-
29. Zimelman, A.P., Zimmerman, H.J., McLean, R., and Weintraub, L.R. Effect
ment with oral iron therapy. JAMA, 1964;190:897–901.
of iron saturation of transferrin on hepatic iron uptake: an in vitro study.
2. Haskins, D., Stevens, A.R., Jr, Finch, S., and Finch, C.A. Iron metabolism;
Gastroenterology, 1977;72(1):129–31.
iron stores in man as measured by phlebotomy. J Clin Invest, 1952;
30. Craven, C.M., Alexander, J., Eldridge, M. et al. Tissue distribution and clear-
31(6):543–7.
ance kinetics of non‐transferrin‐bound iron in the hypotransferrinemic
3. Walters, G.O., Miller, F.M., and Worwood, M. Serum ferritin concentration
mouse: a rodent model for hemochromatosis. Proc Natl Acad Sci U S A,
and iron stores in normal subjects. J Clin Pathol, 1973;26(10):770–2.
1987;84(10):3457–61.
4. Olsson, K.S. Iron stores in normal men and male blood donors. As measured
31. Brissot, P., Wright, T.L., Ma, W.L., and Weisiger, R.A. Efficient clearance of
by desferrioxamine and quantitative phlebotomy. Acta Med Scand, 1972;
non‐transferrin‐bound iron by rat liver. Implications for hepatic iron loading
192(5):401–7.
in iron overload states. J Clin Invest, 1985;76(4):1463–70.
5. Balcerzak, S.P., Westerman, M.P., Heinle, E.W., and Taylor, F.H. Measurement
32. Wright, T.L., Brissot, P., Ma, W.L., and Weisiger, R.A. Characterization of
of iron stores using deferoxamine. Ann Intern Med, 1968;68(3):518–25.
non‐transferrin‐bound iron clearance by rat liver. J Biol Chem,
6. Harrison, P.M. and Arosio, P. The ferritins: molecular properties, iron storage
1986;261(23):10909–14.
function and cellular regulation. Biochim Biophys Acta, 1996;1275(3):
33. Bernstein, S.E. Hereditary hypotransferrinemia with hemosiderosis, a
161–203.
murine disorder resembling human atransferrinemia. J Lab Clin Med,
7. Theil, E.C. Ferritins. Annu Rev Biochem, 1987;56:289–315.
1987;110(6):690–705.
8. De Domenico, I., Vaughn, M.B., Li, L. et al. Ferroportin‐mediated mobiliza-
34. Hayashi, A., Wada, Y., Suzuki, T., and Shimizu, A. Studies on familial
tion of ferritin iron precedes ferritin degradation by the proteasome. EMBO
hypotransferrinemia: unique clinical course and molecular pathology. Am J
J, 2006;25(22):5396–404.
Hum Genet, 1993;53(1):201–13.
9. Mancias, J.D., Pontano Vaites, L., Nissim, S. et  al. Ferritinophagy via
35. Aisen, P., Leibman, A., and Zweier, J. Stoichiometric and site characteristics
NCOA4 is required for erythropoiesis and is regulated by iron dependent
of the binding of iron to human transferrin. J Biol Chem,
HERC2‐mediated proteolysis. Elife, 2015;4.
1978;253(6):1930–1937.
10. Ha, Y., Shi, D., Small, G.W., Theil, E.C., and Allewell, N.M. Crystal struc-
36. Breuer, W., Ronson, A., Slotki, I.N. et al. The assessment of serum nontrans-
ture of bullfrog M ferritin at 2.8 A resolution: analysis of subunit interactions
ferrin‐bound iron in chelation therapy and iron supplementation. Blood,
and the binuclear metal center. J Biol Inorg Chem, 1999;4(3):243–56.
2000;95(9):2975–82.
11. Wally, J. and Buchanan, S.K. A structural comparison of human serum trans-
37. Hider, R.C. Nature of nontransferrin‐bound iron. Eur J Clin Invest,
ferrin and human lactoferrin. Biometals, 2007;20(3–4):249–62.
2002;32(Suppl 1):50–4.
12. Chua, A.C., Graham, R.M., Trinder, D., and Olynyk, J.K. The regulation of
38. Baker, E., Baker, S.M., and Morgan, E.H. Characterisation of non‐transfer-
cellular iron metabolism. Crit Rev Clin Lab Sci, 2007;44(5–6):413–59.
rin‐bound iron (ferric citrate) uptake by rat hepatocytes in culture. Biochim
13. Ohgami, R.S., Campagna, D.R., McDonald, A., and Fleming, M.D. The
Biophys Acta, 1998;1380(1):21–30.
Steap proteins are metalloreductases. Blood, 2006;108(4):1388–94.
39. Trinder, D., Oates, P.S., Thomas, C., Sadleir, J., and Morgan, E.H.
14. Mims, M.P. and Prchal, J.T. Divalent metal transporter 1. Hematology,
Localisation of divalent metal transporter 1 (DMT1) to the microvillus mem-
2005;10(4):339–45.
brane of rat duodenal enterocytes in iron deficiency, but to hepatocytes in
15. Shaw, G.C., Cope, J.J., Li, L. et al. Mitoferrin is essential for erythroid iron
iron overload. Gut, 2000;46(2):270–6.
assimilation. Nature, 2006;440(7080):96–100.
40. Mims, M.P., Guan, Y., Pospisilova, D. et al. Identification of a human muta-
16. Kawabata, H., Yang, R., Hirama, T. et al. Molecular cloning of transferrin
tion of DMT1 in a patient with microcytic anemia and iron overload. Blood,
receptor 2. A new member of the transferrin receptor‐like family. J Biol
2005;105(3):1337–42.
Chem, 1999;274(30):20826–32.
41. Thompson, K., Molina, R.M., Brain, J.D., and Wessling‐Resnick, M.
17. Chloupkova, M., Zhang, A.S., and Enns, C.A. Stoichiometries of transferrin
Belgrade rats display liver iron loading. J Nutr, 2006;136(12):3010–14.
receptors 1 and 2 in human liver. Blood Cells Mol Dis, 2010;44(1): 28–33.
42. Aydemir, T.B. and Cousins, R.J. The multiple faces of the metal transporter
18. Kleven, M.D., Jue, S., and Enns, C.A. Transferrin receptors TfR1 and TfR2 bind
ZIP14 (SLC39A14). J Nutr, 2018;148(2):174–84.
transferrin through differing mechanisms. Biochemistry, 2018;57(9):1552–9.
43. Jenkitkasemwong, S., Wang, C.Y., Coffey, R. et al. SLC39A14 is required
19. West, A.P., Jr., Bennett, M.J., Sellers, V.M. et al. Comparison of the interactions
for the development of hepatocellular iron overload in murine models of
of transferrin receptor and transferrin receptor 2 with transferrin and the heredi-
hereditary hemochromatosis. Cell Metab, 2015;22(1):138–50.
tary hemochromatosis protein HFE. J Biol Chem, 2000; 275(49):38135–8.
44. Tripathi, A.K., Haldar, S., Qian, J. et al. Prion protein functions as a ferrire-
20. Huebers, H.A. and Finch, C.A. The physiology of transferrin and transferrin
ductase partner for ZIP14 and DMT1. Free Radic Biol Med,
receptors. Physiol Rev, 1987;67(2):520–82.
2015;84:322–30.
21. Herbison, C.E., Thorstensen, K., Chua, A.C. et  al. The role of transferrin
45. Randell, E.W., Parkes, J.G., Olivieri, N.F., and Templeton, D.M. Uptake of
receptor 1 and 2 in transferrin‐bound iron uptake in human hepatoma cells.
non‐transferrin‐bound iron by both reductive and nonreductive processes is
Am J Physiol Cell Physiol, 2009;297(6):C1567–75.
modulated by intracellular iron. J Biol Chem, 1994;269(23):16046–53.
22. Chen, J. and Enns, C.A. Hereditary hemochromatosis and transferrin recep-
46. Zhao, N., Zhang, A.S., Worthen, C., Knutson, M.D., and Enns, C.A. An iron‐
tor 2. Biochim Biophys Acta, 2012;1820(3):256–63.
regulated and glycosylation‐dependent proteasomal degradation pathway for
23. Tacchini, L., Bianchi, L., Bernelli‐Zazzera, A., and Cairo, G. Transferrin
the plasma membrane metal transporter ZIP14. Proc Natl Acad Sci U S A,
receptor induction by hypoxia. HIF‐1‐mediated transcriptional activation
2014;111(25):9175–80.
and cell‐specific post‐transcriptional regulation. J Biol Chem,
47. Brissot, P., Pietrangelo, A., Adams, P.C. et al. Haemochromatosis. Nat Rev
1999;274(34):24142–6.
Dis Primers, 2018;4:18016.
24. Lok, C.N. and Ponka, P. Identification of a hypoxia response element in the
48. Naito, M., Hasegawa, G., Ebe, Y., and Yamamoto, T. Differentiation and
transferrin receptor gene. J Biol Chem, 1999;274(34):24147–52.
function of Kupffer cells. Med Electron Microsc, 2004;37(1):16–28.
25. Addess, K.J., Basilion, J.P., Klausner, R.D., Rouault, T.A., and Pardi, A.
49. McKie, A.T., Marciani, P., Rolfs, A. et al. A novel duodenal iron‐regulated
Structure and dynamics of the iron responsive element RNA: implications
transporter, IREG1, implicated in the basolateral transfer of iron to the circu-
for binding of the RNA by iron regulatory binding proteins. J Mol Biol,
lation. Mol Cell, 2000;5(2):299–309.
1997;274(1):72–83.
50. Donovan, A., Brownlie, A., Zhou, Y. et  al. Positional cloning of zebrafish
26. Ghosh, M.C., Zhang, D.L., and Rouault, T.A. Iron misregulation and neuro-
ferroportin1 identifies a conserved vertebrate iron exporter. Nature,
degenerative disease in mouse models that lack iron regulatory proteins.
2000;403(6771):776–81.
Neurobiol Dis, 2015;81:66–75.
51. Abboud, S. and Haile, D.J. A novel mammalian iron‐regulated protein involved
27. Fleming, R.E., Migas, M.C., Holden, C.C. et al. Transferrin receptor 2: con-
in intracellular iron metabolism. J Biol Chem, 2000;275(26):19906–12.
tinued expression in mouse liver in the face of iron overload and in hereditary
52. Knutson, M. and Wessling‐Resnick, M. Iron metabolism in the reticuloen-
hemochromatosis. Proc Natl Acad Sci U S A, 2000;97(5):2214–19.
dothelial system. Crit Rev Biochem Mol Biol, 2003;38(1):61–88.
226 THE LIVER:  REFERENCES

53. Zhang, A.S., Xiong, S., Tsukamoto, H., and Enns, C.A. Localization of iron 78. Meynard, D., Kautz, L., Darnaud, V. et al. Lack of the bone morphogenetic
metabolism‐related mRNAs in rat liver indicate that HFE is expressed pre- protein BMP6 induces massive iron overload. Nat Genet, 2009;
dominantly in hepatocytes. Blood, 2004;103(4):1509–14. 41(4):478–81.
54. Hellman, N.E. and Gitlin, J.D. Ceruloplasmin metabolism and function. 79. Canali, S., Wang, C.Y., Zumbrennen‐Bullough, K.B., Bayer, A., and Babitt,
Annu Rev Nutr, 2002;22:439–58. J.L. Bone morphogenetic protein 2 controls iron homeostasis in mice inde-
55. Harris, Z.L., Durley, A.P., Man, T.K., and Gitlin, J.D. Targeted gene disrup- pendent of Bmp6. Am J Hematol, 2017;92(11):1204–13.
tion reveals an essential role for ceruloplasmin in cellular iron efflux. Proc 80. Du, X., She, E., Gelbart, T. et al. The serine protease TMPRSS6 is required
Natl Acad Sci U S A, 1999;96(19):10812–17. to sense iron deficiency. Science, 2008;320(5879):1088–92.
56. Xu, X., Pin, S., Gathinji, M., Fuchs, R., and Harris, Z.L. Aceruloplasminemia: 81. Finberg, K.E., Heeney, M.M., Campagna, D.R. et  al. Mutations in
an inherited neurodegenerative disease with impairment of iron homeostasis. TMPRSS6 cause iron‐refractory iron deficiency anemia (IRIDA). Nat
Ann N Y Acad Sci, 2004;1012:299–305. Genet, 2008;40(5):569–71.
57. Fuqua, B.K., Lu, Y., Darshan, D. et al. Deletion of the multicopper ferroxi- 82. Maxson, J.E., Chen, J., Enns, C.A., and Zhang, A.S. Matriptase‐2‐ and propro-
dases hephaestin and ceruloplasmin disrupts systemic iron distribution with- tein convertase‐cleaved forms of hemojuvelin have different roles in the down‐
out affecting intestinal iron absorption. Cell Mol Gastroenterol Hepatol, regulation of hepcidin expression. J Biol Chem, 2010;285(50):39021–8.
2018. https://doi.org/10.1016/j.jcmgh.2018.06.006 83. Silvestri, L., Pagani, A., Nai, A. et  al. The serine protease matriptase‐2
58. De Domenico, I., Ward, D.M., di Patti, M.C. et  al. Ferroxidase activity is (TMPRSS6) inhibits hepcidin activation by cleaving membrane hemojuve-
required for the stability of cell surface ferroportin in cells expressing GPI‐ lin. Cell Metab, 2008;8(6):502–11.
ceruloplasmin. EMBO J, 2007;26(12):2823–31. 84. Wahedi, M., Wortham, A.M., Kleven, M.D. et al. Matriptase‐2 suppresses
59. Ji, C., Steimle, B.L., Bailey, D.K., and Kosman, D.J. The ferroxidase hepcidin expression by cleaving multiple components of the hepcidin
hephaestin but not amyloid precursor protein is required for ferroportin‐sup- induction pathway. J Biol Chem, 2017;292(44):18354–71.
ported iron efflux in primary hippocampal neurons. Cell Mol NeuroBiol, 85. Feder, J.N., Gnirke, A., Thomas, W. et al. A novel MHC class I‐like gene is
2018;38(4):941–54. mutated in patients with hereditary haemochromatosis. Nat Genet,
60. Drakesmith, H., Nemeth, E., and Ganz, T. Ironing out ferroportin. Cell 1996;13(4):399–408.
Metab, 2015;22(5):777–87. 86. Kautz, L., Meynard, D., Besson‐Fournier, C. et al. BMP/Smad signaling is
61. Nemeth, E., Preza, G.C., Jung, C.L. et  al. The N‐terminus of hepcidin is not enhanced in Hfe‐deficient mice despite increased Bmp6 expression.
essential for its interaction with ferroportin: structure‐function study. Blood, Blood, 2009;114(12):2515–20.
2006;107(1):328–33. 87. Ryan, J.D., Ryan, E., Fabre, A., Lawless, M.W., and Crowe, J. Defective
62. Qiao, B., Sugianto, P., Fung, E. et al. Hepcidin‐induced endocytosis of fer- bone morphogenic protein signaling underlies hepcidin deficiency in HFE
roportin is dependent on ferroportin ubiquitination. Cell Metab, hereditary hemochromatosis. Hepatology, 2010;52(4):1266–73.
2012;15(6):918–24. 88. Bridle, K.R., Frazer, D.M., Wilkins, S.J. et al. Disrupted hepcidin regula-
63. Preza, G.C., Pinon, R., Ganz, T., and Nemeth, E. Cellular catabolism of the tion in HFE‐associated haemochromatosis and the liver as a regulator of
iron‐regulatory peptide hormone hepcidin. PLoS One, 2013;8(3):e58934. body iron homoeostasis. Lancet, 2003;361(9358):669–73.
64. Clark, R.J., Tan, C.C., Preza, G.C. et al. Understanding the structure/activity 89. Nemeth, E., Roetto, A., Garozzo, G., Ganz, T., and Camaschella, C.
relationships of the iron regulatory peptide hepcidin. Chem Biol, Hepcidin is decreased in TFR2 hemochromatosis. Blood, 2005;105(4):
2011;18:336–43. 1803–6.
65. Frazer, D.M. and Anderson, G.J. The orchestration of body iron intake: how 90. Corradini, E., Rozier, M., Meynard, D. et al. Iron regulation of hepcidin
and where do enterocytes receive their cues? Blood Cells Mol Dis, despite attenuated Smad1,5,8 signaling in mice without transferrin receptor
2003;30(3):288–97. 2 or Hfe. Gastroenterology, 2011;141(5):1907–14.
66. Nemeth, E., Tuttle, M.S., Powelson, J. et al. Hepcidin regulates cellular iron 91. West, A.P., Jr., Giannetti, A.M., Herr, A.B. et al. Mutational analysis of the
efflux by binding to ferroportin and inducing its internalization. Science, transferrin receptor reveals overlapping HFE and transferrin binding sites.
2004;306(5704):2090–3. J Mol Biol, 2001;313(2):385–97.
67. Delaby, C., Pilard, N., Goncalves, A.S., Beaumont, C., and Canonne‐ 92. Giannetti, A.M. and Bjorkman, P.J. HFE and transferrin directly compete
Hergaux, F. Presence of the iron exporter ferroportin at the plasma mem- for transferrin receptor in solution and at the cell surface. J Biol Chem,
brane of macrophages is enhanced by iron loading and down‐regulated by 2004;279(24):25866–75.
hepcidin. Blood, 2005;106(12):3979–84. 93. Schmidt, P.J., Toran, P.T., Giannetti, A.M., Bjorkman, P.J., and Andrews,
68. Gulec, S., Anderson, G.J., and Collins, J.F. Mechanistic and regulatory N.C. The transferrin receptor modulates Hfe‐dependent regulation of hep-
aspects of intestinal iron absorption. Am J Physiol Gastrointest Liver cidin expression. Cell Metab, 2008;7(3):205–14.
Physiol, 2014;307(4):G397–409. 94. Goswami, T. and Andrews, N.C. Hereditary hemochromatosis protein,
69. Mastrogiannaki, M., Matak, P., Keith, B. et  al. HIF‐2alpha, but not HFE, interaction with transferrin receptor 2 suggests a molecular mecha-
HIF‐1alpha, promotes iron absorption in mice. J Clin Invest, nism for mammalian iron sensing. J Biol Chem, 2006;281(39):28494–8.
2009;119(5):1159–66. 95. Waheed, A., Britton, R.S., Grubb, J.H., Sly, W.S., and Fleming, R.E. HFE
70. Chaston, T., Chung, B., Mascarenhas, M. et  al. Evidence for differential association with transferrin receptor 2 increases cellular uptake of transfer-
effects of hepcidin in macrophages and intestinal epithelial cells. Gut, rin‐bound iron. Arch Biochem Biophys, 2008;474(1):193–7.
2008;57(3):374–82. 96. Zhao, N., Zhang, A.S., and Enns, C.A. Iron regulation by hepcidin. J Clin
71. Oates, P.S. The role of hepcidin and ferroportin in iron absorption. Histol Invest, 2013;123(6):2337–43.
Histopathol, 2007;22(7):791–804. 97. Wallace, D.F., Summerville, L., Crampton, E.M. et al. Combined deletion
72. Laftah, A.H., Ramesh, B., Simpson, R.J. et al. Effect of hepcidin on intesti- of Hfe and transferrin receptor 2 in mice leads to marked dysregulation of
nal iron absorption in mice. Blood, 2004;103(10):3940–4. hepcidin and iron overload. Hepatology, 2009;50(6):1992–2000.
73. Vujic Spasic, M., Kiss, J., Herrmann, T. et  al. Hfe acts in hepatocytes to 98. Wu, X.G., Wang, Y., Wu, Q. et al. HFE interacts with the BMP type I recep-
prevent hemochromatosis. Cell Metab, 2008;7(2):173–8. tor ALK3 to regulate hepcidin expression. Blood, 2014;124(8):1335–43.
74. Parrow, N.L. and Fleming, R.E. Bone morphogenetic proteins as regulators 99. Wilkins, S.J., Frazer, D.M., Millard, K.N., McLaren, G.D., and Anderson,
of iron metabolism. Annu Rev Nutr, 2014;34:77–94. G.J. Iron metabolism in the hemoglobin‐deficit mouse: correlation of difer-
75. Papanikolaou, G., Samuels, M.E., Ludwig, E.H. et al. Mutations in HFE2 ric transferrin with hepcidin expression. Blood, 2006;107(4):1659–64.
cause iron overload in chromosome 1q‐linked juvenile hemochromatosis. 100. Lee, P., Peng, H., Gelbart, T., Wang, L., and Beutler, E. Regulation of hep-
Nat Genet, 2004;36(1):77–82. cidin transcription by interleukin‐1 and interleukin‐6. Proc Natl Acad Sci U
76. Wang, R.H., Li, C., Xu, X. et  al. A role of SMAD4 in iron metabolism S A, 2005;102(6):1906–10.
through the positive regulation of hepcidin expression. Cell Metab, 101. Darshan, D. and Anderson, G.J. Liver‐gut axis in the regulation of iron
2005;2(6):399–409. homeostasis. World J Gastroenterol, 2007;13(35):4737–45.
77. Andriopoulos, B., Jr., Corradini, E., Xia, Y. et al. BMP6 is a key endogenous 102. Anderson, G.J., Darshan, D., Wilkins, S.J., and Frazer, D.M. Regulation of
regulator of hepcidin expression and iron metabolism. Nat Genet, systemic iron homeostasis: how the body responds to changes in iron
2009;41(4):482–7. demand. Biometals, 2007;20(3–4):665–74.
19:  The Central Role of the Liver in Iron Storage and Regulation of Systemic Iron Homeostasis 227

103. Mirciov, C.S.G., Wilkins, S.J., Hung, G.C.C. et al. Circulating iron levels 128. Lanzara, C., Roetto, A., Daraio, F. et  al. Spectrum of hemojuvelin gene
influence the regulation of hepcidin following stimulated erythropoiesis. mutations in 1q‐linked juvenile hemochromatosis. Blood, 2004;103(11):
Haematologica, 2018;103(10):1616–26. 4317–21.
104. Nicolas, G., Chauvet, C., Viatte, L. et al. The gene encoding the iron regu- 129. Herpin, A. and Cunningham, C. Cross‐talk between the bone morphoge-
latory peptide hepcidin is regulated by anemia, hypoxia, and inflamma- netic protein pathway and other major signaling pathways results in tightly
tion. J Clin Invest, 2002;110(7):1037–44. regulated cell‐specific outcomes. FEBS J, 2007;274(12):2977–85.
105. Ganz, T. and Nemeth, E. Iron homeostasis in host defence and inflamma- 130. Halbrooks, P.J., Ding, R., Wozney, J.M., and Bain, G. Role of RGM core-
tion. Nat Rev Immunol, 2015;15(8):500–10. ceptors in bone morphogenetic protein signaling. J Mol Signal, 2007;2:4.
106. Kim, A., Fung, E., Parikh, S.G. et al. A mouse model of anemia of inflam- 131. Niederkofler, V., Salie, R., and Arber, S. Hemojuvelin is essential for die-
mation: complex pathogenesis with partial dependence on hepcidin. tary iron sensing, and its mutation leads to severe iron overload. J Clin
Blood, 2014;123(8):1129–36. Invest, 2005;115(8):2180–6.
107. Wrighting, D.M. and Andrews, N.C. Interleukin‐6 induces hepcidin 132. Babitt, J.L., Huang, F.W., Wrighting, D.M. et al. Bone morphogenetic pro-
expression through STAT3. Blood, 2006;108(9):3204–9. tein signaling by hemojuvelin regulates hepcidin expression. Nat Genet,
108. Verga Falzacappa, M.V., Vujic Spasic, M., Kessler, R. et al. STAT3 medi- 2006;38(5):531–9.
ates hepatic hepcidin expression and its inflammatory stimulation. Blood, 133. Lin, L., Valore, E.V., Nemeth, E. et al. Iron transferrin regulates hepcidin
2007;109(1):353–8. synthesis in primary hepatocyte culture through hemojuvelin and BMP2/4.
109. Schmidt, P.J. Regulation of iron metabolism by hepcidin under conditions Blood, 2007;110(6):2182–9.
of inflammation. J Biol Chem, 2015;290(31):18975–83. 134. Huang, F.W., Pinkus, J.L., Pinkus, G.S., Fleming, M.D., and Andrews, N.C. A
110. Mayeur, C., Lohmeyer, L.K., Leyton, P. et al. The type I BMP receptor mouse model of juvenile hemochromatosis. J Clin Invest, 2005;115(8):
Alk3 is required for the induction of hepatic hepcidin gene expression by 2187–91.
interleukin‐6. Blood, 2014;123(14):2261–8. 135. Pietrangelo, A. The ferroportin disease. Blood Cells Mol Dis,
111. Enns, C.A., Ahmed, R., Wang, J. et al. Increased iron loading induces 2004;32(1):131–8.
Bmp6 expression in the non‐parenchymal cells of the liver independent of 136. Pietrangelo, A. Ferroportin disease: pathogenesis, diagnosis and treatment.
the BMP‐signaling pathway. PLoS One. 2013;8(4):e60534. Haematologica, 2017;102(12):1972–84.
112. Canali, S., Zumbrennen‐Bullough, K.B., Core, A.B. et  al. Endothelial 137. Pootrakul, P., Kitcharoen, K., Yansukon, P. et  al. The effect of erythroid
cells produce bone morphogenetic protein 6 required for iron homeostasis hyperplasia on iron balance. Blood, 1988;71(4):1124–9.
in mice. Blood, 2017;129(4):405–14. 138. Cazzola, M., Barosi, G., Bergamaschi, G. et al. Iron loading in congenital
113. Pietrangelo, A. Hereditary hemochromatosis. Biochim Biophys Acta, dyserythropoietic anaemias and congenital sideroblastic anaemias. Br J
2006;1763(7):700–10. Haematol, 1983;54(4):649–54.
114. Simon, M., Bourel, M., Fauchet, R., and Genetet, B. Association of HLA‐ 139. Gordeuk, V.R., McLaren, G.D., and Samowitz, W. Etiologies, conse-
A3 and HLA‐B14 antigens with idiopathic haemochromatosis. Gut, quences, and treatment of iron overload. Crit Rev Clin Lab Sci,
1976;17(5):332–4. 1994;31(2):89–133.
115. Merryweather‐Clarke, A.T., Pointon, J.J., Shearman, J.D., and Robson, 140. Camaschella, C. Iron and hepcidin: a story of recycling and balance.
K.J. Global prevalence of putative haemochromatosis mutations. J Med Hematology Am Soc Hematol Educ Program, 2013;2013:1–8.
Genet, 1997;34(4):275–8. 141. Pak, M., Lopez, M.A., Gabayan, V., Ganz, T., and Rivera, S. Suppression
116. Lebron, J.A., Bennett, M.J., Vaughn, D.E. et al. Crystal structure of the of  hepcidin during anemia requires erythropoietic activity. Blood,
hemochromatosis protein HFE and characterization of its interaction with 2006;108(12):3730–5.
transferrin receptor. Cell, 1998;93(1):111–23. 142. Kautz, L., Jung, G., Valore, E.V. et al. Identification of erythroferrone as an
116a. Lebrón, J.A., West, A.P., and Bjorkman, P.J. The hemochromatosis protein erythroid regulator of iron metabolism. Nat Genet, 2014;46(7):678–84.
HFE competes with transferrin for binding to the transferrin receptor. 143. Kautz, L., Meynard, D., Monnier, A. et al. Iron regulates phosphorylation
J Mol Biol, 1999;294:239–45. of Smad1/5/8 and gene expression of Bmp6, Smad7, Id1, and Atoh8 in the
117. Waalen, J. and Beutler, E. Hereditary hemochromatosis: screening and mouse liver. Blood, 2008;112(4):1503–9.
management. Curr Hematol Rep, 2006;5(1):34–40. 144. Wang, C.Y., Core, A.B., Canali, S. et  al. Smad1/5 is required for
118. Allen, K.J., Gurrin, L.C., Constantine, C.C. et al. Iron‐overload‐related dis- erythropoietin‐mediated suppression of hepcidin in mice. Blood,
­
ease in HFE hereditary hemochromatosis. N Engl J Med, 2017;130(1):73–83.
2008;358(3):221–30. 145. Rivera, S., Nemeth, E., Gabayan, V. et al. Synthetic hepcidin causes rapid
119. Merryweather‐Clarke, A.T., Cadet, E., Bomford, A. et al. Digenic inherit- dose‐dependent hypoferremia and is concentrated in ferroportin‐containing
ance of mutations in HAMP and HFE results in different types of haemo- organs. Blood, 2005;106(6):2196–9.
chromatosis. Hum Mol Genet, 2003;12(17):2241–7. 146. Nicolas, G., Bennoun, M., Porteu, A. et al. Severe iron deficiency anemia
120. Constantine, C.C., Anderson, G.J., Vulpe, C.D. et al. A novel association in transgenic mice expressing liver hepcidin. Proc Natl Acad Sci U S A,
between a SNP in CYBRD1 and serum ferritin levels in a cohort study of 2002;99(7):4596–601.
HFE hereditary haemochromatosis. Br J Haematol, 2009;147(1):140–9. 147. Rivera, S., Liu, L., Nemeth, E. et al. Hepcidin excess induces the sequestra-
121. McLaren, C.E., Emond, M.J., Subramaniam, V.N. et al. Exome sequencing in tion of iron and exacerbates tumor‐associated anemia. Blood, 2005;105(4):
HFE C282Y homozygous men with extreme phenotypes identifies a GNPAT 1797–802.
variant associated with severe iron overload. Hepatology, 2015;62(2):429–39. 148. Weinstein, D.A., Roy, C.N., Fleming, M.D. et al. Inappropriate expression
122. Waalen, J., Felitti, V.J., Gelbart, T., and Beutler, E. Screening for hemo- of hepcidin is associated with iron refractory anemia: implications for the
chromatosis by measuring ferritin levels: a more effective approach. anemia of chronic disease. Blood, 2002;100(10):3776–81.
Blood, 2008;111(7):3373–6. 149. Theurl, I., Theurl, M., Seifert, M. et al. Autocrine formation of hepcidin
123. Camaschella, C., Roetto, A., Cali, A. et al. The gene TFR2 is mutated in a induces iron retention in human monocytes. Blood, 2008;111(4):2392–9.
new type of haemochromatosis mapping to 7q22. Nat Genet, 2000;25(1): 150. Eskew, J.D., Vanacore, R.M., Sung, L., Morales, P.J., and Smith, A. Cellular
14–15. protection mechanisms against extracellular heme. heme‐hemopexin, but
124. Roetto, A., Totaro, A., Piperno, A. et al. New mutations inactivating trans- not free heme, activates the N‐terminal c‐jun kinase. J Biol Chem,
ferrin receptor 2 in hemochromatosis type 3. Blood, 2001;97(9):2555–60. 1999;274(2):638–48.
125. Loreal, O., Ropert, M., Mosser, A. et al. [Pathophysiology and genetics of clas- 151. Baker, H.M., Anderson, B.F., and Baker, E.N. Dealing with iron: common
sic HFE (type 1) hemochromatosis]. Presse Med, 2007;36(9 Pt 2):1271–7. structural principles in proteins that transport iron and heme. Proc Natl
126. Fleming, R.E., Ahmann, J.R., Migas, M.C. et al. Targeted mutagenesis of Acad Sci U S A, 2003;100(7):3579–83.
the murine transferrin receptor‐2 gene produces hemochromatosis. Proc 152. Hvidberg, V., Maniecki, M.B., Jacobsen, C. et al. Identification of the receptor
Natl Acad Sci U S A, 2002;99(16):10653–8. scavenging hemopexin‐heme complexes. Blood, 2005; 106(7):2572–9.
127. Roetto, A., Papanikolaou, G., Politou, M. et al. Mutant antimicrobial pep- 153. Tolosano, E., Hirsch, E., Patrucco, E. et al. Defective recovery and severe
tide hepcidin is associated with severe juvenile hemochromatosis. Nat renal damage after acute hemolysis in hemopexin‐deficient mice. Blood,
Genet, 2003;33(1):21–2. 1999;94(11):3906–14.
228 THE LIVER:  REFERENCES

154. Fagoonee, S., Gburek, J., Hirsch, E. et  al. Plasma protein haptoglobin 161. Moyo, V.M., Gangaidzo, I.T., Gomo, Z.A. et al. Traditional beer consump-
modulates renal iron loading. Am J Pathol, 2005;166(4):973–83. tion and the iron status of spouse pairs from a rural community in
155. Schaer, D.J., Alayash, A.I., and Buehler, P.W. Gating the radical hemo- Zimbabwe. Blood, 1997;89(6):2159–66.
globin to macrophages: the anti‐inflammatory role of CD163, a scavenger 162. Moyo, V.M., Mandishona, E., Hasstedt, S.J. et  al. Evidence of
receptor. Antioxid Redox Signal, 2007;9(7):991–9. genetic transmission in African iron overload. Blood, 1998;91(3):
156. Fletcher, L.M., Halliday, J.W., and Powell, L.W. Interrelationships of alcohol 1076–82.
and iron in liver disease with particular reference to the iron‐binding proteins, 163. Tsukamoto, H., Horne, W., Kamimura, S. et al. Experimental liver cirrhosis
ferritin and transferrin. J Gastroenterol Hepatol, 1999;14(3):202–14. induced by alcohol and iron. J Clin Invest, 1995;96(1):620–30.
157. Suzuki, Y., Saito, H., Suzuki, M. et al. Up‐regulation of transferrin receptor 164. Fletcher, L.M. and Powell, L.W. Hemochromatosis and alcoholic liver dis-
expression in hepatocytes by habitual alcohol drinking is implicated in ease. Alcohol, 2003;30(2):131–6.
hepatic iron overload in alcoholic liver disease. Alcohol Clin Exp Res, 165. Kohgo, Y., Ohtake, T., Ikuta, K. et al. Iron accumulation in alcoholic liver
2002;26(8 Suppl):26S–31S. diseases. Alcohol Clin Exp Res, 2005;29(11 Suppl):189S–93S.
158. Gordeuk, V.R., Boyd, R.D., and Brittenham, G.M. Dietary iron overload 166. Flanagan, J.M., Peng, H., and Beutler, E. Effects of alcohol consumption on
persists in rural sub‐Saharan Africa. Lancet, 1986;1(8493):1310–13. iron metabolism in mice with hemochromatosis mutations. Alcohol Clin
159. Gordeuk, V., Mukiibi, J., Hasstedt, S.J. et  al. Iron overload in Africa. Exp Res, 2007;31(1):138–43.
Interaction between a gene and dietary iron content. N Engl J Med, 167. Ohtake, T., Saito, H., Hosoki, Y. et al. Hepcidin is down‐regulated in alco-
1992;326(2):95–100. hol loading. Alcohol Clin Exp Res, 2007;31(1 Suppl):S2–8.
160. Friedman, B.M., Baynes, R.D., Bothwell, T.H. et al. Dietary iron overload
in southern African rural blacks. S Afr Med J, 1990;78(6):301–5.
Disorders of Bilirubin
20 Metabolism
Namita Roy Chowdhury, Yanfeng Li, and Jayanta Roy Chowdhury
Departments of Medicine and Genetics and Marion Bessin Liver Research Center, Albert Einstein College of
Medicine, Bronx, New York, NY, USA

INTRODUCTION bilirubin produced daily in normal adults, ~80% is derived from


the hemoglobin, while the remainder comes from more rapidly
Bilirubin is the end product of degradation of the heme moiety turning‐over hemoproteins, predominantly in the liver.
of hemoproteins. Hemoglobin, derived from senescent erythro- Microsomal heme oxygenase (HO) cleaves heme at the α‐
cytes, is the major source of bilirubin. Significant fractions are methene bridge by a reaction requiring oxygen and a reducing
also derived from other hemoproteins of liver and other organs. agent, such as NADPH. HO activity is rate limiting in bilirubin
Historically, hyperbilirubinemia has attracted the attention of production and its inhihition by nonmetabolized “dead‐end”
clinicians as a marker of liver dysfunction. Subsequently, the inhibitors, such as tin‐protoporphyrin or tin‐mesoporphyrin,
studies of bilirubin chemistry, synthesis, transport, metabolism, reduces serum bilirubin levels in neonates [1, 2]. The reaction
distribution, and excretion have provided important insights opens the tetrapyrrole ring of heme, giving rise to a linear
into the transport, metabolism, and excretion of biologically tetrapyrrole, biliverdin, and releasing 1 mol each of carbon
important organic anions, particularly those with limited aque- monoxide (CO) and iron. Normally, oxidation of the α‐chain
ous solubility. carbon of heme accounts for the great majority of the body’s
Bilirubin is cytoprotective at relatively low concentrations, endogenous CO production, only a small fraction being contrib-
but is potentially toxic at higher levels. It is normally rendered uted by intestinal bacteria [3]. Therefore, measurement of CO
harmless by tight binding to albumin in the circulation, efficient production by breath analysis can be used to quantify bilirubin
extraction and storage by hepatocytes, enzyme‐catalyzed detox- production and heme breakdown, which, at a steady state,
ification, and subsequent active biliary excretion. Patients with equals heme synthesis.
very high levels of unconjugated hyperbilirubinemia are at risk Of the two HO isoforms, HO‐1, expressed from the HMOX1
for bilirubin encephalopathy (kernicterus). Kernicterus is found gene, is ubiquitous and inducible by protoheme IX and various
in some cases of severe neonatal jaundice and in inherited disor- stress signals. The kelchlike ECH‐associated protein 1 (KEAP1)–
ders associated with severe unconjugated hyperbilirubinemia. nuclear factor E2‐related factor 2 (Nrf2) pathway regulates HO‐1
This chapter will provide a brief summary of bilirubin metabo- expression. At resting state, KEAP1 degrades NRF2. Oxidative/
lism and its inherited disorders, with emphasis on some signifi- electrophilic modification of KEAP1 or its sequestration in
cant recent developments. autophagosomes permits stabilization of NRF2 and its transloca-
tion to the nucleus, where it activates target genes, including
HO‐1 [4].
HO‐1 and its products, biliverdin and bilirubin, provide tissue
BILIRUBIN METABOLISM protection against reactive oxygen radicals, and CO released by
the reaction regulates the vascular tone in the liver and in other
Bilirubin is derived from the breakdown of hemoglobin released organs, such as the heart, under conditions of stress. HO‐1
from senescent erythrocytes and other hemoproteins, particu- induction in the gastrointestinal tract protects against ischemia–
larly cytochromes and several enzymes. Of the 250–400 mg of reperfusion injury, lipopolysaccharide (LPS)‐associated sepsis,

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
230 THE LIVER:  POTENTIAL BENEFICIAL EFFECTS OF BILIRUBIN

as well as tissue injury caused by indometacin, trinitrobenzene dipyrrolic halves are joined by a central methane bridge. Each
sulfonic acid, and dextran sulfate [5]. Inherited HO‐1 deficiency half has a propionic acid side‐chain. X‐ray diffraction crystal-
is associated with extensive endothelial injury, causing con- lography and nuclear magnetic resonance studies have revealed
sumption coagulopathy and microangiopathic hemolytic ane- that the propionic acid side‐chains of bilirubin are internally
mia. In addition, these patients have growth retardation and hydrogen‐bonded to the pyrrolic and lactam sites on the oppo-
hyperlipidemia [6]. The cytoprotective function of HO‐1 induc- site half of the molecule [14]. The internal hydrogen bonding
tion is also important in the renal vascular endothelium and engages all polar groups and “buries” the central methane
tubular epithelium. Some effects of HO‐1 induction, such as bridge that joins the two dipyrrolic halves of the molecule.
induction of subsets of regulatory T‐lymphocytes may be unre- Physiologically, the hydrogen bonds can be disrupted by
lated to its products, CO and bilirubin [7]. enzyme‐catalyzed conjugation of one or both propionic acid
The released iron is reutilized, but can be deleterious through side‐chains, forming bilirubin mono‐ and diglucuronides,
mitochondrial non‐transferrin iron sequestration that can con- respectively.
tribute to bioenergy failure, such as in Alzheimer disease. HO‐2 In the van den Bergh reaction [15] that is commonly employed
is expressed constitutively, mainly in the brain, where its upreg- in clinical analysis of serum bilirubin levels, the diazo reagents,
ulation in response to anoxia may have a protective function which attack the central methane bridge, act rapidly on the non‐
during ischemic attacks [8]. hydrogen‐bonded conjugated bilirubin (“direct”‐reacting biliru-
In most vertebrates, including early vertebrates, such as bin), whereas the unconjugated fraction reacts rapidly only
early teleost and elasmobranch fish [9], biliverdin is reduced to when the hydrogen bonds are disrupted by a chemical accelera-
bilirubin by the action of two biliverdin reductases, BVRA and tor (“total” bilirubin). The hydrogen bonds can also be disrupted
BVRB. BVRB, which is expressed earlier in embryogenesis, transiently by configurational isomerization of bilirubin, which
catalyzes the reduction of biliverdin IXβ, the product of fetal occurs upon exposure to light. These bilirubin photoproducts
heme‐IXβ, whereas BVRA mediates the reduction of biliverdin can be excreted into bile without conjugation, which explains
IXα, the product of adult heme‐IXα [10]. The gene encoding the efficacy of phototherapy in reducing serum unconjugated
BVRA is located on chromosome 7 (pter>q22) [11], whereas bilirubin levels.
the BVRB‐encoding gene is on chromosome 19. The catalytic
site for biliverdin reduction is located within the N‐terminal
domain of both isozymes [12]. However, the C‐terminal domain
of BVRA contains a basic‐leucine‐zipper (bZiP) domain to POTENTIAL BENEFICIAL EFFECTS
function as a transcription factor, which does not exist in OF BILIRUBIN
BVRB. The C‐terminal domain of BVRA also contains both a
nuclear localization sequence (NLS) and a nuclear export Although clinicians are generally concerned with serum
sequence (NES), enabling BVRA to bind to DNA sequences bilirubin levels as an indicator of liver function and disease,
such as the antioxidant response element (ARE) and hypoxia and the toxicity of bilirubin, within a near‐physiological
response elements (HRE), thereby recruiting Nrf2. Nrf2 range of plasma concentrations, the antioxidative action of
induces HO‐1, which protects against oxidative injury. bilirubin may provide beneficial effects. Serum bilirubin
Additionally, BVRA is a member of the insulin receptor sub- levels were found to be inversely related to the risk of obesity
strate family with serine/threonine/tyrosine kinase activity and and metabolic syndrome [16] and ischemic coronary artery
through its C‐terminal region interacts with both major arms of disease in middle‐aged men [17]. An inverse relationship
the insulin signaling pathway, namely the phosphatidylinositol between serum bilirubin levels and cancer mortality was
3‐kinase (PI3‐kinase)/Akt pathway and the IRK/IRS/PI3‐ reported [18]. The study of a large number of subjects in the
kinase/MAPK pathway [10]. United States revealed that the odds ratio for history of
The evolutionary conservation of the energy‐consuming pro- colorectal cancer was reduced by 0.295 in men and 0.186 in
cess of generation and excretion of the nonpolar bilirubin sug- women per 1 mg dL−1 increment in serum bilirubin concentra-
gests that the stronger antioxidant activity of bilirubin may be tions [19]. Subjects with mildly elevated plasma bilirubin lev-
particularly important during the neonatal period, when concen- els were found to have lower levels of abdominal obesity and
trations of other intracellularly available antioxidants are low in reduced risk of metabolic syndrome. Consistent with this,
body fluids. Another potential advantage of bilirubin formation obese individuals with elevated insulin and visceral adiposity
may be that, being more nonpolar, bilirubin is more efficiently had lower plasma bilirubin [20].
extracted by the placenta in intrauterine life, although the mech- It should be noted that these convincing statistical associa-
anism of bilirubin extraction by the placenta from the fetal cir- tions do not, by themselves, establish a causative role of
culation has not been elucidated fully. Cord blood bilirubin bilirubin in reducing the incidence of a number of common
concentrations of fetuses born to mothers who have unconju- diseases in the population. Mechanistically, HO‐1 induction
gated hyperbilirubinemia due to bilirubin glucuronidation by cobalt‐protoporphyrin administration in leptin‐deficient
deficiency are similar to the maternal serum bilirubin levels, ob/ob mice led to the recruitment of FGF21, PPARα, and
suggesting that the placenta does not pose a barrier to equilibra- Glut1, thereby reducing hepatic heme, body weight gain,
tion of maternal and fetal serum unconjugated bilirubin [13]. plasma glucose, fatty acid synthase, and hepatic steatosis
As heme oxygenase acts specifically at the α‐bridge of heme, [21]. Many of these effects could be reproduced by bilirubin
physiologically generated biliverdin and bilirubin are desig- administration in mice with diet‐induced obesity or leptin
nated as biliverdin IXα and bilirubin IXα, respectively. The two receptor deficiency (db/db) [22].
20:  Disorders of Bilirubin Metabolism 231

HEPATIC METABOLISM because of acquired or inherited liver diseases, a significant


AND ELIMINATION OF BILIRUBIN amount of conjugated bilirubin is excreted in urine. In the pres-
ence of prolonged accumulation of conjugated bilirubin in
Bilirubin circulates in plasma bound to albumin. Albumin bind- plasma, a fraction of the conjugated bilirubin becomes cova-
ing keeps unconjugated bilirubin in solution and prevents its dif- lently bound to albumin. The covalently bound fraction, termed
fusion into tissues and all its toxic effects. Unconjugated delta‐bilirubin, is not excreted in bile or urine, and may persist
bilirubin binds more tightly to albumin than conjugated biliru- in plasma even after biliary obstruction or intrahepatic cholesta-
bin. Therefore, in the absence of proteinuria, unconjugated bili- sis is relieved.
rubin does not undergo glomerular filtration significantly. As In the liver, bilirubin dissociates from albumin and is internal-
albumin is normally present in approximately threefold molar ized by hepatocytes via facilitated diffusion. Although the trans-
excess to bilirubin, there is a significant reserve bilirubin bind- port protein SLC21A6 (OATP‐2) was implicated in bilirubin
ing capacity, which acts as a buffer for fluctuations of serum internalization, this could not be confirmed in subsequent studies
bilirubin levels. However, a number of metabolites and drugs [26]. Within the hepatocyte, binding of bilirubin to glutathione‐
affect albumin binding of bilirubin and thereby risk of neurotox- S‐transferases (GSTs) inhibits its efflux, thereby increasing
icity. Therefore, measurement of unbound plasma bilirubin and the net uptake. Microsomal uridinediphosphoglucuronate glucu-
the reserve bilirubin binding capacity could provide a more ronosyltransferase type 1 (UGT1A1) catalyzes the transfer of glu-
accurate estimate of the risk of bilirubin‐induced neurological curonic acid from UDP‐glucuronate to bilirubin, forming
damage (BIND). This is particularly important in premature mono‐ and diglucuronides. Glucuronidation makes bilirubin
infants, in whom the threshold total plasma bilirubin concentra- water soluble, reduces its toxicity, and promotes its secretion into
tions used for instituting phototherapy and/or exchange transfu- bile. Glucuronide conjugation is critical in biliary excretion of
sion in full‐term infants may be misleading. Directly measured bilirubin. Significant reduction of hepatic bilirubin glucuronidat-
free (unbound) serum bilirubin levels (Bf) are more sensitive ing activity results in the accumulation of unconjugated bilirubin
and specific predictors of BIND, as evidenced by incidence of in plasma (see later). Finally, the bilirubin glucuronides are
hearing defect than total serum bilirubin or bilirubin/albumin transported into the bile canaliculi by an energy‐consuming pro-
ratio [23]. Peroxidase treatment, gel chromatography, electro- cess mediated by ABCC2 (also termed MRP2) (Figure 20.1).
phoretic analysis, and direct fluorometry have been used for Bf Among the many isoforms of uridinediphosphoglucuro-
determination [24], However, these approaches are not in clini- nate glucuronosyltransferase (UGT), UGT1A1 is the only
cal use, except for Bf determination by peroxidase treatment, one that contributes significantly to bilirubin glucuronida-
which is used routinely in Japan [25]. tion. UGT1A1 [27] also accepts estradiol and several drugs
Because of the tight binding of unconjugated bilirubin to as substrates. UGT1A1 is expressed from an unusually
albumin, bilirubin is not excreted in the urine in the absence of organized gene that expresses eight UGT1 isoforms from
proteinuria. Conjugated bilirubin binds less tightly to albumin. eight different promoters, each next to a unique exon encod-
Therefore, when conjugated bilirubin accumulates in plasma ing the N‐terminal half of a specific isoform. The transcript

Bilirubin
Albumin + glucuronides
Albumin-Bilirubin Bilirubin
Sinusoidal 1
surface 6 7
2
3
Contiguous GST’s Bilirubin ABCC3 OATP01B1 OATP01B3
membrane UDPGA
4 UGT1A1
UDP Bilirubin
5 glucuronides
Bilirubin glucuronides ABCC2
(MRP2)
ABCC2 Canalicular
(MRP2) surface

Figure 20.1  Summary of hepatic metabolism of bilirubin. Bilirubin is strongly bound to albumin in the circulation (1). At the sinusoidal surface
of the hepatocyte, this complex dissociates, and bilirubin enters hepatocytes by fascilitated diffusion (2). This process is non‐adenosine triphosphate
(ATP)‐dependent and bidirectional. Within the hepatocyte, bilirubin binds to a group of cytosolic proteins, mainly to glutathione‐S‐transferases
(GSTs) (3). GST binding inhibits the efflux of bilirubin from the cell, thereby increasing the net uptake. A specific form of uridine diphosphoglu-
curonate glucuronosyltransferase, UGT1A1, located in the endoplasmic reticulum, catalyzes the transfer of the glucuronic acid moiety from UDP‐
glucuronic acid (UDPGA) to bilirubin, forming bilirubin glucuronides (diglucuronide and monoglucuronide) (4). Glucuronidation is necessary for
efficient excretion of bilirubin in bile. Canalicular excretion of bilirubin and other organic anions (except most bile acids) is primarily an energy‐
dependent process, mediated by the ATP‐utilizing transporter ABCC2, also known as multidrug‐resistance‐related proteins (MRP2) (5). Excess
bilirubin glucuronides are pumped back into the plasma by ABCC3 located at the sinusoidal membrane (6) and undergo reuptake by hepatocytes
located downstream to the portal blood flow via sinusoidal surface organic anion transport proteins, OATP01B1 and OATP01B3 (7).
232 THE LIVER:  DISORDERS OF BILIRUBIN METABOLISM

initiated from each promoter is spliced to four common unconjugated hyperbilirubinemia and subjects with inherited
region exons (exons 2–5) that encode the identical C‐termi- disorders of bilirubin conjugation.
nal half of all UGT1A isoforms [28]. The presence of a dif-
ferent promoter for each isoform permits its independent
regulation. Mutations in any of the five exons (1A1–5)
Neonatal hyperbilirubinemia
encoding UGT1A1 can cause complete or partial deficiency Normally, newborns have higher serum bilirubin levels than adults.
of bilirubin glucuronidation, resulting in Crigler–Najjar Eighty percent of all term infants exhibit clinical jaundice during
syndrome type 1 (CN1) or Crigler–Najjar syndrome type 2 the first 5 days of life. For clinical management, serum bilirubin
(CN2), which are characterized by hyperbilirubinemia with levels in newborns need to be interpreted according to the infant’s
potential cerebral injury [29, 30]. Gilbert syndrome, a milder age in hours. Bilirubin levels increase during the first few days of
and much more common type of inherited hyperbilirubine- life, typically peaking at the age of 96 hours, when the 50th percen-
mia, results from a variant TATA element within the tile in healthy newborns in Western countries ranges from 8 to 9 mg
UGT1A1 promoter, which leads to reduced synthesis of dL−1 and the 95th percentile is 15–17.5 mg dL−1 [36]. These levels
structurally normal UGT1A1 protein [31]. are considered to be harmless. After this, serum bilirubin levels
Finally, the bilirubin glucuronides are transported into the decline to less than 1 mg dL−1 in 7–10 days. Exaggeration of this
bile canaliculi against a steep concentration gradient by the physiological jaundice increases the risk of BIND.
ATP‐hydrolyzing canalicular pump ABCC2 (also termed Neonatal hyperbilirubinemia results from a combination of
MRP2), which is rate limiting in bilirubin throughput. Thus, increased bilirubin production and lower hepatic bilirubin excre-
efficient uptake and conjugation of bilirubin by the periportal tory capacity. Exacerbation of these factors, with or without
(zone 1) hepatocytes that are initially exposed to bilirubin in additional complicating disorders, increases BIND risk. In
portal blood may saturate the storage capacity of these cells. addition to prematurity, common risk factors for severe hyperbili-
A portion of the bilirubin glucuronides produced in these rubinemia in babies born at 35 weeks or more gestation are exclusive
cells is secreted into the sinusoidal plasma by the multispe- breastfeeding (particularly with excessive weight loss), clinical
cific sinusoidal export pump MRP3 (ABCC3). Conjugated jaundice noted within the first 24 hours, hemolytic diseases (e.g.
bilirubin flowing toward the central vein undergoes reuptake glucose 6‐phosphate dehydrogenase (G6PD) deficiency), and
via the organic anion‐transporting polypeptides OATP1B1 cephalohematoma or significant bruising during birth [37].
(SLC01B1) and OATP1B3 (SLC01B3) located in hepatocyte Contribution of genetic factors is suggested by increased
sinusoidal membranes. This process recruits additional hepat- incidence of hyperbilirubinemia in infants of East Asian ethnicity
ocytes, thereby increasing the bilirubin‐excreting capacity of and history of neonatal jaundice in a previous sibling [38].
the liver [32]. Mechanisms of neonatal jaundice are briefly considered here.
Bilirubin is degraded by intestinal bacteria into a series of
urobilinogen and related products. Most of the urobilinogen Increased bilirubin production
reabsorbed from the intestine is excreted in bile, but a small
Bilirubin production, as measured by CO excretion in breath, is
fraction is excreted in urine. Absence of urobilinogen in stool
increased during the newborn period [39]. The excess bilirubin is
and urine indicates complete obstruction of the bile duct. In
derived from shortened erythrocyte half‐life and also from non-
liver disease and states of increased bilirubin production, uri-
erythroid sources [40]. Mother–fetus Rh incompatibility has
nary urobilinogen excretion is increased. Urobilinogen is color-
become infrequent since the availability of anti‐Rh immunoglobu-
less; its oxidation product, urobilin, contributes to the color of
lins [41], but ABO blood group incompatibility continues to be a
normal urine and stool.
common cause of exaggerated neonatal hyperbilirubinemia. Other
Normally, the capacity of the liver for bilirubin uptake,
common hemolytic disorders include sickle cell disease, glucose
conjugation, and excretion is closely balanced, so that reduc-
6‐phosphate dehydrogenase deficiency, hereditary spherocytosis,
tion of any of these processes can limit the rate of bilirubin
and toxic or allergic drug reactions. Ineffective erythropoiesis, as in
throughput by the liver. On the other hand, increase of the
thalassemia, vitamin B12 deficiency, and congenital dyserythro-
bilirubin handling capacity of the liver, for example, in
poietic anemias can also result in excessive bilirubin production.
response to increased bilirubin load, requires that all these
pathways are coordinately upregulated. Several nuclear
receptor proteins, such as CAR and PXR, may regulate such Low hepatic bilirubin uptake
coordinated regulation [33–35]. Low uptake rate during the first few days of life may result from
delayed closure of the ductus venosus and low levels of cyto-
solic GSTs [42], which increase net bilirubin uptake by reduc-
ing its efflux.
DISORDERS OF BILIRUBIN
METABOLISM Reduced bilirubin glucuronidation
At higher concentrations, unconjugated bilirubin is toxic to Both full‐term and premature infants are born with approxi-
many cells and organelles. Since albumin binding abrogates the mately 1% of adult hepatic UGT1A1 activity, which increases
toxic effect of bilirubin, the harmful effects occur when uncon- to adult levels by 14 weeks, regardless of the gestational age at
jugated bilirubin is present in a molar excess over albumin. Such birth [43]. Inhibition of UGT1A1 exacerbates and prolongs
conditions are usually limited to neonates with a high level of ­neonatal jaundice.
20:  Disorders of Bilirubin Metabolism 233

Maternal milk jaundice common benign disorder called Gilbert syndrome, which is
characterized by a mild, fluctuating, and often intermittent
In general, breastfed infants have higher serum bilirubin levels
unconjugated hyperbilirubinemia.
than formula‐fed babies [44]. Mild maternal milk jaundice may
abate despite continuation of breastfeeding and even in more
severe cases, resolves promptly upon discontinuation of breast- Crigler–Najjar syndrome type 1
feeding, If breastfeeding is continued, the hyperbilirubinemia
This rare syndrome with autosomal recessive inheritence was
may persist for weeks and, in some cases, may increase to 15–24
described in 1952 by Crigler and Najjar in six infants from three
mg dL−1 by the age of 10–19 days. Although maternal milk jaun-
unrelated families [51] and was later found to result from lack of
dice is usually benign [45], kernicterus can occur in rare
UGT1A1 activity [52]. All patients had lifelong severe non-
cases  [46]. The mechanism of maternal milk jaundice is not
hemolytic unconjugated hyperbilirubinemia resulting in bilirubin‐
clear. Polyunsaturated free fatty acids produced by lipolytic
induced encephalopathy and death within 15 months in five of
enzymes in some maternal milk samples may be responsible for
the six originally reported cases. The remaining patient devel-
the inhibition of bilirubin glucuronidation, as suggested by
oped kernicterus for the first time at the age of 15 years, and
increased inhibitory effect of maternal milk upon storage and
died six months after that [52]. A related patient remained
marked reduction of the inhibitory effect by heating the milk to
without brain damage until 18 years of age, but then developed
56°C [47].
kernicterus and died at the age of 24 [53]. As the original fami-
lies described by Crigler and Najjar had a high degree of
Maternal serum jaundice
consanguinity, several other recessively inherited disorders,
­
Lucey and associates [48] described a syndrome manifested by including Morquio syndrome, homocystinuria, metachromatic
moderate to severe unconjugated hyperbilirubinemia (8.9–65 leukodystrophy, and bird‐headed dwarfism existed in these fam-
mg dL−1) within the first 4 days of life, which is thought to be ilies. However, in other families CN1 was not associated with
caused by an unidentified inhibitor of UGT1A1 present in additional inherited disorders. Several hundred CN1 patients
maternal serum. Jaundice may persist several weeks and is were described subsequently in all races.
occasionally associated with kernicterus. After Arias reported a milder variant of this condition (see
below under Crigler–Najjar syndrome type 2), the original
Delayed maturation of canalicular bilirubin excretion potentially lethal Crigler–Najjar syndrome was designated
Crigler–Najjar syndrome type 1 (CN1), whereas the milder var-
Maturation of canalicular excretion mechanism may lag behind
iant of the disease was termed Crigler–Najjar syndrome type 2
the maturation of uptake and conjugation, so that in the late
(CN2). Jaundice is often the only clinical finding, although
newborn period, canalicular excretion becomes rate limiting in
some patients may have residual neurologic abnormalities, from
hepatic bilirubin throughput. In these cases, conjugated biliru-
previous episodes of bilirubin encephalopathy. With routine use
bin may accumulate in serum [49].
of phototherapy and plasmapheresis during acute bilirubin
encephalopathy, many patients with CN1 now survive beyond
Increased intestinal reabsorption
childhood, but many survivors develop kernicterus around
Deconjugation of bilirubin by intestinal β‐glucuronidase puberty or in early adult life [54]. Orthotopic or auxilliary liver
releases unconjugated bilirubin, which is not further degraded transplantation results in normalization of serum bilirubin.
because of the absence of an established intestinal microbiota in Because of the relatively high concentration of unconjugated
the newborn. This results in increased bilirubin absorption [50], bilirubin excreted in bile as a result of phototherapy, pigment
which may be enhanced by feeding maternal milk. gallstones are common.

Laboratory tests
DISORDERS ASSOCIATED Serum bilirubin levels usually range from 20 to 25 mg dL−1, but
WITH UNCONJUGATED may reach 50 mg dL−1 [51, 52]. Serum bilirubin is all unconju-
gated and tightly bound to albumin, therefore bilirubinuria is
HYPERBILIRUBINEMIA
absent. The bile contains only small amounts of unconjugated
bilirubin [55]. Although fecal urobilinogen excretion is reduced,
Predominantly unconjugated hyperbilirubia can result from
the stool color remains normal [51]. Normal bile canalicular
increased bilirubin production due to hemolytic disorders or
transport is evidenced by normal plasma clearance of bromosul-
ineffective erythropoiesis. In the presence of normal liver func-
fophthalein and indocyanin green, as well as visualization of the
tion, increased bilirubin production does not lead to plasma lev-
biliary tree by cholecystographic agents [51].
els of bilirubin exceeding 5 mg dL−1.
Three inherited disorders associated with UGT1A1 defi-
Liver histology
ciency and consequent reduction of bilirubin glucuronidation
have been described. A near‐complete deficiency of UGT1A1 Historically, liver histology has been reported as normal other
activity results in Crigler–Najjar syndrome type 1 (CN1); severe than bilirubin plugs in bile canaliculi and bile ducts [51, 55].
but incomplete deficiency of UGT1A1 activity is termed However, a recent systematic analysis of 22 CN1 cases under-
Crigler–Najjar syndrome type 2 (CN2), also known as Arias going liver transplantation at a single center showed liver fibro-
syndrome; and a mild reduction of UGT1A1 activity results in a sis of various degrees in 41% of the explanted livers [56]. The
234 THE LIVER:  DISORDERS ASSOCIATED WITH UNCONJUGATED HYPERBILIRUBINEMIA

liver fibrosis was not associated with portal hypertension and polymerase chain reaction (PCR) amplification of the five
there was no significant correlation with gallstones. The trans- UGT1A1 exons and their flanking splice sites. The same strat-
plant recipients with liver fibrosis were older in age, suggesting egy can be used for prenatal diagnosis by analyzing DNA
an accrual of the risk of fibrosis with age. extracted from amniotic cells or chorionic villus samples [64].
As in other recessively inherited inherited disorders, in nearly
Abnormalities of hepatic UGTs all CN1 and CN2 cases, one mutant allele is derived from each
heterozygous parent, giving rise to homozygous or compound
Hepatic UGT activity toward bilirubin is virtually absent in all
homozygous states. However, a person with CN1 with unipa-
patients with CN1. In addition, many of these patients have
rental isodisomy has been reported, in whom both mutant alleles
reduced glucuronidation of phenolic substrates [57], which can
were inherited from the father [57]. The mother’s UGT1A1
be explained by the location of mutation in UGT1A1 exons (see
genotype was normal. This case highlights the desirability of
later). By immunological analysis, expression of UGT isoforms
analyzing the genotype of both parents to determine the mode of
may vary in the livers of different CN1 patients [58].
inheritance of CN syndromes.
Molecular bases of CN1 and CN2
Animal models of CN1
Description of the layout of the UGT1A locus in 1992 by Ritter
et al. [59] and Bosma et al. [60] explained the structural charac- A mutant strain of Wistar rats exhibiting lifelong nonhemolytic
teristics and different substrate specificities of various isoforms unconjugated hyperbilirubinemia inherited as an autosomal
of the UGT1A family. Each isoform has a unique N‐terminal recessive characteristic was reported by Gunn in 1938 [65].
domain but they all have identical C‐terminal domains. Subsequently, the cause of jaundice was found to be deficiency
Processed UGT1 mRNAs consist of five exons. The UGT1A of UGT‐mediated glucuronidation of bilirubin. Jaundiced Gunn
locus comprises four exons encoding the common C‐terminal rats are homozygous for the deletion of a single guanosine resi-
domain of all UGT1A isoforms that contains the binding site of due in the common region exon 4, causing a frameshift and a
the glucuronic acid donor substrate uridine diphosphoglucu- premature termination codon that results in the expression of a
ronic acid and the single transmembrane region of the protein. truncated protein lacking 150 amino acid residues at the C‐
Twelve unique exon 1 sequences are located 5′ to these common terminus. This results in the loss of activity of all UGT isoforms
region exons, only one of which is used in a given UGT1A iso- expressed from the UGT1 locus, but UGT isoforms expressed
form, encoding its unique N‐terminal domain that imparts agly- from other loci (e.g. UGT2) are not affected [66]. Experiments
cone substrate specificity to specific isoforms. Each unique using Gunn rats have provided important information on biliru-
exon 1 has a separate proximal upstream promoter, so that bin toxicity and have helped in developing novel cell and gene‐
depending on which promoter is used for transcription, RNA based therapies of CN1 [67–69].
transcripts of different lengths are produced. In each transcript A mouse knockout model of CN1 was created by disrupting
only the exon 1 located at the 5′ end of the transcript is spliced exon 4 of the mouse UGT1 locus [70]. The knockout mice have
to the common region exon 2, and all other exon 1’s are treated higher bilirubin levels than Gunn rats, and have a high spontane-
as intervening sequences and deleted. This results in the ous mortality rate unless treated with intensive phototherapy.
translation of nine UGT1 isoforms, of which only UGT1A1 The UGT1‐KO mice have enabled the pathophysiological study
contributes significantly to bilirubin glucuronidation [61]. of BIND and development of novel therapeutic strategies.
Consequently, genetic lesions within any of the five exons com-
prising the UGT1A1 gene may lead to complete or near‐com- Treatment of CN1
plete loss of hepatic bilirubin glucuronidation. Management of CN1 centers on maintaining serum bilirubin
Such genetic lesions may consist of point mutations, dele- concentrations below neurotoxic levels. Partial or whole liver
tions, or insertions within the coding region or introns at splice transplantation cures the disease, but commits the pateint to pro-
donor or acceptor sites [29, 54]. The genetic lesions may result longed immunosuppressive therapy. Therapies based on hepato-
in mutation of a single critical amino acid or deletion of seg- cyte transplantation and gene therapy are still considered
ments of the enzyme. The various genetic lesions described in experimental. A brief discussion of these treatment modalities
the literature have been reviewed [29, 62]. In cases where the follows.
genetic lesions are present in exons 2–5, all isoforms expressed
from the UGT1A locus are affected, whereas mutation located
Phototherapy
in the unique first exon of UGT1A1 affects the gluronidation of
only the substrates of this isoform. Phototherapy is routinely used to reduce the level of plasma
As CN1 or CN2 can be caused by any of a large number of unconjugated bilirubin [55]. Banks of fluorescent lamps, or
mutations, deletions, or insertions, no particular mutation is more recently light‐emitting diode (LED) lamps, are used with
very common in any ethnic group or community. An exception devices for shielding the eyes. LED “light blankets,” or “light
to this is seen where there is a strong founder effect and a high jackets,” have been also devised. Light converts bilirubin IXα‐
level of consanguinity, such as in the Amish–Mennonite com- ZZ to its photoisomers (see the section on Bilirubin chemistry),
munity, in which there is a high incidence of CN1 and all people which are excreted in bile and partly degraded. During the neo-
with CN1 carry a specific nonsense mutation in exon 1 of natal period, use of phototherapy is based on age‐related serum
UGT1A1 [63]. In nearly all cases molecular diagnosis of CN1 bilirubin concentrations: 15 mg dL−1 (260 μM) at age 24–48
and CN2 can be made by sequence determination of products of hours, 18 mg dL−1 (310 μM) at 49–72 hours, and 20 mg dL−1
20:  Disorders of Bilirubin Metabolism 235

(340 μM) above 72 hours [24]. If serum bilirubin remains above disorders [75]. Because of this and the increasing shortage of
these levels and is not reduced by at least 1–2 mg dL−1 within good‐quality donor livers for hepatocyte isolation, novel strate-
4–6 hours despite intensive phototherapy, plasmapheresis is gies are being explored to induce preferential proliferation of
considered. The efficiency of phototherapy is reduced beyond transplanted normal hepatocytes over the mutant host cells. As
the age of 3 or 4 years, because of the thickening of skin, pig- adult hepatocytes retain a remarkable capacity to proliferate
mentation, and reduction of surface area relative to body mass, and the liver to body weight ratio is regulated tightly by physi-
requiring readjustment of photherapy intensity and duration. ological mechanisms, transplanted hepatocytes must compete
with host liver hepatocytes for preferential proliferation.
Plasmapheresis Controlled regional irradiation of the liver, in combination with
a variety of mitotic stimuli can enhance both initial hepatocyte
During neurologic emergencies, serum bilirubin concentration
engraftment in Gunn rats and subsequent proliferation of the
can be acutely reduced by plasmapheresis [55]. If serum biliru-
donor cells, leading to normalization of serum bilirubin levels
bin levels exceed the target levels described in the previous
[76]. Following initial evaluation in non‐human primates [77],
paragraph by 5 mg dL−1, despite intensive phototherapy, plas-
a clinical trial has been initiated to evaluate preparative hepatic
mapheresis is added to continued intensive photherapy, because
irradiation for hepatocyte transplantation in patients with inher-
after removal of albumin‐bound bilirubin from blood, bilirubin
ited metabolic disorders of the liver [78]. Hepatocyte‐like cells
is mobilized from tissue stores to the plasma, leading to second-
(iHep) generated by differentiating human‐induced pluripotent
ary increase of bilirubin levels.
stem cells derived by reprogramming somatic cells (e.g. skin
fibroblasts, bone marrow cells, peripheral blood mononuclear
Orthotopic liver transplantation
cells, or epithelial cells shed in the urine) can be transplanted
At present, the transplantation of whole liver or a segment of the into the livers of experimental animals. Transplantation of
liver is the only available curative therapy for CN1 [71]. human iHep cells into immunosuppressed Gunn rats subjected
Although this procedure commits the patient to prolonged to preparative X‐irradiation resulted in significant reduction of
immunosuppressive therapy, liver transplantation has dramati- serum bilirubin levels [69].
cally improved the outlook for CN1 patients.
Gene therapy
Experimental methods for reducing serum bilirubin levels Gene therapy approaches aimed at reconstitution of the missing
UHT1A1 activity include: (i) ex vivo gene therapy, which con-
Inhibiting heme oxygenase activity
sists of transduction of primary hepatocytes using viral vectors,
Non‐iron metalloporphyrins are dead‐end inhibitors of microso- followed by transplantation into the liver; (ii) systemic adminis-
mal heme oxygenase [72]. Tin‐protoporphyrin injection sup- tration of viral or nonviral vectors to transduce hepatocytes in
presses neonatal hyperbilirubinemia in rhesus monkeys [73]. vivo with transcription units expressing UGT1A1; and (iii) tar-
Injection of tin‐mesoporphyrin (0.5 μmol kg−1, three times a geted gene editing through homologous recombination for cor-
week for 13–23 weeks) in two 17‐year‐old male CN1 patients rection of a specific mutation in the UGT1A1 gene, or for
reduced serum bilirubin concentrations modestly. However, insertion of a UGT1A1 transcription unit at a genomic “safe
the  duration and safety of this treatment of CN1 are not yet haven” site of choice, or for targeted insertion of a UGT1A1
established. open reading frame downstream to a highly expressed gene (e.g.
albumin) to take advantage of the strong endogenous promoter.
Hepatocyte transplantation After validation in preclinical experiments, a clinical trial has
been initiated in CN1 patients using recombinant adeno‐associated
As UGT1A1 activity is present in excess in normal liver, partial
viral vectors.
replacement of the enzyme activity in livers of CN1 patients
should reduce serum bilirubin to nontoxic levels. After exten-
sive validation in Gunn rats [68], isolated allogeneic human
CN2 (Arias syndrome)
hepatocytes were transplanted into the liver of an adolescent
CN1 patient through a catheter placed percutaneously into the In this milder variant of Crigler–Najjar syndrome described by
portal vein [74]. Transplantation of 7.5 × 109 hepatocytes Arias in 1962 [79], serum bilirubin, which is mostly unconju-
reduced bilirubin levels by about 50% and enabled reduction of gated, usually ranges from 8 to 18 mg dL−1. During intercurrent
the duration of phototherapy [74]. However, although bilirubin illness, general anesthesia, or prolonged fasting, serum bilirubin
glucuronides were detectable in bile for up to two and a half can increase to 40 mg dL−1 [80]. Kernicterus is unusual, but has
years, serum bilirubin level gradually increased to pretrans- been reported during episodes of exacerbated hyperbilirubine-
plantation levels. The patient received an auxiliary liver trans- mia [79–81]. As in CN1, there is no evidence of hemolysis or
plantation, which rapidly reduced the serum bilirubin levels other liver dysfunction. CN2 can be clinically differentiated
to normal (J. Roy Chowdhury, personal communication). from CN1 by at least 25% decrease of serum bilirubin after
Experience in this case and a number of other cases of hepato- induction of the residual UGT1A1 activity by administration of
cyte transplantation for CN1, as well as for a number of other drugs, such as phenobarbital. Also in contrast to CN1, bile con-
inherited liver diseases indicated that the number of adult tains a significant amount of bilirubin glucuronides. In CN2, as
hepatocytes that can be transplanted at a single procedure is in Gilbert syndrome (see later), bilirubin monoglucuronide
insufficient for fully curing inherited liver‐based metabolic exceeds 30% of total conjugated bilirubin (normal, ~10%),
236 THE LIVER:  DISORDERS ASSOCIATED WITH UNCONJUGATED HYPERBILIRUBINEMIA

reflecting a reduced hepatic UGT1A1 activity. Liver histology within the proximal promoter region of UGT1A1 provides a
is normal, and UGT1A1 activity is usually reduced to 10% simple means of diagnosis (see section on Genetic basis of
normal [82]. Gilbert syndrome). If confirmation of diagnosis is required,
chromatographic analysis of bile for determination of the biliru-
Genetic lesions bin monoglucuronide to diglucuronide ratio can be used. As in
CN2, bile in Gilbert syndrome contains an increased proportion
Molecular genetic studies are consistent with autosomal reces-
(over 10%) of bilirubin monoglucuronide [82], reflecting
sive inheritance [83]. In CN2, UGT1A1 coding region mutations
reduced hepatic UGT1A1 activity in these syndromes.
always result in single amino acid transitions that significantly
Reduction of caloric intake to 400 kcal day−1 for 2 days or nico-
reduce the UGT1A1 activity, without completely abolishing it.
tinic acid administration [89] increase serum bilirubin levels in
In some cases, the mutation has been shown to increase the Km
Gilbert syndrome as well as in normal subjects, therefore these
for bilirubin [54]. The various lesions causing CN2 have been
tests do not provide definitive diagnosis. Needle biopsy of the
reviewed [29].
liver is not recommended for the diagnosis.

Genetic basis of Gilbert syndrome


Gilbert syndrome
Gilbert syndrome is associated with a varient TATAA box in the
Gilbert and Lereboullet described this common disorder, associ-
promoter upstream to exon 1 of UGT1A1. The normal TATAA
ated with a mild, fluctuating unconjugated hyperbilirubinemia
element has the sequence A[TA]6TAA, whereas subjects with
in 1901 [84]. More than a century later, Bosma and associates
Gilbert syndrome are homozygous for insertion of two nucleo-
discovered that this syndrome is caused by a polymorphism in
tides, making its sequence T[TA]7TAA, which reduces the
the proximal promoter region that reduces the expression of
expression of UGT1A1 to approximately 30% of normal [85].
UGT1A1 [85]. Gilbert syndrome is usually diagnosed in young
The Gilbert‐type TATAA element has been designated as
adults during blood tests for routine check up, screening, or
UGT1A1*28 [90]. Approximately 9% of most populations are
investigation of unrelated illnesses. In many cases, hyperbiliru-
homozygous for UGT1A1*28 and about 42% are heterozygous
binemia is intermittent and is usually less than 3 mg dL−1.
carriers. However, all subjects who are homozygous for the
Bilirubin levels increase during intercurrent illness, stress, fast-
UGT1A1*28 allele do not exhibit the clinical phenotype,
ing, or menstruation [86]. Mild icterus is the only positive clini-
indicating that other variables, particularly the rate of bilirubin
cal finding, and predominantly unconjugated hyperbilirubinemia
production may be necessary for manifestation of hyperbiliru-
is the only abnormality found on routine blood tests. Some
binemia. Thus, although the inheritance is autosomal (chromo-
patients report fatigue and abdominal discomfort, which may be
some 2q37), jaundice is uncomon in women, probably because
manifestations of anxiety or superimposed illnesses. Oral chol-
of lower daily bilirubin production. The gene frequency for
ecystrography visualizes the gallbladder normally, but there
UGT1A1*28 may be lower in Japan. Some UGT1A1 coding
may be a higher incidence of gallstones. Liver biopsy is not nec-
region mutations may cause mild hyperbilirubinemia, compati-
essary for diagnosis, but when performed, shows normal histol-
ble with the clinical diagnosis of Gilbert syndrome [91, 92].
ogy, except for some nonspecific lipofuscin accumulation in the
These mutations have been reported only in populations of East
centrilobular zone.
Asian origin. Some of these mutations were rported to be domi-
nant negative, suggesting that they reduce the activity of a nor-
Incidence
mal allele [93].
Gilbert syndrome is one of the most common inherited disorder Some heterozygous carriers of structural mutations (CN1 or
in humans, the reported incidence ranging from 3% to 7% of CN2) may also carry the Gilbert‐type promoter on the structur-
most populations [87]. As males have higher average serum bili- ally normal allele, which reduces the expression of the only
rubin levels than females, Gilbert syndrome is diagnosed more structurally normal allele. Such combinations may give rise to
frequency in males [87]. Gilbert syndrome is often recognized intermediate levels of hyperbilirubinemia [94, 95], explaining
around puberty, which may be related to increased red cell mass the long‐standing observation that intermediate levels of hyper-
and consequent increased bilirubin production, as well as inhi- bilirubinemia are common among relatives of patients with
bition of bilirubin glucuronidation by endogenous steroid CN1 or CN2. Because of the rarity of CN1 and CN2 mutations
hormones. and the very high frequency of the UGT1A1*28 allele, this type
of combination is a more common cause of intermediate levels
Diagnosis of hyperbilirubinemia than the presence of a coding region
mutation on both UGT1A1 alleles.
Gilbert syndrome is diagnosed in individuals with mild uncon-
jugated hyperbilirubinemia in the absence of normal liver serol-
Health implications of Gilbert syndrome
ogy and without evidence of hemolysis. Hepatic UGT1A1
activity is consistently low (around 30% of normal) in Gilbert Gilbert syndrome is innocuous, but its recognition is considered
syndrome [88]. Although hemolysis is not a feature of Gilbert important mainly for reassuring the patient and the physician
syndrome, coexistent hemolytic disorders, such as glucose 6‐ that no underlying liver disease is responsible for the mild
phosphate dehydrogenase deficiency can make jaundice clini- hyperbilirubinemia. In fact, mild elevation of serum bilirubin
cally obvious, thereby bringing the patient to the attention of a may have health benefits (see section on Potential beneficial
physician. Sequence determination of the TATAA element effects of bilirubin). However, reduced UGT1A1 activity can
20:  Disorders of Bilirubin Metabolism 237

affect detoxification of certain drugs [96]. Gilbert syndrome is usually made after puberty and, in some cases, during pregnancy
associated with a high incidence of diarrhea in patients treated or contraceptive use [105, 107].
with the anticancer drug irinotecan [97]. Oxidative paracetamol
metabolism is associated with drug toxicity. There is conflicting Laboratory tests
evidence in literature for increased oxidative metabolism of par-
Liver function tests, including serum bile acid levels, are normal
acetamol (acetaminophen) and its decreased glucuronidation in
[107]. Serum bilirubin levels usually range from 2 to 5 mg dL−1,
subjects with the UGT1A1*28 allele [98, 99]. Nilotinib and
but can rarely reach 20–25 mg dL−1. Over 50% of total serum
probably other tyrosine kinase inhibitors used in the treatment
bilirubin is direct‐reacting, and bilirubin is excreted in urine.
of leukemia are not metabolized by UGT1A1, but may inhibit
Because the hepatocellular canalicular excretion is abnormal for
the enzyme activity, thereby enhancing hyperbilirubinemia in
many organic anions, except bile acids, oral cholecystography,
patients with Gilbert syndrome, as well as CN2 [100].
even using a “double dose” of the contrast material, fails to visu-
alize the gallbladder. Gallbladder may be visualized 4 to 6 hours
Animal model
after intravenous administration of meglumine iodipamide
The Bolivian population of squirrel monkeys (Saimiri sciureus) (Biligrafin) [108]. Macroscopically, the liver is black, and light
has higher serum unconjugated bilirubin concentrations and a microscopy reveals a dense pigment. After intravenous infusion
greater hyperbilirubinemic response to fasting than does a of 3H‐epinephrine into mutant Corriedale sheep (an animal
closely related Brazilian population [101, 102]. Plasma clear- model for Dubin–Johnson syndrome), radioactivity is incorpo-
ance of intravenously administered bilirubin is lower in the rated into the dark brown pigment [109], which is not authentic
Bolivian squirrel monkey population. As in subjects with Gilbert melanin, but may consist of polymers of epinephrine metabo-
syndrome, hepatic UGT activity toward bilirubin, and bilirubin lites [110]. Following liver regeneration after hepatocellular
diglucuronide:monoglucuronide ratio in bile are lower in the apoptosis, such as after acute viral hepatitis, the pigment is
Bolivian population. Fasting hyperbilirubinemia is rapidly cleared from the liver and reaccumulates slowly after recovery.
reversed by oral or intravenous administration of carbohydrates,
but not lipids [101]. Organic anion transport
Organic anions other than bile acids are transported into the bile
canaliculus from the hepatocyte against a concentration gradi-
ent by an ATP‐dependent energy‐consuming process, mediated
DISORDERS PREDOMINANTLY by a protein that had been originally termed the canalicular mul-
ASSOCIATED WITH CONJUGATED tispecific organic anion transporter (cMOAT), and is now
HYPERBILIRUBINEMIA termed MRP2 or ABCC2 [111–114]. These anions include bili-
rubin glucuronides, the leukotriene LTC4, oxidized and reduced
Normally, by chromatographic measurement, approximately glutathione, and numerous glucuronide and glutathione conju-
4% of bilirubin in plasma is conjugated. The proportion of con- gates. In contrast, with some exceptions, bile acids are secreted
jugated bilirubin increases in plasma when bilirubin glucuron- normally. MRP2 is one of the ABC transporters [112]. Direct
ides formed inside hepatocytes are transferred back to plasma as evidence for its involvement in canalicular transport came from
a result of biliary obstruction, inflammatory or ischemic injury the discovery of a frameshift mutation in the gene encoding
of the liver, inherited disorders of bile canalicular excretion, or Mrp2 in the TR− rat [115]. Studies in TR− mice indicated that
failure of reuptake of bilirubin glucuronides transported out of some sulfated or glucuronidated bile acids require Mrp2 for
the hepatocytes into sinusoidal blood. In addition, there are sev- biliary excretion [116]. Consistent with this, in a multicenter
eral disorders, collectively termed progressive familial intrahe- study in Japan, neonates with Dubin–Johnson syndrome had
patic cholestasis, that are caused by inherited abnormalities of significantly increased serum total bile acid levels [117].
bile canalicular transport proteins or a tight junction protein. In Despite the lack of MRP2 function, the serum bilirubin is
another genetic condition, paucity of bile ducts leads to conju- only mildly elevated in Dubin–Johnson syndrome, suggesting
gated hyperbilirubinemia. A brief description of these disorders the presence of additional mechanisms of biliary excretion of
follows. bilirubin conjugates. In addition, accumulation of organic ani-
ons within the hepatocytes due to MRP2 deficiency leads to
upregulation of the expression of MRP1 and MRP3 in the baso-
Dubin–Johnson syndrome
lateral surface of the hepatocytes. These and possibly additional
Dubin and Johnson [103] and Sprinz and Nelson [104] described ATP‐consuming pumps may actively export both unconjugated
a syndrome characterized by chronic conjugated hyperbiliru- and conjugated bilirubin from the hepatocyte to plasma via the
binemia and grossly pigmented, but otherwise histologically space of Disse [118].
normal liver. Mild icterus is the only consistent physical finding, After intravenous injection of the organic anion bromosul-
but rarely, hepatosplenomegaly has been observed [105, 106]. fophthalein (BSP), plasma BSP concentration decreases at near‐
Patients are usually asymptomatic, but an occasional patient normal rate for 45 minutes, indicating normal uptake across the
complains of weakness and vague abdominal pain. Serum bile sinusoidal surface of hepatocytes and normal storage capacity
acid levels are nearly normal [105], and pruritus is absent. of hepatocytes. However, in 90% of patients, plasma BSP con-
Serum bilirubin levels are increased during intercurrent illness, centration exhibits a secondary increase, so that plasma BSP
intake of oral contraceptives, and pregnancy [105]. Diagnosis is concentration at 90 minutes exceeds that at 45 minutes. This
238 THE LIVER:  DISORDERS PREDOMINANTLY ASSOCIATED WITH CONJUGATED HYPERBILIRUBINEMIA

secondary rise is due to reflux of glutathione‐conjugated BSP conjugates [129] and many other organic anions [130],
from hepatocytes into the circulation [119]. A similar secondary Coproporphyrin I is the major porphyrin isomer excreted in
rise occurs after intravenous administration of bilirubin. urine [131]. The liver is not normally pigmented, but intracel-
However, such secondary rise of plasma BSP can also occur in lular pigments are deposited upon feeding a diet enriched in
some other hepatobiliary disorders [120], therefore, this phe- tryptophan, tyrosine, and phenylalanine. Impaired excretion of
nomenon is not pathognomonic of Dubin–Johnson syndrome. anionic metabolites of these amino acids may result in their
retention, oxidation, polymerization, and subsequent lysosomal
Urinary coproporphyrin excretion accumulation in hepatocytes [132]. Experiments in these rat
models showed that the pathway for secretion of bilirubin con-
Coproporphyrins excreted in urine consists of two isomers,
jugates and many other organic anions into the bile canaliculus
coproporphyrin I and coporophyrin III. Normally, in adults
differed from that for secretion of most bile acids. Bile acids
approximately 75% of the urinary coproporphyrin is isomer III,
with free 3‐OH groups are transported by the bile salt export
which is the precursor of heme. Total urinary coproporphyrin
pump (BSEP), but sulfated or glucuronide‐conjugated bile acids
excretion is normal in Dubin–Johnson syndrome, but over 80%
are transported by Mrp2 [133].
is isomer I [121]. Neonates have a higher proportion of urinary
The golden lion tamarin monkey (Leontopithecus rosalia),
coproporphyrin I than adults, but the proportion is not as high as
which manifests elevated serum conjugated bilirubin levels, is a
in Dubin–Johnson syndrome. The relationship of the abnormal
non‐human primate model of Dubin–Johnson syndrome [134].
pattern of urinary porphyrin excretion to the organic anion
transport defect is not fully understood. When correlated with
history and physical examination, the urinary coproporphyrin Rotor syndrome
excretion pattern is diagnostic of Dubin–Johnson syndrome.
Rotor, Manahan, and Florentin reported two families with sev-
eral patients with lifelong predominantly conjugated hyperbili-
Genetic basis and inheritance rubinemia without evidence of hemolysis [135]. Other routine
Dubin–Johnson syndrome is inherited as an autosomal recessive blood biochemistries and hematologic tests were normal and, in
trait, and has been reported in all races and both sexes. It is rare contrast to Dubin–Johnson syndrome, there was no pigmenta-
in all races, except in Iranian, Iraqi, and Moroccan Jews living tion in the liver. Liver histology was normal. Rotor syndrome is
in Israel, who have an incidence of 1 in 1300 [106], in whom it harmless [135]. Although rare, it has been described in several
is associated with clotting factor VII deficiency [122]. Dubin– races.
Johnson syndrome may be relatively common in some areas of
Japan, where consanguinity is frequent. Organic anion excretion
Dubin–Johnson syndrome is caused by insertions, deletions, The organic anion excretion defect in Rotor syndrome is differ-
and nonsense mutations of the MRP2 (ABCC2) gene, or abnor- ent from that in Dubin–Johnson syndrome. As in many acquired
mal splicing of the RNA transcript that causes loss of MRP2 liver diseases, over 25% of injected BSP is retained in serum at
expression, or missense mutations of the gene that interfere with 45 minutes, and there is no secondary rise of plasma BSP level
normal localization of MRP2 in the bile canalicular plasma [136]. Plasma clearance of intravenously administrated uncon-
membrane of hepatocytes [115, 117, 123, 124]. Human MRP2 jugated bilirubin and indocyanine green is also delayed. In con-
gene has been localized to chromosome 10q23–q24 [125]. trast to the findings in Dubin–Johnson syndrome, bile canalicular
Single amino acid transitions most commonly involve the cru- export pumps are normal in Rotor syndrome, therefore, the gall-
cial ATP‐binding region. Some mutations may lead to impaired bladder is visualized by oral cholecystography [137, 137A].
glycosylation of MRP2, leading to premature proteasome‐
dependent degradation [126].
Urinary coproporphyrin excretion

Animal models Total urinary coproporphyrin is increased two‐ to fivefold over


normal in Rotor syndrome, of which approximately 65% is iso-
In mutant Corriedale sheep, biliary excretion of conjugated biliru- mer I [138]. These findings are similar to those seen in many
bin, glutathione‐conjugated BSP, iopanoic acid, and indocyanine other hepatobiliary disorders, and distinguish Rotor syndrome
green is decreased, whereas the transport of taurocholate [127] from Dubin–Johnson syndrome. However, two brothers with
and unconjugaed BSP [128] is normal. The secretion of organic clinical features of Rotor syndrome were reported to have over
cations, such as procaine amide ethobromide, is unaffected. 80% of urinary coproporphyrins as isomer I [139], raising doubts
Serum bilirubin is mildly increased, with 60% of the bilirubin about the usefulness of using urinary coproporphyrin analysis in
being conjugated. Other than dark brown pigmentation, liver his- distinguishing the two syndromes. More recently, the discovery
tology is normal [109]. As in human Dubin–Johnson syndrome, of the genetic basis of Rotor syndrome has clarified the defect in
total urinary coproporphyrin excretion is normal, but the propor- organic anion handling by hepatocytes in this disorder (see later).
tion of coproporphyrin I is increased. Thus, the mutant Corriedale
sheep is a model of human Dubin–Johnson syndrome.
Molecular basis of Rotor syndrome
As in humans with Dubin–Johnson syndrome, Eisai hyper-
bilirubinemic (EHBR) and TR− rats lack Mrp2 and exhibit Van de Steeg et  al. found a tight linkage between simultaneous
reduced biliary excretion of conjugated bilirubin, leukotriene mutations in the SLC01B1 and SLC01B3 genes and Rotor syn-
LTC4 and glutathione, as glucuronic acid and glutathione drome in six families [140]. These mutations result in deficiency
20:  Disorders of Bilirubin Metabolism 239

of the organic anion transporting polypeptides OATP1B1 and the secretion of bile acids or other components of bile across the
OATP1B3, respectively. Using mice deficient in the multispecific bile canaliculi of hepatocytes [145]. Benign recurrent intrahe-
sinusoidal export pump Abcc3, as well as Oatp1a/1b, van de Steeg patic cholestasis (BRIC) is a disorder that is genetically related
et  al. demonstrated that Abcc3 secretes bilirubin glucuronides to PFIC1 or PFIC2. The PFIC syndromes usually present during
from hepatocytes to blood, whereas Oatp1a/1b mediate their reup- infancy or childhood, and often lead to growth failure and pro-
take. Because of the efficient uptake and conjugation of bilirubin gressive liver disease. Several additional genetic cholestatic
by zone 1 hepatocytes from blood entering the hepatic sinusoids, disorders have been described during the last few years. PFIC
these bilirubin glucuronides formed in these cells may exceed syndromes and the other inherited cholestatic diseases have been
their canalicular excretory capacity. A portion of bilirubin glucuro- reviewed recently [146] and are discussed in brief here.
nides formed in the zone 1 hepatocytes is transported back into the
sinusoidal blood and is subsequently taken up by hepatocytes Progressive familial intrahepatic cholestasis type 1
located downstream to the blood flow (toward Zzone 3) via
Oatp1b1/3. By recruiting additional hepatocytes, this mechanism PFIC1 was originally described in an Amish–Mennonite family
increases the capacity of liver to handle a bilirubin load. and was named Byler disease, after the family name of the first
In humans, OATP1B1 and OATP1B3 belong to the 11‐member patient [147]. PFIC1 is associated with severe life‐threatening
OATP family that have 12 membrane‐spanning domains and cholestasis. It is caused by mutations in the P‐type ATPase gene
mediate the uptake of a wide variety of compounds. They are FIC1 (also termed ATP8B1), which is located on chromosome
almost exclusively localized in the sinusoidal domain of hepato- 18q21 [148]. FIC1‐mediated ATP hydrolysis is coupled with the
cyte plasma membranes. OATP1B1 and OATP1B3 transport dif- translocation of acidic phospholipids. How FIC1 mutation
ferent, but partly overlapping substrates. In addition to bilirubin causes cholestasis is not fully understood [149]. FIC1 defi-
glucururonides, substrates for both transporters include drugs such ciency may interfere with the nuclear translocation of the
as hydroxymethylglutaryl (HMG)‐CoA reductase inhibitors nuclear receptor FXR [150], thereby downregulating the expres-
(statins), angiotensin II receptor blockers (sartans), angiotensin‐ sion of the bile salt export protein (BSEP) (see section on
converting enzyme (ACE) inhibitors, and antidiabetes (glinides) Progressive familial intrahepatic cholestasis type 2).
as substrates [141]. In addition to organic anions, OATP1B1 and
OATP1B3 accept neutral compunds, such as digoxin, ouabain, Benign recurrent intrahepatic cholestasis
lopinavir, as well as zwitter ionic drugs such as fexofenadine. BRIC was first described in 1959 [151]. It presents in adoles-
OATP1B1 preferentially recognizes estrone‐3‐sulfate , whereas cence or early adulthood with recurrent episodes of cholestasis,
cholecystokinin octapeptide CCK‐8, telmisartan, paclitaxel, and characterized by conjugated hyperbilirubinemia, together with
docetaxel are specifically transported via OATP1B3. OATPs are malaise, anorexia, pruritus, weight loss, and malabsorption.
structurally unrelated to ABCC2 (MRP2), which is located in the During the attacks, which last weeks to months, laboratory tests
bile canalicular domain of hepatocyte plasma membranes, but reveal biochemical evidence of cholestasis without severe hepa-
they share many of their substrates, including bilirubin glucuron- tocellular injury [152–154]. These attacks are followed by a
ides, which coordinates hepatocellular uptake with and canalicular complete clinical, biochemical, and histological remission. The
excretion. For certain substrates, OATP1B1 and OATP1B3 cannot clinical features and duration of the episodes in a given patient
compensate for the lack of each other. For example, simvastatin‐ resemble those in previous attacks. Liver biopsy shows non-
associated myopathy is strongly linked to a single nucleotide poly- inflammatory intrahepatic cholestasis without fibrosis, not-
morphism of SLC01B1 [142]. On the other hand, for the reuptake withstanding the number and severity of the attacks. During
of bilirubin glucuronides, OATP1B1 and OATP1B3 can compen- remission, light or electron microscopy of the liver tissue returns
sate for the lack of one another, whereby simultaneous mutation of to normal [155]. Like PFIC1, the inheritance shows an autoso-
both is required for the manifestation of Rotor syndrome. mal recessive pattern. Interestingly, this relatively benign disor-
der is also caused by certain missense mutations in the FIC1
Animal model gene, other mutations of which cause the much more severe
A subgroup of Southdown sheep exhibits mild conjugated ­disorder PFIC1.
hyperbilirubinemia and photosensitivity [143]. Southdown There is no specific treatment for BRIC. Some cases are asso-
sheep with these clinical features were found to have a missense ciated with mutations of ABCB11, which is associated with
(glycine to arginine) mutation in Slc01b3 [144]. This glycine PFIC2. Because of this, some authors classify BRIC into BRIC‐I
residue is conserved in seven other mammalian species and is and BRIC‐II.
thought to be functionally critical.
Progressive familial intrahepatic cholestasis type 2
This disorder, which resembles Byler disease clinically, occurs in
INHERITED CHOLESTASIS SYNDROMES mainly in Middle Eastern and European races. It is caused by
defects of the ABCB11 gene that encodes the bile salt export
pump (BSEP), previously termed sister P‐glycoprotein. BSEP is
Progressive familial intrahepatic cholestasis a canalicular ATP‐dependent transporter of bile acids from the
Three life‐threatening disorders, collectively termed progressive hepatocytes into the bile [156]. The ABCB11 gene is located on
familial intrahepatic cholestasis (PFIC1, PFIC2, PFIC3, and chromosome 2q24 [157]. A large number of different point muta-
PFIC4), can cause cholestasis of varying severities by affecting tions in ABCB11 have been described in patients with PFIC2.
240 THE LIVER:  ACKNOWLEDGMENT

Although both PFIC1 and PFIC2 are associated with life‐ OTHER INHERITED CHOLESTATIC
threatening cholestasis, serum γ‐glutamyl transpeptidase (GGT) DISORDERS
levels are normal or nearly so in both disorders, which differen-
tiates them from PFIC3 [145, 157]. Mutation of CIRH1A, the gene encoding a mitochondrial scaf-
BSEP is expressed specifically in the liver and, as expected, fold protein, has been reported to manifest as cholangiopathy of
liver transplantation ameliorates all manifestations of PFIC2. North American Indian childhood cirrhosis. Clinically, this
However, in several cases, PFIC‐like symptoms recur after liver disorder resembles extrahepatic biliary atresia, but there is no
transplantation consequent to development of immunoglobulin evidence of biliary tract obstruction [168].
G‐type antibodies against BSEP. Although antibodies develop
against both the N‐terminal and the C‐terminal regions of the
protein, only sera recognizing the first extracellular loop (ECL1) GRACILE syndrome
inhibit transepithelial taurocholate transport [158].
Genetic lesions of another mitochondrial scaffold gene,
Progressive familial intrahepatic cholestasis type 3 BCS1L, presents potentially lethal intrahepatic cholestasis
associated with fetal growth retardation, amino aciduria, iron
PFIC3 involves mutations in ABCB4, an ATP‐utilizing bile can- overload, and lactic acidosis (GRACILE) [169]. Interestingly,
alicular pump (also known as the multidrug resistance protein‐3 neither North American Indian childhood cirrhosis nor
P‐glycoprotein (PGY3, MDR3). ABCB4 transports phosphati- GRACILE exhibit the usual features of mitochondriopathy,
dylcholine from the inner lipid leaflet of the bile canaliculus to such as central nervous system (CNS) lesions or hepatocellu-
the outer leaflet, thereby replenishing the phospholipid in the lar steatosis.
outer leaflet, which is continuously removed by contact with
bile acids. Failure of phosphatidylcholine translocation, the
canalicular membrane, and small bile ducts are chronically Alagille syndrome
injured by bile salts [158, 159]. In contrast to PFIC1 and PFIC2,
in which the canalicular membrane is protected by reduced Alagille syndrome affects the development of multiple organ sys-
secretion of bile salts in the canalicular bile, serum GGT activity tems, including liver, heart, kidneys, eyes, vertebrae (sagittal,
is increased in PFIC3 [145]. Human ABCB4 deficiency can pre- cleft), and CNS [170]. Paucity of bile ducts is found in 89% of
sent with a variety of hepatobiliary disorders, including small‐ cases [171]. Most patients exhibit a characteristic facies. Although
duct primary sclerosing cholangitis [160] and cholesterol the CNS lesions can be the result of malnutrition, even after care-
gallstones [161]. Interestingly, asymptomatic female heterozy- ful nutritional balance, intracranial bleeding is the most common
gous carriers of nonsense or missense mutations of the ABCB4 CNS complication [171]. Alagille syndrome is inherited as an
gene have been reported to manifest familial intrahepatic chol- autosomal dominant characteristic with markedly variable pene-
estasis of pregnancy [162, 163], the cause of which was trance. Lesions of the Jagged1 (JAG1) gene, on chromosome
unknown heretofore. 20p12 [172] are responsible for this disease. Jagged‐1 is a ligand
Liver transplantation is the only treatment option for PFIC3 of Notch, and is important in the Notch signaling pathway, which
available at this time. Transplantation of normal hepatocytes in is critical in organ development. Coding region mutations of
a knockout mouse model of PFIC3 has resulted in spontaneous JAG1 have been identified in about 70% of the patients with
massive liver repopulation with the transplanted cells, resulting Alagille syndrome. Surprisingly, 50–70% of these mutations are
in long‐term amelioration of the phospholipid transport defect de novo, and not found in the parents. About 21–50% of patients
[164, 165]. with Alagille syndrome who manifest hepatic symptoms in
infancy eventually require liver transplantation.
Progressive familial intrahepatic cholestasis type 4
As discussed earlier, PFIC syndromes can be classified clinically Villin disease
by serum GGT levels that reflect bile salt‐induced injury of bile
canaliculi or smallest bile ductules. In PFIC1 and PFIC2 serum Villin is a tissue‐specific actin‐modifying protein, critical in
GGT is normal, whereas in PFIC3 it is elevated. In 2014, the bundling, nucleation, capping, and severing of actin fila-
Sambrota et al. [166] reported a multi‐institution collaborative ments [173]. In vertebrates, villin is expressed at the apical
study of 33 children from 29 families who had severe chronic surface of epithelial cells and is a major structural component
cholestatic liver disease, with low GGT for the degree of choles- of the brush border cytoskeleton. Several children with fea-
tasis. However, these children did not have mutations of either tures of cholestasis, who later developed cirrhosis of the liver
ABCB11 and ATP8B1, which cause PFIC1 and PFIC2, respec- requiring liver transplantation, were found to have lesions of
tively. In these 29 families, 18 parents were consanguineous. the VILLIN1 gene [174].
Genetic analysis by targeted resequencing (TRS), whole‐exome
sequencing (WES), or both revealed protein‐truncating muta-
tions in the tight junction protein 2 gene (TJP2), which resulted
in failure of incorporation of TJP2 protein in desmosomes delim- ACKNOWLEDGMENT
iting the bile canaliculi. PFIC resulting from TJP2 mutations has
been designated PFIC4. Two cases of hepatocellular carcinoma This work was supported in part by NIH grants RO1‐DK‐092469,
in infants with this disorder have been reported [167]. RO1‐DK 100490 and P30‐DK‐41296.
20:  Disorders of Bilirubin Metabolism 241

REFERENCES 25. Nakamura, H., Yonetani, M., Uetani, Y., Funato, M., and Lee, Y.
Determination of serum unbound bilirubin for prediction of kernicterus in
low birthweight infants. Acta Paediatr Jpn, 1992;34: 642–7.
1. Kappas, A. and Drummond, G.S. Direct comparison of Sn‐mesoporphyrin,
26. Wang, P., Kim, R.B., Roy‐Chowdhury, J., and Wolkoff, A.W. The human
an inhibitor of bilirubin production, and phototherapy in controlling hyper-
organic anion transport protein SLc21A6 is not sufficient for bilirubin trans-
bilirubinemia in term and near‐term newborns. Pediatrics, 1995;95:468.
port. J Biol Chem, 2003;278:20695.
2. Valaes, T., Petmezaki, S., Henschke, C., Drummond, G.S., and Kappas, A.
27. Bosma, P.J., Seppen, J., Goldhoorn, B. et al. Bilirubin UDP‐glucuronosyl-
Control of jaundice in preterm newborns by an inhibitor of bilirubin produc-
transferase 1 is the only relevant bilirubin glucuronidating isoform in
tion: studies with tin‐mesoporphyrin. Pediatrics, 1994;93:1.
man. J Biol Chem, 1994;269:17960.
3. Westlake, D.W.S., Roxburgh, J.M., and Talbot, G. Microbial production of
28. Ritter, J.K., Chen, F., Sheen, Y.Y. et  al. A novel complex locus UGT1
carbon monoxide from flavonoids. Nature, 1961;189:510.
encodes human bilirubin, phenol, and other UDP‐glucuronosyltransferase
4. Itoh, K., Chiba, T., Takahashi, S. et al. An Nrf2/small Maf heterodimer medi-
isozymes with identical carboxyl termini. J Biol Chem, 1992;267:3257.
ates the induction of phase II detoxifying enzyme genes through antioxidant
29. Kadakol, A., Ghosh, S.S., Sappal, B.S. et  al. Genetic lesions of bilirubin
response elements. Biochem Biophys Res Commun, 1997;236(2):313–22.
uridinediphosphoglucuronate glucuronosyltransferase causing Crigler‐
5. Naito, Y., Takagi, T., Uchiyama, K., and Yoshikawa, T. Heme oxygenase‐1: a
Najjar and Gilbert’s syndromes: correlation of genotype to phenotype. Hum
novel therapeutic target for gastrointestinal diseases. J Clin Biochem Nutr,
Mutat, 2000;16:297.
2011;48:126.
30. Gantla, S., Bakker, C.T.M., Deocharan, B. et  al. Splice site mutations: a
6. Yachie, A., Niida, Y., Wada, T. et  al. Oxidative stress causes enhanced
novel genetic mechanism of Crigler‐Najjar syndrome type 1. Am J Hum
endothelial cell injury in human heme oxygenase‐1 deficiency. J Clin Invest,
Genet, 1998;62:585.
1999;103:129–35.
31. Bosma, P.J., Roy Chowdhury, J., Bakker, C. et al. A sequence abnormality in
7. Mougiakakos, D., Jitschin, R., Johansson, C.C. et al. The impact of inflam-
the promoter region results in reduced expression of bilirubin‐UDP‐glucu-
matory licensing on heme oxygenase‐1‐mediated induction of regulatory T
ronosyltransferase‐1 in Gilbert syndrome. N Engl J Med, 1995;333:1171.
cells by human mesenchymal stem cells. Blood, 2011;117:4826–35.
32. van de Steeg, E., Stranecky, V., Hartmannova, H. et al. Complete OATP1B1
8. Basuroy, S., Bhattacharya, S., Tcheranova, D. et al. HO‐2 provides endogenous
and OATP1B3 deficiency causes human Rotor syndrome by interrupting
protection against oxidative stress and apoptosis caused by TNF‐α in cerebral
conjugated bilirubin reuptake into the liver. J Clin Invest, 2012;122:519.
vascular endothelial cells. Am J Physiol Cell Physiol, 2006;291:C897–C908.
33. Huang, W., Zhang, J., Chua, S.S. et al. Induction of bilirubin clearance by the
9. Chowdhury, R., Roy Chowdhury, N., and Arias, I.M. Bilirubin conjugation
constitutive androstane receptor. Proc Natl Acad Sci U S A, 2003;100:4156.
in the spiny dogfish Squalus acanthias, the small skate Raja erinacea and the
34. Xie, W., Yeuh, M.F., Radominska‐Pandya, A. et al. Control of steroid, heme,
winter flounder Pseudopleuronectes americanus. Comp Biochem Physiol,
and carcinogen metabolism by nuclear pregnane X receptor and constitutive
1980;668:523.
androstane receptor. Proc Natl Acad Sci U S A, 2003;100:4150.
10. O’Brien, L., Hosick, P.A., John, K., Stec, D.E., and Hinds, T.D., Jr. Biliverdin
35. Roy‐Chowdhury, J., Locker, J., and Roy‐Chowdhury, N. Nuclear receptors
reductase isozymes in metabolism. Trends Endocrinol Metab, 2015;26:212–20.
orchestrate detoxification pathways. Dev Cell, 2003;4:607.
11. Parkar, M., Jeremiah, S.J., Povey, S. et al. Confirmation of the assignment of
36. Maisels, M.J. Managing the jaundiced newborn: a persistent challenge. Can
human biliverdin reductase to chromosome 7. Ann Human Genet, 1984;48:
Med Assoc J, 2015;187:335–43.
57–60.
37. American Academy of Pediatrics Subcommittee on Hyperbilirubinemia.
12. Fu, G., Liu, H., and Doerksen, R.J. Molecular modeling to provide insight
Management of hyperbilirubinemia in the newborn infant 35 or more weeks
into the substrate binding and catalytic mechanism of human biliverdin‐
of gestation. Pediatrics, 2004;114:297–316. Erratum October 01, 2004.
IXalpha reductase. J Phys Chem B, 2012;116:9580–94.
38. Bhutani, V.K., Stark, A.R., Lazzeroni, L.C. et al. Initial clinical testing evalua-
13. Hannam, S., Moriaty, P., O’Reilly, H. et al. Normal neurological outcome in
tion and risk assessment for universal screening for Hyperbilirubinemia
two infants treated with exchange transfusions born to mothers with Crigler‐
Screening Group. Predischarge screening for severe neonatal hyperbilirubine-
Najjar Type 1 disorder. Eur J Pediatr, 2009;168:427–9.
mia identifies infants who need phototherapy. J Pediatr, 2013;162:477–82.
14. Bonnet, R.J., Davis, E., and Hursthouse, M.B. Structure of bilirubin. Nature,
39. Maisels, M.J., Pathak, A., Nelson, N.M. et  al. Endogenous production of
1976;262:326.
carbon monoxide in normal and erythroblastic newborn infants. J Clin
15. Van den Bergh, A.A.H. and Muller, P. Uber eine direkte und eine indirekte
Invest, 1971;50:1.
Diazoreaktion auf Bilirubin. Biochem Z, 1916;77:90.
40. Vest, M., Strebel, L., and Hauensiein, D. The extent of “shunt” bilirubin and
16. Choi, S.H., Yun, K.E., and Choi, H.J. Relationships between serum total
erythrocyte survival in the newborn infant measured by the administration of
bilirubin levels and metabolic syndrome in Korean adults. Nutr Metab
(15N) glycine. Biochem J, 1965;95:11c.
Cardiovasc Dis, 2013;23:31.
41. Clarke, C.A., Donohoe, W.T.A., Finn, R. et al. Prevention of Rh hemolytic
17. Breimer, L.H., Wannamethee, G., Ebrahim, S., and Shaper, A.G. Serum bili-
disease: Final results of the “high risk” clinical trial. A combined study from
rubin and risk of ischemic heart disease in middle‐aged British men. Clin
centers in England and Baltimore. Br Med J, 1971;2:607.
Chem, 1995;41:1504.
42. Levi, A.J., Gatmaitan, Z., and Arias, I.M. Deficiency of hepatic organic
18. Temme, E.H.N., Zhang, J., Schouten, E.G. et al. Serum bilirubin and 10‐year
anion‐binding protein, impaired organic anion uptake by liver and “physio-
mortality risk in a Belgian population. Cancer Causes Control, 2001;12:887.
logic” jaundice in newborn monkeys. N Engl J Med, 1970;283:1136.
19. Zucker, S.D., Horn, P.S., and Sherman, K.E. Serum bilirubin levels in the US
43. Kawade, N. and Onishi, S. The prenatal and postnatal development of
population: gender effect and inverse correlation with colorectal cancer.
UDP‐glucuronosyltransferase activity towards bilirubin and the effect of
Hepatology, 2004;40:827.
premature birth on this activity in human liver. Biochem J, 1981;196:257.
20. Torgerson, J.S., Lindroos, A.K., Sjostrom, C.D. et  al. Are elevated ami-
44. Arthur, L.J., Bevan, B.R., and Holton, J.B. Neonatal hyperbilirubinemia and
notransferases and decreased bilirubin additional characteristics of the meta-
breast feeding. Dev Med Child Neurol, 1966;8:279.
bolic syndrome? Obesity Res, 1997;5:105–14.
45. Arias, I.M., Gartner, L.M., Seifter, S., and Furman, M. Prolonged neonatal
21. Hinds, T.D. Jr. Sodhi, K., Meadows, C. et al. Increased HO‐1 levels amelio-
unconjugated hyperbilirubinemia associated with breast feeding and a ster-
rate fatty liver development through a reduction of heme and recruitment of
oid, pregnane‐3α, 20β‐diol, in maternal milk that inhibits glucuronide forma-
FGF21. Obesity, 2014;22:705–12.
tion in vitro. J Clin Invest, 1964;42:2037.
22. Dong, H., Huang, H., Yun, X. et al. Bilirubin increases insulin sensitivity in
46. Maisels, M.J. and Newman, T.B. Kernicterus in otherwise healthy breast‐fed
leptin‐receptor deficient and diet‐induced obese mice through suppression of
term newborns. Pediatrics, 1995;96:730.
ER stress and chronic inflammation. Endocrinology, 2014;155:818–28.
47. Foliot, A., Ploussard, J.P., Housett, E. et al. Breast milk jaundice: In vitro
23. Amin, S.B., Saluja, S., Saili, A. et al. Chronic auditory toxicity in late pre-
inhibition of rat liver bilirubin‐uridine diphosphate glucuronosyltransferase
term and term infants with significant hyperbilirubinemia. Pediatrics,
activity and Z protein‐bromosulfophthalein binding by human breast milk.
2017;(4):pii:e20164009.
Pediatr Res, 1976;10:594.
24. Ahlfors, C.E., Vreman, H.J., Wong, R.J. et  al. Effects of sample dilution,
48. Lucey, J.F. and Driscol, J.J. Physiological jaundice re‐examined, in
peroxidase concentration, and chloride ion on the measurement of unbound
Kernicterus (ed. A. Sass‐Kortsak), University of Toronto Press, Toronto,
bilirubin in premature newborns. Clin Biochem, 2007;40:261–7.
1961.
242 THE LIVER:  REFERENCES

49. Nies, A.T., Gatmaitan, Z., and Arias, I.M. ATP‐dependent phosphatidylcho- 73. Galbraith, R.A., Drummond, G.S., and Kappas A. Suppression of bilirubin
line translocation in rat liver canalicular plasma membrane vesicles. J Lipid production in the Crigler‐Najjar, type I syndrome: studies with heme oxyge-
Res, 1996;37:1125. nase inhibitor tin‐mesoporphyrin. Pediatrics, 1992;89:175.
50. Poland, R.L. and Odell, G.B. Physiologic jaundice: the enterohepatic circu- 74. Fox, I.J., Roy Chowdhury, J., Kaufman, S.S. et  al. Treatment of Crigler‐
lation of bilirubin. N Engl J Med, 1971;284:1 Najjar syndrome type I with hepatocyte transplantation. N Engl J Med,
51. Crigler, J.F. and Najjar, V.A. Congenital familial non‐hemolytic jaundice 1998;333:1422.
with kernicterus. Pediatrics, 1952;10:169. 75. Iansante, V., Mitry, R.R., Filippi, C., Fitzpatrick, E., and Dhawan, A. Human
52. Childs, B., Sidbury, J.B., and Migeon, C.J. Glucuronic acid conjugation by hepatocyte transplantation for liver disease: current status and future per-
patients with familial non‐hemolytic jaundice and their relatives. Pediatrics, spectives. Pediatr Res, 2018;83:232–40.
1959;23:903. 76. Zhou, H., Dong, X., Kabarriti, R. et al. Single liver lobe repopulation with
53. Berk, P.D., Martin, F., Blaschke, T.F. et al. Unconjugated hyperbilirubine- wildtype hepatocytes using regional hepatic irradiation cures jaundice in
mia: physiological evaluation and experimental approaches to therapy. Ann Gunn rats. PLoS One, 2012;7:e46775.
Intern Med, 1975;82:552. 77. Yannam, G.R., Han, B., Setoyama, K. et al. A nonhuman primate model of
54. Seppen, J., Bosma, P.J., Goldhoorn, B.G. et  al. Discrimination between human radiation‐induced venocclusive liver disease and hepatocyte injury.
Crigler‐Najjar type I and II by expression of mutant bilirubin uridine diphos- Int J Radiat Oncol Biol Phys, 2014;88:404–11.
phate‐glucuronosyltransferase. J Clin Invest, 1994;94:2385. 78. Soltys, K.A., Setoyama, K., Tafaleng, E.N. et  al. Host conditioning and
55. Wolkoff, A.W., Roy Chowdhury, J., Gartner, L.A. et al. Crigler‐Najjar syn- rejection monitoring in hepatocyte transplantation in humans. J Hepatol,
drome (type I) in an adult male. Gastroenterology, 1979;76:3380. 2016;66:987–1000.
56. Mitchell, E., Ranganathan, S., McKiernan, P. et  al. Hepatic parenchymal 79. Arias, I.M. Chronic unconjugated hyperbilirubinemia without overt signs of
injury in Crigler‐Najjar type I. J Pediatr Gastroenterol Nutr, 2018;66: hemolysis in adolescents and adults. J Clin Invest, 1962;41:2233.
588–94. 80. Gollan, J.L., Huang, S.M., Billing, B., and Sherlock, S. Prolonged survival in
57. Petit, F.M., Gajdos, V., Parisot, F. et al. Parental isodisomy for chromosome three brothers with severe type II Crigler‐Najjar syndrome: ultrastructural
2 as the cause of Crigler‐Najjar syndrome type I syndrome. Eur J Hum and metabolic studies. Gastroenterology, 1975;68:1543.
Genet, 2005;13:278–82. 81. Arias, I.M., Gartner, L.M., Cohen, M. et al. Chronic nonhemolytic unconju-
58. Van Es, H.H.G., Goldhoorn, B., Paul‐Abrahamse, M. et al. Immunochemical gated hyperbilirubinemia with glucuronosyltransferase deficiency: clinical,
characterization of UDP‐glucuronosyltransferase in four patients with the biochemical, pharmacologic, and genetic evidence for heterogeneity. Am J
Crigler‐Najjar type I syndrome. J Clin Invest, 1990;85:1199. Med 1969;47:395.
59. Ritter, J.K., Crawford, J.M., and Owens, I.S. Cloning of two human liver 82. Fevery, J., Blanckaert, N., Heirwegh, K.P.M. et al. Unconjugated bilirubin
bilirubin‐UDP‐glucuronosyltransferase cDNAs with expression in COS‐1 and an increased proportion of bilirubin monoconjugates in the bile of
cells. J Biol Chem, 1991;266:1043. patients with Gilbert’s syndrome and Crigler‐Najjar syndrome. J Clin Invest,
60. Bosma, P.J., Roy Chowdhury, N., Goldhoorn, B.G. et al. Sequence of exons 1977;60:970.
and the flanking regions of human bilirubin‐UDP‐glucuronosyltransferase 83. Bosma, P.J., Golhoorn, B., Oude Elferink, R.P. et al. A mutation in bilirubin
gene complex and identification of a genetic mutation in a patient with uridine 5′‐diphosphate glucuronosyltransferase isoforms 1 causing Crigler‐
Crigler‐Najjar syndrome, type I. Hepatology, 1992;15:941–7. Najjar syndrome type II. Gastroenterology, 1993;105:216.
61. Bosma, P.J., Seppen, J., Goldhoorn, B. et al. Bilirubin UDP‐glucuronosyl- 84. Gilbert, A. and Lereboullet, P. La cholamae simple familiale. Semin Med,
transferase 1 is the only relevant bilirubin glucuronidating isoform in man. 1901;21:241.
J Biol Chem, 1994;269:17960. 85. Bosma, P.J., Roy Chowdhury, J., Bakker, C. et al. A sequence abnormality in
62. Servedio, V., d’Apolito, M., Maiorano, N. et al. Spectrum of UGT1A1 muta- the promoter region results in reduced expression of bilirubin‐UDP‐glucu-
tions in Crigler‐Najjar (CN) syndrome patients: identification of twelve ronosyltransferase‐1 in Gilbert syndrome. N Engl J Med, 1995;333:1171.
novel alleles and genotype‐phenotype correlation. Hum Mutat, 2005;25:325. 86. Thompson, R.P.H. (1981) Genetic transmission of Gilbert’s syndrome, in
63. Strauss, K.A., Robinson, D.L., Vreman, H.J. et al. Management of hyperbili- Familial Hyperbilirubinemia (ed. L. Okolicsanyl), Wiley, New York, 1981,
rubinemia and prevention of kernicterus in 20 patients with Crigler‐Najjar p. 91.
disease. Eur J Pediatr, 2006;165:306–19. 87. Powell, L.W., Hemingway, E., Billing, B.H., and Sherlock, S. Idiopathic
64. Kadakol, A., Deocharan, B., Mukhopadhyay, L. et al. Rapid prenatal diagno- unconjugated hyperbilirubinemia (Gilbert’s syndrome): a study of 42 fami-
sis of Crigler‐Najjar syndrome type 1 by genetic analysis of chorionic villus lies. N Engl J Med, 1967;277:1108.
samples. Hepatology, 1998;28:316A. 88. Arias, I.M. and London, I.M. Bilirubin glucuronide formation in vitro:
65. Gunn, C.H. Hereditary acholuric jaundice in a new mutant strain of rats. Demonstration of a defect in Gilbert’s disease. Science, 1957;126:563.
J Hered, 1938;29:137. 89. Davidson, A.R., Rojas‐Beuno, A., Thompson, R.P.H., and Williams, R.
66. Roy Chowdhury, N., Kondapalli, R., and Roy Chowdhury, J. The Gunn rat: Reduced caloric intake and nicotinic acid provocation tests in diagnosis of
An animal model for inherited deficiency of bilirubin glucuronidation, in Gilbert’s syndrome. Br Med J, 1975;2:480.
Animal Models in Liver Research (eds. C.E. Cornelius, R.R. Marshak, and 90. Bosma, P.J., Roy Chowdhury, J., Bakker, C. et al. A sequence abnormality in
E.C. Melby), Academic Press, New York, 1993, p. 150. the promoter region results in reduced expression of bilirubin‐UDP‐glucu-
67. Roy Chowdhury, N., Kondapalli, R., and Roy Chowdhury, J. The Gunn rat: ronosyltransferase‐1 in Gilbert syndrome. N Engl J Med, 1995;333:1171.
an animal model for inherited deficiency of bilirubin glucuronidation, in 91. Maruo, Y., Sato, H., Yamano, T. et al. Gilbert’s syndrome caused by homozy-
Animal Models in Liver Research (eds. C.E. Cornelius, R.R. Marshak, and gous missense mutation (Tyr486Asp) of bilirubin‐UDP glucuronosyl trans-
E.C. Melby), Academic Press, New York, 1993, pp. 150–75. ferase. J Pediatr, 1998;132:1045.
68. Polgar, Z., Li, Y., Li Wang, X. et al. Gunn rats as a surrogate model for evalu- 92. Soeda, Y., Yamamoto, K., Adachi, Y. et al. Predicted homozygous mis‐sense
ation of hepatocyte transplantation‐based therapies of Crigler‐Najjar syn- mutation in Gilbert’s syndrome. Lancet, 1995;346:1494.
drome type 1. Methods Mol Biol, 2017;1506:131–47. 93. Koiwai, O., Nishizawa, M., Hasada, K. et al. Gilbert’s syndrome is caused by
69. Chen, Y., Li, Y., Wang, X. et al. Amelioration of hyperbilirubinemia in Gunn a heterozygous missense mutation in the gene for bilirubin UDP‐glucurono-
rats after transplantation of human induced pluripotent stem cell‐derived syltransferase. Hum Mol Genet, 1995;4:1183.
hepatocytes. Stem Cell Reports, 2015;5:22–30. 94. Kadakol, A., Sappal, B.S., Ghosh, S.S. et  al. Interaction of coding region
70. Nguyen, N., Bonzo, J.A., Chen, S. et al. Disruption of the Ugt1 locus in mice mutations and the Gilbert‐type promoter abnormality of the UGT1A1 gene
resembles human Crigler‐Najjar type I disease. J Biol Chem, 2008; causes moderate degrees of unconjugated hyperbilirubinemia and may lead
283:7901–11. to neonatal kernicterus. J Med Genet, 2001;38:244.
71. Mazariegos, G., Shneider, B., Burton, B. et al. Liver transplantation for pedi- 95. Labrune, P., Myara, A., Chalas, J. et al. Association of a homozygous (TA)8
atric metabolic disease. Mol Genet Metab, 2014;111:418–27. promoter polymorphism and a N400D mutation of UGT1A1 in a child with
72. Kappas, A., Drummond, G.S., Manola, T., Petmezaki, S., and Valaes T Sn‐ Crigler‐Najjar type II syndrome. Hum Mutat, 2002;20:399.
protoporphyrin use in the management of hyperbilirubinemia in term new- 96. Wells, P.G., Mackenzie, P.I., Roy‐Chowdhury, J. et al. Glucuronidation and
borns with direct Coombs‐positive ABO incompatibility. Pediatrics, the UDP‐glucuronosyltransferases in health and disease. Drug Metab
1988;81(4):485–97. Dispos, 2004;32:281.
20:  Disorders of Bilirubin Metabolism 243

97. Iyer, L., King, C.D., Whitington, P.F. et al. Genetic predisposition to the 121. Kaplowitz, N., Javitt, N., and Kappas, A. Coproporphyrin I and III excre-
metabolism of irinotecan (CPT‐11): Role of uridine diphosphate glucu- tion in bile and urine. J Clin Invest, 1972;51:2895.
ronosyltransferase isoform 1A1 in the glucuronidation of its active metabo- 122. Seligsohn, U., Shani, M., Ramot, B. et al. Dubin‐Johnson syndrome in
lite (SN‐38) in human liver microsomes. J Clin Invest, 1998;101:847. Israel. II. Association with factor‐VII deficiency. Q J Med, 1970;39:569.
98. de Morais, S.M., Uetrecht, J.P., and Wells, P.G. Decreased glucuronidation 123. Kamisako, T., Leier, I., Cui, Y. et al. Transport of monoglucuronosyl and
and increased bioactivation of acetaminophen in Gilbert’s syndrome. bisglucuronosyl bilirubin by recombinant human and rat multidrug resist-
Gastroenterology, 1992;102:577. ance protein 2. Hepatology, 1999;30:485.
99. Ullrich, D., Sieg, A., Blume, R. et al. Normal pathways for glucuronidation, 124. Kartenbeck, J., Leuschner, U., Mayer, R., and Keppler, D. Absence of the
sulphation and oxidation of paracetamol in Gilbert’s syndrome. Eur J Clin canalicular isoform of the MRP gene‐encoded conjugate export pump
Invest, 1987;17:237. from the hepatocytes in Dubin‐Johnson syndrome. Hepatology, 1996;
100. Abumiya, M., Takahashi, N., Niioka, T. et al. Influence of UGT1A1 6, 27, 23:1061.
and 28 polymorphisms on nilotinib‐induced hyperbilirubinemia in Japanese 125. Allikmets, R., Gerrard, B., Hutchinson, A., and Dean, M. Characterization
patients with chronic myeloid leukemia. Drug Metab Pharmacokinet, of the human ABC superfamily: isolation and mapping of 21 new genes
2014;29:449–54. using the Expressed Sequence Tags database. Hum Mol Gen, 1996;5:1649.
101. Portman, O.W., Alexander, M., Roy Chowdhury, J. et al. Effects of nutri- 126. Keitel, V., Karenbeck, J., Nies, A.T. et al. Impaired protein maturation of the
tion on hyperbilirubinemia in Bolivian squirrel monkeys. Hepatology, conjugate export pump multidrug resistance protein 2 as a consequence of a
1984;4:454. deletion mutation in Dubin‐Johnson syndrome. Hepatology, 2000;32:1317.
102. Portman, O.W., Roy Chowdhury, J., Roy Chowdhury, N. et  al. A non‐ 127. Cornelius, C.E., Arias, I.M., and Osburn, B.I. Hepatic pigmentation with
human primate model for Gilbert’s syndrome. Hepatology, 1984;4:175. photosensitivity: a syndrome in Corriedale sheep resembling Dubin‐
103. Dubin, I.N. and Johnson, F.B. (1954) Chronic idiopathic jaundice with uni- Johnson syndrome in man. J Am Vet Med Assoc, 1965;146:709.
dentified pigment in liver cells: a new clinicopathologic entity with a report 128. Barnhart, J.L., Gronwall, R.R., and Combes, B. Biliary excretion of
of 12 cases. Medicine (Baltimore) 1954;33:155. ­sulfobromophthalein compounds in normal and mutant Corriedale sheep:
104. Sprinz, H. and Nelson, R.S. Persistent nonhemolytic hyperbilirubinemia evidence for a disproportionate transport defect for conjugated sulfo-
associated with lipochrome‐like pigment in liver cells: report of four cases. bromophthalein. Hepatology, 1981;1:441.
Ann Intern Med, 1954;41:952. 129. Bohan, A. and Boyer, J.L. Mechanisms of hepatic transport of drugs:
105. Dubin, I.N. Chronic idiopathic jaundice: a review of fifty cases. Am J Med, Implications for cholestatic drug reactions. Semin Liver Dis, 2002;22:123.
1958;23:268. 130. Jansen, P.L.M., van Klinken, J.W., van Gelder, M. et al. Preserved organic
106. Shani, M., Seligshon, U., Gilon, E. et  al. Dubin‐Johnson syndrome in anion transport in mutant TR‐ rats with a hepatobiliary secretion defect.
Israel: clinical, laboratory, and genetic aspects of 101 cases. West J Med, Am J Physiol, 1993;265:G445.
1970;39:549. 131. Jansen, P.L.M., Peters, W.H.M., and Lamers, W.H. Hereditary chronic
107. Come, S.E., Shohet, S.B., and Robinson, S.H. Surface remodeling vs. conjugated hyperbilirubinemia in mutant rats caused by defective hepatic
whole‐cell hemolysis of reticulocytes produced with erythroid stimulation anion transport. Hepatology, 1985;5:573.
or iron deficiency anemia. Blood, 1974;44:817. 132. Kitamura, T., Alroy, J., Gatmaitan, Z. et al. Defective biliary excretion of
108. Morita, M. and Kihava, T. Intravenous cholecystography and metabolism epinephrine metabolites in mutant (TR‐) rats: Relation to the pathogene-
of meglumine iodipamide (biligrafin) in Dubin‐Johnson syndrome. sis of black liver in the Dubin‐Johnson syndrome and Correidale sheep
Radiology, 1971;95:57. with an analogous excretory defect. Hepatology, 1992;15:1154.
109. Arias, I.M., Bernstein, L., Roffler, R., and Ben Ezzer, J. Black liver diseases 133. Takikawa, H., Nishikawa, K., Sano, N. et  al. Mechanisms of biliary
in Corriedale sheep: metabolism of tritiated epinephrine and incorporation excretion of lithocholate‐3‐sulfate in Eisai hyperbilirubinemic rats
of isotope into the hepatic pigment in vivo. J Clin Invest, 1965;44:1026. (EHBR). Dig Dis Sci, 1995;40:1792.
110. Swartz, H.M., Sarna, T., and Varma, R.R. On the nature and excretion of the 134. Schulman, F.Y., Montali, R.J., Bush, M. et al. Dubin‐Johnson‐like syn-
hepatic pigment in the Dubin‐Johnson syndrome. Gastroenterology, 1979;76:958. drome in Golden Lion tamarins (Leontopithecus rosalia rosalia). Vet
111. Ishikawa, T., Muller, M., Klunemann, C. et  al. ATP‐dependent primary Pathol, 1993;30:491.
active transport of cysteinyl leukotrienes across liver canalicular mem- 135. Rotor, A.B., Manahan, L., and Florentin, A. Familial nonhemolytic jaun-
branes. J Biol Chem, 1990;265:19279. dice with direct van den Bergh reaction. Acta Med Philos, 1948;5:37.
112. Jedlitschky, G., Leier, I., Buchholz, U. et al. ATP‐dependent transport of 136. Wolpert, E., Pascasio, F.M., Wolkoff, A.W., and Arias, I.M. Abnormal
glutathione S‐conjugates by the multidrug resistance‐associated protein. sulfobromophthalein metabolism in Rotor’s syndrome and obligate het-
Cancer Res, 1994;54:4833. erozygotes. N Engl J Med, 1977;296:1099.
113. Kobayashi, K., Sogame, Y., Hara, H., and Hayashi, K. Mechanism of glu- 137. Schiff, L., Billing, B.H., and Oikawa Y (1959) Familial nonhemolytic
tathione S‐conjugate transport in canalicular and basolateral rat liver jaundice with conjugated bilirubin in the serum; a case study. N Engl J
plasma membranes. J Biol Chem, 1990;265:7737. Med, 1959;260(26):1315–18.
114. Nishida, T., Hardenbrook, C., Gatmaitan, Z., and Arias, I.M. ATP‐depend- 137A. Porush, J.G., Delman, A.J., and Feuer, M.M. Chronic idiopathic jaundice
ent organic anion transport system in normal and TR− rat liver canalicular with normal liver histology. Arch Intern Med, 1962;109:102.
membranes. Am J Physiol, 1992;262:G629. 138. Wolkoff, A.W., Wolpert, E., Pascasio, F.N., and Arias, I.M. Rotor’s syn-
115. Paulusma, C.C., Bosma, P.J., Zaman, G.J.R. et al. Congenital jaundice in drome: A distinct inheritable pathophysiologic entity. Am J Med, 1976;60:173
rats with a mutation in a multidrug resistance‐associated protein gene. 139. Rapacini, G.L., Topi, G.C., Anti, M. et al. Porphyrins in Rotor syndrome:
Science, 1996;271:1126. A study on an Italian family. Hepatogastroenterology, 1986;33:11.
116. Takikawa, H., Sano, N., Narita, T. et al. Biliary excretion of bile acid con- 140. van de Steeg, E., Stranecky’, V., Hartmannova, H. et al. Complete OATP1B1
jugates in a hyperbilirubinemic mutant sprague‐dawley rat. Hepatology, and OATP1B3 deficiency causes human Rotor syndrome by interrupting
1991;14:352–60. conjugated bilirubin reuptake into the liver. J Clin Invest, 2012;122:519–28.
117. Togawa, T., Mizuochi, T., Sugiura, T. et al. Clinical, pathologic, and genetic 141. Maeda, K. Organic anion transporting polypeptide (OATP)1B1 and
features of neonatal Dubin‐Johnson syndrome: a multicenter study in OATP1B3 as important regulators of the pharmacokinetics of substrate
Japan. J Pediatr, 2018;196:161–7. drugs. Biol Pharm Bull, 2015;38:155–68.
118. Konig, J., Rost, D., Cui, Y., and Keppler, D. Characterization of the human 142. Link, E., Parish, S., Armitage, J. et al. SEARCH Collaborative Group.
multidrug resistance protein isoform MRP3 localized to the basolateral SLC01B1 variants and statin‐induced myopathy: a genomewide study. N
hepatocyte membrane. Hepatology, 1999;29:1156. Engl J Med, 2008;359:789–99.
119. Charbonnier, A. and Brisbois, P. Etude chromatographique de la BSP au 143. Cornelius, C.E. and Gronwall, R.R. Congenital photosensitivity and
cours de l’epreuve clinique d’epuration plasmatique de ce colorant. Rev hyperbilirubinemia in Southdown sheep in the United States. Am J Vet
Intern Hepatol, 1960;10:1163. Res, 1968;29:291–5.
120. Rodes, J., Zubizarreta, A., and Bruguera, M. Metabolism of bromsulphoph- 144. Posbergh, C.J.,.Kalla, S.E., Sutter, N.B., Tennant, B.C., and Huson, H.J.
thalein in Dubin‐Johnson syndrome: Diagnostic value of the paradoxical Mutation responsible for congenital photosensitivity and hyperbilirubine-
increase in plasma levels of BSP. Am J Dig Dis, 1972;17:545. mia in Southdown sheep. Am J Vet Res, 2018;79:538–45.
244 THE LIVER:  REFERENCES

145. Jansen, P.L. and Muller, M.M. Progressive familial intrahepatic cholestasis Proceedings of Falk Symposium 129 (eds. D. Keppler, U. Leuschner, G.
types 1, 2, and 3. Gut, 1998;42:766. Paumgartner, and A. Stiehl), Kluwer, Dordrecht, 2003, p. 184.
146. Sticova, E., Jirsa, M., and Pawlowska, J. New insights in genetic cholestasis: 161. Rosmorduc, O., Hermelin, B., and Poupon, R. MDR3 gene defect in adults
from molecular mechanisms to clinical implications. Can J Gastroenterol with symptomatic intrahepatic and gallbladder cholesterol cholelithiasis.
Hepatol, 2018, Article ID 2313675. https://doi.org/10.1155/2018/2313675 Gastroenterology, 2001;120:1459.
147. Clayton, R.J., Iber, F.L., Ruebner, B.H. et al. Byler disease: Fatal familial 162. Dixon, P.H., Weerasekera, N., Linton, K.J. et al. Heterozygous MDR3 mis-
intrahepatic cholestasis in an Amish kindred. Am J Dis Child, 1969;117:112. sense mutation associated with intrahepatic cholestasis of pregnancy: evi-
148. Bull, L.N., Van Eijk, M.J., Pawlikowaska, L. et al. A gene encoding a P‐ dence for a defect in protein trafficking. Hum Mol Genet, 2000;9:1209.
type ATPase mutated in two forms of hereditary cholestasis. Nat Genet, 163. Jacquemin, E., Cresteil, D., Manouvrier, S. et al. Heterozygous non‐sense
1998;18:219. mutation associated with intrahepatic cholestasis of pregnancy. Lancet,
149. Klomp, L.W., Vargas, J.C., van Mil, S.W. et al. Characterization of muta- 1999;353:210.
tions in ATP8B1 associated with hereditary cholestasis. Hepatology, 164. de Vree, J., Jacquemin, E., Strum, E. et al. Mutations in the MDR3 gene
2004;40:27. cause progressive familial intrahepatic cholestasis. Proc Acad Natl Sci U S
150. Chen, F., Ananthanarayanan, M., Emre, S. et al. Progressive familial intra- A, 1998;25:282.
hepatic cholestasis, type 1, is associated with decreased farnesoid X receptor 165. de Vree, J.M.L., Ottenhoff Smith, A.J., Aten, J. et al. Rapid correction of
activity. Gastroenterology, 2004;126:756. Mdr2 deficiency by transplantation of Mdr3 transgenic hepatocytes.
151. Summerskill, W.H.J. and Walshe, J.M. Benign recurrent intrahepatic Hepatology, 1998;28:387A.
obstructive jaundice. Lancet 1959;2:686. 166. Sambrotta, M., Strautnieks, S., Papouli, E. et al. Mutations in TJP2 cause
152. Tygstrup, N. and Jensen, B. Intermittent intrahepatic cholestasis of progressive cholestatic liver disease. Nat Genet, 2014;46:326–8.
unknown etiology in five young males from the Faroe Islands. Acta Med 167. Zhou, S., Hertel, P.M., Finegold, M.J. et al. Hepatocellular carcinoma associ-
Scand, 1969;185:523. ated with tight‐junction protein 2 deficiency. Hepatology, 2015;62:1914–16.
153. Summerfield, J.A., Scott, J., Berman, M. et al. Benign recurrent intrahepatic 168. Drouin, E., Russo, P., Tuchweber, B. et al. North American Indian cirrhosis
cholestasis: Studies of bilirubin kinetics, bile acids, and cholangiography. in children: a review of 30 cases. J Pediatr Gastroenterol Nutr, 2000;31:395.
Gut, 1980;21:154. 169. Rapola, J., Heikkila, P., and Fellman, V. (2002) Pathology of lethal fetal
154. Summerskill, W.H.J. The syndrome of benign recurrent cholestasis. Am J growth retardation syndrome with aminoaciduria, iron overload and lactric
Med, 1965;38:298. acidosis (GRACILE). Pediatr Pathol Mol Med, 2002;21:183.
155. Biempica, L., Gutstein, S., and Arias, I.M. Morphological and biochemi- 170. Emerick, E.M., Rand, E.B., Goldmuntz, E. et al. Features of Alagille syn-
cal studies of benign recurrent cholestasis. Gastroenterology, 1967; drome in 92 patients: frequency and relation to prognosis. Hepatology,
52:521. 1999;29:822.
156. Thompson, R. and Strautnieks, S. BSEP: Function and role in progressive 171. Piccoli, D.A. and Spinner, N.A. Alagille syndrome and the Jagged1 gene.
familial intrahepatic cholestasis. Semin Liver Dis, 2001;21:545. Semin Liver Dis, 2001;21:525.
157. Strautnieks, S.S., Kagalwalla, A.F., Tanner, M.S. et al. Identification of a 172. Oda, T., Elkaholoun, A.G., Meltzer, P.S., and Chandrashekharappa, S.C.
locus for progressive familial intrahepatic cholestasis (PFIC2) on chromo- Identification and cloning of the human homolog (Jag1) of the rat jagged1
some 2q24. Am J Hum Genet, 1997;61:630. gene from the Alagille syndrome critical region at 20p12. Genomics,
158. Stindt, J., Kluge, S., Dröge, C. et al. Bile salt export pump‐reactive antibod- 1997;43:376.
ies form a polyclonal, multi‐inhibitory response in antibody‐induced bile 173. Friederich, E., Vancompernolle, K., Louvard, D., and Vandekerckhove, J.
salt export pump deficiency. Hepatology, 2016;63:524–37. Villin function in the organization of the actin cytoskeleton. Correlation of
159. Nies, A.T., Gatmaitan, Z., and Arias, I.M. ATP‐dependent phosphatidylcho- in vivo effects to its biochemical activities in vitro. J Biol Chem,
line translocation in rat liver canalicular plasma membrane vesicles. J Lipid 1999;274:26751–60.
Res, 1996;37:1125. 174. Phillips, M.J., Azuma, T., Meredith, S.L. et al. Abnormalities in villin gene
160. Strautnieks, S.S., Knisley, A.S., Gerred, S. et al. Mutations in MDR3 in adult expression and canalicular microvillus structure in progressive cholestatic
onset cholangiopathy, in Bile Acids: From Genomics to Disease and Therapy. liver disease if childhood. Lancet, 2003;362:1090.
Hepatic Lipid Droplets
21 in Liver Function
and Disease
Douglas G. Mashek1,2, Wenqi Cui1, Linshan Shang1, and Charles P. Najt1
1
Department of Biochemistry, Molecular Biology and Biophysics, University of Minnesota,
Minneapolis, MN, USA
2
Department of Medicine, Division of Diabetes, Endocrinology and Metabolism, University of Minnesota,
Minneapolis, MN, USA

INTRODUCTION with an increase in the contribution of DNL and remnant uptake,


both of which are highly influenced by diet composition.
Lipid droplet (LD) accumulation is the defining characteristic of Interestingly, the contribution of FAs from DNL is markedly
non‐alcoholic and alcoholic fatty liver disease. Historically con­ increased in subjects with non‐alcoholic fatty liver disease
sidered to be inert and simply a sign of disease, LDs are increas­ (NAFLD) and can account for up to nearly 30% of hepatic FAs
ingly recognized as etiological factors in numerous liver diseases [2]. The increased flux through the DNL pathway likely contrib­
as well as having important non‐pathological roles in cell signal­ utes to NAFLD development and its comorbidities. Of special
ing and function. These dynamic properties of LDs are highly note is the role of dietary sucrose or its monosaccharide con­
regulated by the hundreds of proteins that coat the LD surface stituent fructose in DNL flux. Numerous studies have linked
and control lipid trafficking and flux. In this chapter, we will sucrose and/or fructose to NAFLD etiology, which results from
highlight the major pathways of lipid metabolism that influence high rate of flux of fructose carbons to the liver and its unique
LD accumulation, explore key proteins that regulate LD turnover metabolism relative to glucose (see [3] for review).
and that link LDs to cellular dysfunction and liver disease. Intracellular FAs must first be activated to their respective
acyl‐CoAs prior to entrance in downstream metabolic pathways.
This occurs in an ATP‐dependent reaction catalyzed by a family
of long‐chain acyl‐CoA synthetases (ACSL). In addition, the
LIPID UPTAKE AND SYNTHESIS family of fatty acid transport proteins (FATP) also possess acyl‐
CoA synthetase activity and contribute to FA uptake. Importantly,
Fatty acids (FAs) serve as the building blocks for the biosynthesis both the ACSL and FATP families consist of multiple isoforms
of many complex lipids including triacylglycerols (TAGs), with different intracellular localizations and substrate specificity
phospholipids, and cholesterol esters. The liver derives these [4]. As examples, several isoforms including ACSL3 and 5 and
FAs from three main sources  –  direct uptake from the blood, FATP2 and 5 promote TAG synthesis in the liver [5–8], highlighting
uptake and degradation of lipoprotein remnants, and de novo important roles in trafficking FAs to anabolic pathways. ACSL1
lipogenesis (DNL). Under fasting conditions, nearly all of the activates and channels FAs to the mitochondria for oxidation in
FAs entering the liver are derived from direct uptake as a result heart and adipose tissue [9, 10], but has minimal effects on FA
of enhanced adipose tissue lipolysis [1]. In the fed state, there is oxidation in the liver [11]. The isoform responsible for this initial
a decreased reliance from adipose‐derived FAs that coincides step in hepatic FA β‐oxidation has yet to be elucidated.

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
246 THE LIVER:  LIPOLYSIS

TRIACYLGLYCEROL BIOSYNTHESIS LIPOLYSIS


AND LIPID DROPLET BIOGENESIS
Catabolism of TAG stored within LDs is a major determinant of
The primary route for disposal of FAs is their esterification to LD size and number. Numerous proteins directly or indirectly
TAG and its subsequent storage in LDs. The coordination of influence this process and, as a result, impact the development
numerous acyltransferase enzymes and the phosphatase enzyme or resolution of hepatic steatosis. Although named after the tis­
lipin are required for the synthesis of TAG [12]. Each of the sue where it has its highest expression, adipose triglyceride
enzymes in this pathway has numerous isoforms with different lipase (ATGL) is also involved in hepatic TAG turnover. First
substrate specificity, which influence TAG composition. discovered in 2004, ATGL is now widely recognized as the pri­
Examples are the initial (glycerol‐3‐phosphate acyltransferase, mary cytosolic TAG hydrolase in numerous tissues [27].
GPAT) and terminal (diacylglycerol acyltransferase, DGAT) Ablation of ATGL in the liver results in TAG accumulation and
enzymes in the classical Kennedy pathway of TAG synthesis. reduced FA oxidation, while ATGL overexpression alleviates
GPAT1 shows substrate specificity for saturated FAs, which hepatic steatosis; ATGL does not influence very low density
accounts for the high percentage of saturated FAs in the sn‐1 lipoprotein (VLDL) secretion [28–30]. The increased FA oxida­
position of TAG [13]. Additionally, DGAT1 channels exoge­ tion is driven via enhanced PGC‐1α/PPAR‐α signaling. We have
nous FAs into TAG, whereas DGAT2 is more selective to FAs shown that ATGL drives this transcriptional network through
derived from DNL [14–16]. Thus, the relative activity levels of the protein deacetylase sirtuin 1 (SIRT1), which is required for
these various isoforms dictate the acyl composition of TAG. It ATGL‐mediated induction of oxidative metabolism [31]. Thus,
should also be noted that more recent work has identified mona­ ATGL appears to be an important signaling node that links LD
cylglycerol acyltransferases (MGAT), historically thought to catabolism to cell signaling as a means to coordinate down­
exist solely in the intestine, to be present in the liver and have stream transcriptional networks that control metabolism. The
increased expression in models of hepatic steatosis [17]. next enzyme in the classical lipolysis pathway is hormone‐­
Although the limited studies on the role of these enzymes are sensitive lipase (HSL), however, the contribution of hepatocyte
not conclusive [18, 19], the contribution of the monacylglycerol HSL to liver TAG catabolism is not known; HSL has been
synthetic pathway to hepatic TAG synthesis will likely be a shown to catalyze cholesterol ester hydrolysis in hepatocytes
focus of future studies. [32]. Monoacylglycerol lipase (MAGL), which catalyzes the
LDs are thought to develop in the endoplasmic reticulum last step in hepatic TAG breakdown, has not been studied exclu­
(ER), where neutral lipids accumulate within the leaflets of the sively in the liver. However, studies in mice lacking MAGL or
membrane bilayer. As TAG begins to accumulate, it forms a utilizing MAGL inhibitors reveal that MAGL plays a key role in
neutral lipid core surrounded by a phospholipid monolayer hepatic injury via decreasing endocannabinoids and, thereby,
embedded with proteins. This core and associated phospholipid promoting inflammation [33].
monolayer eventually bud from the ER, thus forming a cytosolic Given the importance of ATGL in catalyzing the initial step
LD (see [20, 21] for excellent reviews on LD biogenesis). The in cytosolic TAG hydrolysis, perhaps it is not surprising that
nascent LDs formed at the ER are small in diameter and are ATGL activity is highly regulated. A host of proteins that
coated with DGAT1 [22]. For LDs to increase in size, they can directly interact with ATGL to influence its activity have been
either undergo fusion with other LDs or synthesize TAG at the identified. CGI‐58 is widely accepted as the primary coactiva­
surface of LDs. Cell death‐inducing DFF45‐like effector C tor of ATGL [34]. Although ATGL activity is an important
(CIDEC) is a key protein involved in the fusion of LDs [23, 24]. mechanism through which CGI‐58 elicits its effects, it clearly
In response to FAs, CIDEC translocates from the ER to LDs has functions beyond ATGL. For example, CGI‐58 ablation
where it initiates fusion and the formation of large LDs [25]. results in an age‐dependent and robust (8‐ to 52‐fold) increase
The LD monolayer membrane contains numerous enzymes in liver TAG, whereas ATGL ablation only increases hepatic
involved in TAG synthesis. For example, several ACSLs as well TAG by ~3‐fold [29, 35]. Moreover, CGI‐58 ablation regu­
as GPAT4 and DGAT2 are present on the LD surface to promote lates hepatic TAG hydrolysis independent of ATGL, suggest­
TAG synthesis and expansion of preexisting LDs [22]. ing, at least in the liver, a more complicated role of these
As the LD grows with TAG, the phospholipid monolayer proteins in LD turnover [36].
must also expand. As such, the synthesis of phosphatidylcholine The perilipin (PLIN) family were the first LD proteins to be
(PC), the most abundant LD phospholipid, is a critical compo­ characterized. They consist of five family members, with vary­
nent regulating LD growth and dynamics [26]. Deposition of ing structural homology, that act to antagonize ATGL‐catalyzed
TAG in the LD core increases the size of the droplet, thereby lipolysis among other functions [37]. PLIN2 is the isoform with
diluting the relative amount of PC in the phospholipid mon­ the highest expression in the liver and is induced in response to
olayer. As diacylglycerol (DAG) and phosphatidic acid become FA loading in cells and NAFLD in both humans and rodents
enriched at the LD surface, the rate‐limiting enzyme in PC syn­ [38]. Liver‐specific ablation of PLIN2 prevents hepatic steatosis
thesis, CTP:phosphor‐choline cytidylyltransferase α (CCTα), and inflammation [39, 40]. Similar to PLIN2, ablation of PLIN3
translocates from the nucleus and binds to the LD monolayer. in the liver also reduces steatosis, but does not influence inflam­
Once there, CCTα becomes activated to facilitate CDP‐choline matory markers [41]. PLIN5 is highly expressed in oxidative
synthesis. Interestingly, other enzymes critical for PC synthesis tissues including the liver and its expression is robustly upregu­
are not localized on LDs, suggesting a unique role of CCTα in lated in response to fasting [42]. Liver‐specific PLIN5 ablation
controlling LD PC metabolism [26]. reduces steatosis, whereas its overexpression promotes steatosis,
which is likely due to its anti‐lipolytic effects [43, 44].
21:  Hepatic Lipid Droplets in Liver Function and Disease 247

In addition to the perilipin proteins discussed above, a host of microautophagy are precisely regulated and considered as
other proteins have been shown to directly interact with and pro‐survival mechanisms in response to oxidative stress and
antagonize ATGL. G0/G1 switch gene 2 (G0S2) is a potent nutrient limitation, chaperone‐mediated autophagy serves a
inhibitor of ATGL that influences hepatic energy metabolism. housekeeping role in the clearance and recycling of misfolded
G0S2 overexpression in the liver promotes steatosis and reduces protein aggregates to maintain cellular integrity. Through the
FA oxidation, while ablation of hepatic G0S2 does the opposite process of macro‐ and micro‐lipophagy autophagosomes and
[45–47]. Other interacting proteins including pigment epithelial lysosomes, respectively, target LDs for degradation. Once inter­
derived factor [48], hypoxia‐inducible protein 2 [49], and nalized, the lipids are hydrolyzed by lysosomal acid lipase, the
CIDEC [50] inhibit ATGL activity directly while UBXD8 tar­ only known lysosomal lipase with activity towards neutral lipids
gets it for degradation [51]. such as TAG and cholesterol ester. The FAs produced from the
hydrolysis of TAG and cholesterol ester are then available for
β‐oxidation or other downstream pathways.
Very low density lipoprotein secretion Since its initial discovery, a rapidly growing body of litera­
Approximately 70% of hepatic FAs that are destined for VLDL– ture has characterized autophagy/lipophagy in the liver. To date,
TAG first transit through the cytosolic LD pool before lipopro­ most studies have evaluated lipophagy in the broader context of
tein assembly [52, 53]. Members of the carboxylesterase (Ces) global autophagy induction or inhibition. Lipophagy is broadly
family of enzymes contribute to turnover of cytosolic LDs and regulated similar to autophagy; nutrients, insulin, and mTOR
their trafficking to LDs destined for VLDL secretion. In particu­ antagonize autophagy/lipophagy, whereas depletion of nutrients
lar, ablation of hepatic Ces1d (also known as triglyceride hydro­ or increase in activity of low energy sensors (SIRT1, AMPK,
lase, TGH), which colocalizes to ER luminal LDs, reduces etc.) promotes autophagy/lipophagy. At the transcriptional
VLDL secretion and increases the size, but decreases the num­ level, autophagy is regulated by a host of transcription factors
ber, of cytosolic LDs [54, 55]. It appears that Ces1d reduces the coinciding with their known roles in nutrient/energy sensing
transfer of lipids to existing cytosolic LDs, which may be due to [60]. Consistent with nutrient regulation, many small molecules
transfer of ER lipids to VLDL for packaging. While other CES (caffeine, ginsenoside, and quercetin to name a few) reduce
proteins have been studied, their role specifically in liver and hepatic steatosis through alterations in lipophagy [62–64].
VLDL secretion has not been characterized. In addition to the Figure 21.1 provides an overview of the above major pathways
carboxylesterases, the LD protein CIDEB plays an essential role involving LDs and key protein mediators of each pathway.
in VLDL lipidation. Specifically, ablation of CIDEB reduces Lipolysis of TAG via cytosolic lipases and lipophagy are the
VLDL–TAG secretion via an apolipoprotein B (ApoB)‐ two pathways thought to contribute to hepatic LD degradation.
dependent mechanism [56]. In contrast, ancient ubiquitous protein We recently explored the interrelationship between the two
1, which partially colocalizes to LDs, antagonizes ApoB‐medi­ pathways. These studies identified ATGL to act via SIRT1 sign­
ated VLDL lipidation as a means to suppress VLDL production aling as an upstream driver of autophagy/lipophagy [65].
[57]. These studies also highlight the importance of ApoB as a Moreover, autophagy/lipophagy is required for ATGL to drive
nexus between LDs and VLDL lipidation. Indeed, ApoB colocal­ LD degradation. Consistent with these studies, ablation of
izes to LDs on extensions of the ER, which is likely critical in PLIN2 also promotes LD degradation in a lipophagy‐dependent
cytosolic transfer of lipids to the developing VLDL particle [58]. manner [66]. Thus, these studies point towards a dominant role
of lipophagy in hepatic LD degradation. As our understanding
of the lipophagic process and its significance to hepatic LD
turnover evolves, this pathway will be undoubtedly targeted
LIPOPHAGY through small molecules, as mentioned above, or transcript/pro­
tein targeting as a means to alleviate hepatic steatosis.
Autophagy is a highly conserved and well‐characterized mecha­
nism through which cellular organelles, protein aggregates, and
macronutrients are degraded during times of nutrient insuffi­
ciency. Although an extensive body of literature characterizing LIPID DROPLETS AND LIVER DISEASE
autophagy of numerous organelles exists, autophagic degrada­
tion of LDs, termed lipophagy, was relatively more recently Non‐alcoholic fatty liver disease
identified [59] (see [60] for an extensive review). In this pro­
cess, multiple autophagic arms contribute to LD degradation. Lipid storage in LDs is a hallmark of NAFLD. Alterations in the
Macrolipophagy is the classical process through which balance of LD anabolism and catabolism are integral to the
autophagosomes bud off part of the LD prior to fusing with development and resolution of NAFLD. Thus, dozens of studies
lysosomes to form autolysosomes, which can then degrade show that reducing substrates for TAG synthesis, inhibiting
the lipid cargo. Microlipophagy describes the direct interaction enzymes involved in TAG synthesis, or increasing TAG hydrol­
and transfer of lipids between LDs and lysosomes. Finally, ysis can prevent or alleviate hepatic steatosis. Indeed, ongoing
chaperone‐mediated autophagy is characterized by degradation clinical trials are testing these pathways, especially those
of select proteins, which also may indirectly influence lipophagy. involved in TAG synthesis.
For example, degradation of specific LD proteins such as PNPLA3 is an LD protein that belongs to the same family of
PLIN2 via chaperone‐mediated autophagy allows for the subse­ enzymes as ATGL (i.e. PNPLA2). A nucleotide substitution in
quent degradation of LDs [61]. Although macroautophagy and PNPLA3 (rs738409, I148M) is widely recognized as the single
248 THE LIVER:  LIPID DROPLETS AND LIVER DISEASE

Figure 21.1  Key lipid droplet (LD) protein mediators of specific lipid metabolic pathways influencing LDs. Proteins that antagonize specific
pathways are highlighted in red.

largest genetic predictor of NAFLD [67]. The prevalence of this [76–78]. From these studies, we can glean that LDs interact
mutation is ~15–25% in most populations but increases to ~50% with numerous cellular organelles to coordinate metabolism as
in Hispanics [67]. PNPLA3 is highly upregulated in response to evidenced by the abundant presence of numerous organelle‐spe­
high carbohydrate diets and the I148M single‐nucleotide poly­ cific proteins in the LD proteome. Moreover, these studies have
morphism (SNP) is predictive of steatosis, especially when expedited the discovery of novel proteins with pathological
­carriers consume a high carbohydrate/sugar diet [68]. Similarly, roles in liver disease. One such example is 17β‐hydroxysteroid
the PNPLA3 variant is robustly increased on LDs in mice fed dehydrogenase‐13 (17β‐HSD13), which was identified and
high sucrose diets [69]. The mechanism(s) through which characterized to be increased in a LD proteomics screen of liver
PNPLA3 conveys its phenotype is currently under intense biopsies from NAFLD patients relative to non‐steatotic controls
debate. Although ample studies show that the I148M variant [77]. Overexpression of 17β‐HSD13 in the liver of mice
attenuates LD catabolism [70], other studies suggest a lipogenic increases steatosis confirming a direct effect on etiology of the
role of I148M through its lysophosphatidic acid acyltransferase disease [77]. Although the underlying mechanism of action is
activity [71]. Regardless of the mechanism of action, the pres­ still under investigation, 17β‐HSD13 overexpression in hepato­
ence of the I148M variant also progresses NAFLD to liver dam­ cytes increases lipogenesis, suggesting the increased FA supply
age and injury in numerous models of liver disease [72] as and subsequent TAG synthesis may be a driving factor underly­
discussed in more detail later. ing its ability to promote steatosis [77]. Since 17β‐HSD13 is
In addition to PNPLA3, several other LD proteins are altered thought to be involved in estrogen and androgen metabolism, it
in NAFLD and appear to be involved in disease etiology. The has been proposed that it may act to alter local pools of hor­
first discovered and perhaps most studied perilipin isoform is mones, leading to downstream changes in lipogenesis [79].
PLIN1. In adipose tissue, PLIN1 binds CGI‐58 under basal con­ Genome‐wide association studies also identified a polymor­
ditions, but phosphorylation in response to lipolytic stimuli dis­ phism (rs641738 C>T) that contains the membrane‐bound
rupts this interaction allowing CGI‐58 to partner with ATGL to O‐acyltransferase domain‐containing 7 gene (MBOAT7, or
promote lipolysis [73]. Although PLIN1 is normally expressed LPIAT1) and transmembrane channel‐like 4 gene (TMC4) to be
at very low or undetectable levels in the liver, it is upregulated in associated with increased hepatic fat content mediated by
livers of humans with NAFLD [38, 74]. It remains unknown if changes in the phosphatidylinositol acyl‐chain remodeling [80,
this increase is merely a consequence of increased LDs or if 81]. MBOAT7 is localized on intracellular membranes such as
PLIN1 plays an etiological role in NAFLD. PLIN2 and PLIN3 ER, mitochondria, and LDs and functions as a lysophosphoino­
are also increased in human NAFLD [38]. Consistent with their sitol acyltransferase [80]. Another prominent NAFLD‐associated
anti‐lipolytic roles, liver‐specific PLIN2 or PLIN3 ablation pre­ polymorphism (rs58542926, E167K) was identified in trans­
vents hepatic steatosis [39–41]. membrane 6 superfamily member 2 (TM6SF2) [82]. Despite
The surface of the LD is coated with hundreds of proteins being localized in the ER and the ER–Golgi intermediate com­
whose presence and/or abundance are dynamic. Although LD partment, but not on LDs, TM6SF2 regulates cholesterol metab­
proteomics has been conducted in numerous cell types, several olism and incorporation of polyunsaturated FAs into hepatic
studies have characterized the hepatic LD proteome and how TAG [83, 84]. TM6SF2 deficiency causes LD accumulation
it changes in response to fasting/refeeding [75] or NAFLD due, at least in part, to reduced TAG secretion [84].
21:  Hepatic Lipid Droplets in Liver Function and Disease 249

Alcoholic fatty liver disease


Relative to NAFLD, our understanding of LDs in alcoholic fatty
liver disease (AFLD) is less developed. The I148M polymor­
phism of PNPLA3 is positively associated with AFLD in
genome‐wide association studies [85, 86] consistent with its
broad role in promoting liver disease. In addition, TM6SF2
(rs58542926) and MBOAT7 (rs641738) genes are also identi­
fied as risk loci for alcohol‐related cirrhosis [87], suggesting the
importance of lipid metabolism in the pathogenesis of AFLD.
The expression level of PLIN2 increases to coincide with expan­
sion of the LD pool following ethanol exposure [88], suggesting
reduced lipolysis may be one factor contributing to increased
LDs. Lipophagy and expression of RAG GTPase Rab7, which is
required for lipophagy, are greatly inhibited by ethanol, further Figure 21.2  Lipid droplet proteins that link progression of steatosis to
supporting reduced LD catabolism as a contributing factor to non‐alcoholic steatohepatitis (NASH), hepatitis C virus (HCV), hepato­
AFLD [83]. Moreover, ethanol exposure inhibits β‐adrenergic cellular carcinoma (HCC), and insulin/glucose homeostasis.
induced phosphorylation of HSL and the recruitment of ATGL
to the LD surface, leading to decreased lipolysis [89]. More hepatic inflammation, lipid peroxidation, and fibrosis [100].
recently, a novel role of ceramide synthase 6 (CerS6) has been Thus, this terminal step in the TAG synthetic pathway appears
uncovered. Previous studies have shown that increased cera­ to play an important role in coupling/uncoupling steatosis from
mide synthesis and accumulation are linked to AFLD [90]. The the liver dysfunction that precedes NASH.
expression of CerS6, which is involved in ceramide synthesis, is Serine/threonine protein kinase 25 (STK25) is one of the few
increased in response to alcohol [91], and inhibition of ceramide kinases that have been characterized to reside on LDs in hepato­
synthesis attenuates alcohol‐mediated steatosis and liver dys­ cytes [101]. Overexpression of STK25 promotes steatosis, liver
function [92, 93]. Moreover, CerS6 drives PLIN2 expression inflammation, and fibrosis in mice [84]. Moreover, STK25 pro­
and knockdown of PLIN2 attenuates ceramide production fol­ tein abundance correlates with steatosis in humans [101].
lowing ethanol exposure [91]. Finally, CerS6 localizes to LDs Although the proteins downstream of STK25 remain to be elu­
and interacts with ACSL5, suggesting that ceramides are pro­ cidated, these initial studies suggest that STK25 could play a
duced at the LD surface [91, 94]. Although excess efflux of major role in hepatic LD biology and disease progression
ethanol carbons to DNL has long been thought to be the driving through its phosphorylation of LD proteins.
force in AFLD, these more recent yet limited studies suggest a Another important aspect of LDs during NASH development
more complex etiology. is their lipid composition. Specifically, Ioannou et al. observed
the presence of cholesterol crystals in LDs subjects with NASH
but not simple steatosis [102]. Increasing dietary cholesterol
Non‐alcoholic steatohepatitis
intake in rodents also increases LD cholesterol crystals and pro­
The transition from simple steatosis to non‐alcoholic steato­ motes inflammation and fibrosis [103, 104]. These studies are
hepatitis (NASH) is a pivotal point in disease pathology. consistent with human studies that show increased dietary cho­
Inflammation and reactive oxygen species are key components lesterol to be an independent risk factor for NASH and cirrhosis
in advancing progression to NASH. Since LDs directly interact [105]. Activated Kupffer cells surround dead hepatocytes laden
with mitochondria and can generate inflammatory eicosanoids with cholesterol crystals, suggesting that cell death and signal­
at the LD surface [95], perhaps it is not surprising that numerous ing to immune cells may be directly involved in explaining how
LD proteins are implicated in NAFLD progression. The I148M cholesterol crystals promote inflammation [102]. Additionally,
mutation of PNPLA3, in addition to promoting steatosis, is also PC levels are reduced in subjects with NAFLD and NASH,
a major driver of progression from steatosis to NASH in both which suggests that alterations in the LD monolayer could also
adult and pediatric populations [96, 97]. Ablation of CGI‐58 influence disease progression [106]. Figure 21.2 highlights the
causes NASH in mice [35], while ATGL ablation has no effect key LD proteins involved in the development of NASH and
on liver fibrosis [29], suggesting lipolysis‐independent effects other complications downstream of steatosis.
of CGI‐58 on NAFLD progression. Liver‐specific knockout of
PLIN2 alleviates NASH pathologies in a methionine choline‐
deficient (MCD) model of NASH [40]. Although its role in
Insulin resistance and glucose dysregulation
NASH etiology is not known, PLIN1 is highly expressed in Ectopic LD accumulation is correlated with insulin resistance
adults and children with NASH [98]. In contrast to proteins in numerous tissues including the liver. The consequences of
involved in TAG breakdown, the DGAT enzymes appear to have hepatic insulin resistance include excess hepatic glucose pro­
opposing effects on NASH. Treatment of mice with antisense duction and VLDL secretion, which contribute to the hyper­
oligonucleotides (ASOs) targeting DGAT1 attenuates fibrosis, glycemia and hypertriglyceridemia, respectively, commonly
induced by the MCD diet, and reduces stellate cell activation present in prediabetes and type 2 diabetes. Although numerous
without influencing hepatic steatosis [99]. In contrast, DGAT2 mechanisms underlying the development of insulin resistance
ablation in the same model reduces steatosis, but increases exist, alterations in lipid metabolism are a common theme.
250 THE LIVER:  LIPID DROPLETS AND LIVER DISEASE

Accumulation of intermediates in lipid metabolic pathways are accumulation of TAG and other signaling lipids in the liver can
thought to be the key drivers of hepatic insulin resistance rather be dissociated from insulin resistance or glucose intolerance,
than TAG itself. Intermediates in the lipid biosynthesis pathway suggesting that there are likely many mechanisms that can link
including phosphatidic acid, DAG, and ceramides may directly NAFLD to insulin/glucose homeostasis.
antagonize insulin signaling. Reactive oxygen species may In contrast to its effects on liver diseases, the PNPLA3 I148M
result from the oversupply of FAs to mitochondria and/or mito­ polymorphism appears to uncouple NAFLD from its commonly
chondrial dysfunction, leading to insulin resistance. Finally, associated metabolic complications. I148M carriers do not have
production of inflammatory signaling molecules within hepato­ elevated serum TAG, which is common in subjects with NAFLD
cytes or hepatic immune cells may also interfere with normal [67]. Most studies have reported I148M carriers do not have
hepatocyte function and insulin signaling. increased risk for insulin resistance or type 2 diabetes despite
LD dynamics appear to play a major role in the above pro­ the presence of NAFLD [115–117]. The latter study concluded
cesses that promote insulin resistance. PLIN proteins have sig­ that individuals with the I148M SNP, compared to noncarrier
nificant but differing effects on hepatic insulin sensitivity. For steatotic controls, have increased saturated and monosaturated
example, ablation of PLIN2 or PLIN3 in the liver improves hepatic lipids and reduced polyunsaturated lipid species, which
insulin sensitivity, coinciding with reduced steatosis [39–41]. may explain why the lipids are less toxic [89, 90]. Similar to
An SNP within the PLIN2 gene impairs VLDL assembly and PNPLA3 I148M, the TM6SF2 E167K variant also does not
correlates with type 2 diabetes [107]. In contrast, overexpres­ increase type 2 diabetes risk [118]. It will be of interest for
sion of PLIN5 in the liver increases steatosis, but improves insu­ future studies to further clarify the role of the PNPLA3 and
lin sensitivity [43]. These effects are consistent with studies on TM6SF2 polymorphisms in NAFLD‐related complications and
PLIN5 in other tissues, including muscle and heart, where LD identify other dietary, genetic, or environmental modifiers to
accumulation is dissociated from insulin resistance [108, 109]. this relationship.
Although the exact role and function of the perilipin proteins
remains an active area of research, much of their effects are
thought to occur via direct interactions with other LD proteins
Hepatitis C virus
to influence TAG turnover. The majority of hepatitis C virus (HCV)‐infected patients
One mechanism through which perilipin proteins may work have steatosis, which is even more pronounced in those with
is their interaction with lipases, such as ATGL, to inhibit lipol­ genotype 3A [119]. The increase in LD accumulation is likely
ysis under basal conditions [37]. ATGL overexpression in the due to the fact that the HCV life cycle is tightly intertwined
liver improves hepatic insulin signaling and reduces DAGs with LD metabolism. Several HCV proteins bind directly to
and ceramides in response to high fat feeding without altering LDs and the blockade of de novo LD formation via DGAT1
glucose tolerance [30]. Knockdown of liver ATGL in high fat inhibition prevents the translocation of the proteins from the
fed mice does not affect insulin signaling, but reduces gluco­ ER to LDs and blocks viral replication [120, 121]. Similarly,
neogenesis and improves glucose tolerance without changing the LD proteins PLIN3 and Rab18 are also required for direct
DAG levels [110]. ATGL is now recognized to form a regula­ interaction with HCV proteins and viral replication [122,
tory loop with FoxO1 in which ATGL drives FoxO1 activity 123]. HCV itself can promote hepatic steatosis and it appears
and FoxO1 drives ATGL expression reciprocally [111]. This to do so through the direct inhibition of LD catabolism. HCV
regulatory circuitry promotes lipolysis and FA oxidation, infection reduces TAG hydrolysis, in part through indirect
which support hepatic gluconeogenesis [111]. Knockdown of inhibition of ATGL‐mediated TAG hydrolysis [124]. Similarly,
the ATGL coactivator CGI‐58 increases liver TAG, DAG, and HCV suppressed the expression of the putative triglyceride
ceramides but improves glucose tolerance and insulin resist­ lipase arylacetamide deacetylase (AADAC) to reduce
ance [112]. The inhibitor of ATGL, G0S2, also influences lipolysis in the early stages of infection [125]. Although LDs
insulin resistance. G0S2 overexpression improves glucose tol­ are required for viral replication, VLDL secretion is essential
erance in mice with increased hepatic glucose uptake [45], for packaging and export of newly formed virus. Along
whereas G0S2 ablation in liver reduces steatosis, but improves these lines, ablation of AADAC, which also suppresses VLDL
whole‐body insulin sensitivity and glucose tolerance despite secretion, prevents virus production, suggesting that lipolysis
increased hepatic gluconeogenesis [46]. Ablation of hypoxia‐ may be required for late‐stage packaging and export of
inducible factor 2, an ATGL inhibitor, also promotes TAG the virus [125]. Consistent with this logic, the ATGL coacti­
catabolism and FA oxidation while improving glucose toler­ vator CGI‐58 is also essential for viral replication [126].
ance [113]. Taken as a whole, inhibiting lipolysis causes These effects may be due to increased CGI‐58‐mediated
hepatocytes to preferentially take up and burn glucose, which VLDL production. Given its role in VLDL secretion, perhaps
may explain the generally improved glucose tolerance in mod­ it is not surprising that CIDEB is also required for HCV
els of reduced hepatic lipolysis. In contrast, promoting lipoly­ release [127]. Nonetheless, CIDEB also directly interacts
sis causes less steatosis as a result of enhanced FA oxidation, with the core protein of HCV and is essential for entry and
but also increases gluconeogenesis with variable response to replication of the virus [127], suggesting a prominent role for
systemic glucose homeostasis depending on the protein CIDEB in the entire HCV life cycle. Although these data
manipulated to regulate lipolysis. Similarly, DGATs also influ­ clearly highlight an essential role of LD proteins and LD
ence insulin sensitivity. DGAT2 overexpression increases metabolism in regulating the HCV life cycle, much remains to
TAG, DAG, and ceramides but maintains insulin sensitivity be elucidated regarding the molecular underpinnings of this
[114]. Thus, while complicated, these studies suggest that relationship.
21:  Hepatic Lipid Droplets in Liver Function and Disease 251

Hepatocellular carcinoma and CRISPR/Cas9 technologies hold great potential for the
liver‐specific targeting of proteins to reverse or attenuate
Although LD accumulation is a common observation in hepatic ­d isease pathology. Undoubtedly, the LD will likely be a
neoplastic samples, the specific role of LDs and LD proteins in key target to help alleviate the burden of NAFLD and its
development of hepatocellular carcinoma (HCC) has not been comorbidities.
extensively studied. Given its high penetrance, the I148M poly­
morphism in PNPLA3 also is likely a major driver of HCC since
it is positively associated with the development of HCC [128,
129]. These effects are expected based on the robust effects of
the I148M SNP on the development of NASH, a known risk
REFERENCES
factor for HCC [130]. In addition to PNPLA3 (rs738409), 1. Barrows, B.R. and Parks, E.J. Contributions of different fatty acid sources to
TM6SF2 (rs58542926) is also a risk factor for the development very low‐density lipoprotein‐triacylglycerol in the fasted and fed states.
of HCC in alcohol‐related cirrhosis [131]. Similar to its pres­ J Clin Endocrinol Metab, 2006;91(4):1446–52.
ence in human NAFLD, PLIN1 expression is also increased in 2. Donnelly, K.L., Margosian, M.R., Sheth, S.S., Lusis, A.J., and Parks, E.J.
Increased lipogenesis and fatty acid reesterification contribute to hepatic
neoplastic liver cells although it is not clear if this merely
triacylglycerol stores in hyperlipidemic Txnip‐/‐ mice. J Nutr, 2004;
reflects more LDs or if it has a pathological role [132]. Finally, 134(6):1475–80.
MAGL has been shown to promote HCC via increased inflam­ 3. Hannou, S.A., Haslam, D.E., McKeown, N.M., and Herman, M.A. Fructose
mation and has been used as a prognostic indicator of HCC metabolism and metabolic disease. J Clin Invest, 2018;128(2):545–55.
[133, 134]. Clearly, much remains to be understood regarding 4. Grevengoed, T.J., Klett, E.L., and Coleman, R.A. Acyl‐CoA metabolism and
partitioning. Annu Rev Nutr, 2014;34:1–30.
the role of LDs in HCC and related hepatic malignancies.
5. Bu, S.Y. and Mashek, D.G. Hepatic long‐chain acyl‐CoA synthetase 5 medi­
ates fatty acid channeling between anabolic and catabolic pathways. J Lipid
Res, 2010;51(11):3270–80.
Not all fatty liver (or LDs) is created equal 6. Bu, S.Y., Mashek, M.T., and Mashek, D.G. Suppression of long chain acyl‐
As our understanding of NAFLD progresses, there is a growing CoA synthetase 3 decreases hepatic de novo fatty acid synthesis through
decreased transcriptional activity. J Biol Chem, 2009;284(44):30474–83.
appreciation that NAFLD is a very heterogeneous disease. As 7. Falcon, A., Doege, H., Fluitt, A. et al. FATP2 is a hepatic fatty acid trans­
discussed earlier, numerous factors including diet, genetics, pre­ porter and peroxisomal very long‐chain acyl‐CoA synthetase. AJP
disposing diseases (obesity, diabetes, etc.) and HCV can lead to Endocrinol Metab, 2010;299:E384–93.
NAFLD through different etiological paths. As a result, mani­ 8. Doege, H., Baillie, R.A., Ortegon, A.M. et al. Targeted deletion of FATP5
festation of the disease itself, as well as its complications, can reveals multiple functions in liver metabolism: alterations in hepatic lipid
homeostasis. Gastroenterology, 2006;130:1245–58.
vary widely among individuals. These differences also likely 9. Ellis, J.M., Mentock, S.M., Depetrillo, M.A. et al. Mouse cardiac acyl coen­
contribute to the difficulties in identifying efficacious treatment zyme a synthetase 1 deficiency impairs fatty acid oxidation and induces car­
regiments for NAFLD. Clearly, an important focus of research diac hypertrophy. Mol Cell Biol, 2011;31(6):1252–62.
and diagnostic medicine moving forward should be towards 10. Ellis, J.M., Li, L.O., Wu, P.‐C. et al. Adipose acyl‐CoA synthetase‐1 directs
fatty acids toward beta‐oxidation and is required for cold thermogenesis.
characterizing these distinct forms of NAFLD and targeting
Cell Metab, 2010;12(1):53–64.
therapies specific for each case. 11. Li, L.O., Ellis, J.M., Paich, H.A. et al. Liver‐specific loss of long chain acyl‐
Analogous to NAFLD, there is also immense variability in CoA synthetase‐1 decreases triacylglycerol synthesis and beta‐oxidation and
LDs both across the liver and within cells. LD proteins are com­ alters phospholipid fatty acid composition. J Biol Chem, 2009;284(41):
monly observed to have uneven distribution across LDs within 27816–26.
12. Coleman, R.A. and Mashek, D.G. Mammalian triacylglycerol metabolism:
an individual cell. Similarly, the lipid composition of LDs varies
synthesis, lipolysis, and signaling. Chem Rev, 2011;111(10):6359–86.
greatly; Raman spectroscopy of NAFLD reveals substantial het­ 13. Vancura, A. and Haldar, D. Purification and characterization of glycerophos­
erogeneity of LD composition [135]. Moreover, the liver is phate acyltransferase from rat liver mitochondria. J Biol Chem, 1994;
highly zonated with periportal and perivenous hepatocytes hav­ 269(44):27209–15.
ing unique and very different gene signatures and metabolic 14. Wurie, H.R., Buckett, L., and Zammit, V.A. Diacylglycerol acyltransferase 2
acts upstream of diacylglycerol acyltransferase 1 and utilizes nascent diglyc­
functions [136]. Beyond differential expression of PLIN pro­ erides and de novo synthesized fatty acids in HepG2 cells. FEBS J,
teins [132], almost nothing is known regarding the differences 2012;279(17):3033–47.
in LDs among these different populations of cells. It would be 15. Qi, J., Lang, W., Geisler, J.G. et al. The use of stable isotope‐labeled glycerol
logical to assume that LD metabolism and signaling differ and oleic acid to differentiate the hepatic functions of DGAT1 and ‐2. J Lipid
greatly among and within hepatocytes, which is an area that Res, 2012;53(6):1106–16.
16. Villanueva, C.J., Monetti, M., Shih, M. et  al. Specific role for acyl
begs for future investigation. CoA:diacylglycerol acyltransferase 1 (Dgat1) in hepatic steatosis due to
exogenous fatty acids. Hepatology, 2009;50(2):434–42.
17. Hall, A.M., Kou, K., Chen, Z. et al. Evidence for regulated monoacylglyc­
Perspective erol acyltransferase expression and activity in human liver. J Lipid Res,
2012;53(5):990–9.
LDs are the distinguishing trait of NAFLD and play important 18. Soufi, N., Hall, A.M., Chen, Z. et al. Inhibiting monoacylglycerol acyltrans­
roles in both the development and progression of NAFLD, a ferase 1 ameliorates hepatic metabolic abnormalities but not inflammation
disease without efficacious and approved treatments. Under­ and injury in mice. J Biol Chem, 2014;289(43):30177–88.
standing how alterations in the LD proteome and lipidome 19. Hall, A.M., Soufi, N., Chambers, K.T. et al. Abrogating monoacylglycerol
acyltransferase activity in liver improves glucose tolerance and hepatic insu­
change to influence disease development and severity and define lin signaling in obese mice. Diabetes, 2014;63(7):2284–96.
unique subtypes of NAFLD will be a critical area of research 20. Wilfling, F., Haas, J.T., and Walther, T.C., and Farese, R.V. Jr. Lipid droplet
moving forward. New therapeutic strategies including ASOs biogenesis. Curr Opin Cell Biol, 2014;29:39–45.
252 THE LIVER:  REFERENCES

21. Pol, A., Gross, S.P., and Parton, R.G. Review: biogenesis of the multifunc­ 44. Wang, C., Zhao, Y., Gao, X. et al. Perilipin 5 improves hepatic lipotoxicity
tional lipid droplet: lipids, proteins, and sites. J Cell Biol, 2014;204(5): by inhibiting lipolysis. Hepatology, 2015;61(3):870–82.
635–46. 45. Wang, Y., Zhang, Y., Qian, H. et al. The g0/g1 switch gene 2 is an important
22. Wilfling, F., Wang, H., Haas, J.T. et al. Triacylglycerol synthesis enzymes regulator of hepatic triglyceride metabolism. PloS One, 2013;8(8):e72315.
mediate lipid droplet growth by relocalizing from the ER to lipid droplets. 46. Zhang, X., Xie, X., Heckmann, B.L. et al. Targeted disruption of G0/G1 switch
Dev Cell, 2013;24(4):384–99. gene 2 enhances adipose lipolysis, alters hepatic energy balance, and alleviates
23. Xu, W., Wu, L., Yu, M. et al. Differential Roles of Cell Death‐inducing DNA high‐fat diet‐induced liver steatosis. Diabetes, 2014;63(3):934–46.
fragmentation factor‐α‐like effector (CIDE) proteins in promoting lipid 47. Sugaya, Y. and Satoh, H. Liver‐specific G0 /G1 switch gene 2 (G0s2) expres­
droplet fusion and growth in subpopulations of hepatocytes. J Biol Chem, sion promotes hepatic insulin resistance by exacerbating hepatic steatosis in
2016;291(9):4282–93. male Wistar rats. J Diabetes, 2017;9(8):754–63.
24. Gao, G., Chen, F.‐J., Zhou, L. et  al. Control of lipid droplet fusion and 48. Chung, C., Doll, J.A., Gattu, A.K. et  al. Anti‐angiogenic pigment epithe­
growth by CIDE family proteins. Biochim Biophys Acta Mol Cell Biol lium‐derived factor regulates hepatocyte triglyceride content through adi­
Lipids, 2017;1862(10, Part B):1197–204. pose triglyceride lipase (ATGL). J Hepatol, 2008;48(3):471–8.
25. Li, H., Chen, A., Shu, L. et  al. Translocation of CIDEC in hepatocytes 49. Das, K.M.P., Wechselberger, L., Liziczai, M. et al. Hypoxia‐inducible lipid
depends on fatty acids. Genes Cells, 2014;19(11):793–802. droplet‐associated protein inhibits adipose triglyceride lipase. J Lipid Res,
26. Krahmer, N., Guo, Y., Wilfling, F. et al. Phosphatidylcholine synthesis for 2018;59(3):531–41.
lipid droplet expansion is mediated by localized activation of 50. Grahn, T.H.M., Kaur, R., Yin, J. et al. Fat‐specific protein 27 (FSP27) inter­
CTP:phosphocholine cytidylyltransferase. Cell Metab, 2011;14(4):504–15. acts with adipose triglyceride lipase (ATGL) to regulate lipolysis and insulin
27. Zimmermann, R., Strauss, J.G., Haemmerle, G. et  al. Fat mobilization in sensitivity in human adipocytes. J Biol Chem, 2014;289(17):12029–39.
adipose tissue is promoted by adipose triglyceride lipase. Science, 51. Olzmann, J.A., Richter, C.M., and Kopito, R.R. Spatial regulation of UBXD8
2004;306(5700):1383–6. and p97/VCP controls ATGL‐mediated lipid droplet turnover. Proc Natl
28. Ong, K.T., Mashek, M.T., Bu, S.Y., Greenberg, A.S., and Mashek, D.G. Acad Sci U S A, 2013;110(4):1345–50.
Adipose triglyceride lipase is a major hepatic lipase that regulates triacylg­ 52. Yang, L.Y., Kuksis, A., Myher, J.J., and Steiner, G. Origin of triacylglycerol
lycerol turnover and fatty acid signaling and partitioning. Hepatology, moiety of plasma very low density lipoproteins in the rat: structural studies.
2011;53(1):116–26. J Lipid Res, 1995;36(1):125–36.
29. Wu, J.W., Wang, S.P., Alvarez, F. et al. Deficiency of liver adipose triglycer­ 53. Yang, L.Y., Kuksis, A., Myher, J.J., and Steiner, G. Contribution of de novo
ide lipase in mice causes progressive hepatic steatosis. Hepatology, fatty acid synthesis to very low density lipoprotein triacylglycerols: evidence
2011;54(1):122–32. from mass isotopomer distribution analysis of fatty acids synthesized from
30. Turpin, S.M., Hoy, A.J., Brown, R.D. et al. Adipose triacylglycerol lipase is [2H6]ethanol. J Lipid Res, 1996;37(2):262–74.
a major regulator of hepatic lipid metabolism but not insulin sensitivity in 54. Wang, H., Wei, E., Quiroga, A.D. et  al. Altered lipid droplet dynamics in
mice. Diabetologia, 2011;54(1):146–56. hepatocytes lacking triacylglycerol hydrolase expression. Mol Biol Cell,
31. Khan, S.A., Sathyanarayan, A., Mashek, M.T. et al. ATGL‐catalyzed lipoly­ 2010;21(12):1991–2000.
sis regulates SIRT1 to control PGC‐1α/PPAR‐α signaling. Diabetes, 55. Lian, J., Wei, E., Wang, S.P. et  al. Liver specific inactivation of carboxy­
2015;64(2):418–26. lesterase 3/triacylglycerol hydrolase decreases blood lipids without causing
32. Sekiya, M., Osuga, J.‐I., Yahagi, N. et  al. Hormone‐sensitive lipase is severe steatosis in mice. Hepatology, 2012;56(6):2154–62.
involved in hepatic cholesteryl ester hydrolysis. J Lipid Res, 2008;49(8): 56. Ye, J., Li, J.Z., Liu, Y. et al. Cideb, an ER‐ and lipid droplet‐associated pro­
1829–38. tein, mediates VLDL lipidation and maturation by interacting with apolipo­
33. Cao, Z., Mulvihill, M.M., Mukhopadhyay, P. et al. Monoacylglycerol lipase protein B. Cell Metab, 2009;9(2):177–90.
controls endocannabinoid and eicosanoid signaling and hepatic injury in 57. Zhang, J., Zamani, M., Thiele, C. et al. AUP1 (ancient ubiquitous protein 1)
mice. Gastroenterology, 2013;144(4):808–17.e15. is a key determinant of hepatic very‐low‐density lipoprotein assembly and
34. Lass, A., Zimmermann, R., Haemmerle, G. et al. Adipose triglyceride lipase‐ secretion. Arterioscler Thromb Vasc Biol, 2017;37(4):633–42.
mediated lipolysis of cellular fat stores is activated by CGI‐58 and defective 58. Ohsaki, Y., Cheng, J., Suzuki, M., Fujita, A., and Fujimoto, T. Lipid droplets
in Chanarin‐Dorfman syndrome. Cell Metab, 2006;3(5):309–19. are arrested in the ER membrane by tight binding of lipidated apolipoprotein
35. Guo, F., Ma, Y., Kadegowda, A.K.G. et al. Deficiency of liver Comparative B‐100. J Cell Sci, 2008;121(14):2415–22.
Gene Identification‐58 causes steatohepatitis and fibrosis in mice. J Lipid 59. Singh, R., Kaushik, S., Wang, Y. et al. Autophagy regulates lipid metabolism.
Res, 2013;54(8):2109–20. Nature, 2009;458(7242):1131–5.
36. Lord, C.C., Ferguson, D., Thomas, G. et al. Regulation of hepatic triacylg­ 60. Schulze, R.J., Sathyanarayan, A., and Mashek, D.G. Breaking fat: the regula­
lycerol metabolism by CGI‐58 does not require ATGL co‐activation. Cell tion and mechanisms of lipophagy. Biochim Biophys Acta, 2017;1862(10 Pt
Rep, 2016;16(4):939–49. B):1178–87.
37. Kimmel, A.R. and Sztalryd, C. The perilipins: major cytosolic lipid droplet‐ 61. Kaushik, S. and Cuervo, A.M. AMPK‐dependent phosphorylation of lipid
associated proteins and their roles in cellular lipid storage, mobilization, and droplet protein PLIN2 triggers its degradation by CMA. Autophagy,
systemic homeostasis. Annu Rev Nutr, 2016;36:471–509. 2016;12(2):432–8.
38. Straub, B.K., Stoeffel, P., Heid, H., Zimbelmann, R., and Schirmacher, P. 62. Sinha, R.A., Farah, B.L., Singh, B.K. et al. Caffeine stimulates hepatic lipid
Differential pattern of lipid droplet‐associated proteins and de novo per­ metabolism by the autophagy‐lysosomal pathway in mice. Hepatology,
ilipin expression in hepatocyte steatogenesis. Hepatology, 2008;47(6): 2014;59(4):1366–80.
1936–46. 63. Huang, Q., Wang, T., Yang, L., and Wang, H.‐Y. Ginsenoside Rb2 alleviates
39. Imai, Y., Boyle, S., Varela, G.M. et al. Effects of perilipin 2 antisense oligo­ hepatic lipid accumulation by restoring autophagy via induction of Sirt1 and
nucleotide treatment on hepatic lipid metabolism and gene expression. activation of AMPK. Int J Mol Sci, 2017;18(5).
Physiol Genomics, 2012;44(22):1125–31. 64. Zhu, X., Xiong, T., Liu, P. et al. Quercetin ameliorates HFD‐induced NAFLD
40. Najt, C.P., Senthivinayagam, S., Aljazi, M.B. et  al. Liver‐specific loss of by promoting hepatic VLDL assembly and lipophagy via the IRE1a/XBP1s
perilipin 2 alleviates diet‐induced hepatic steatosis, inflammation, and fibro­ pathway. Food Chem Toxicol, 2018;114:52–60.
sis. Am J Physiol Gastrointest Liver Physiol, 2016;310(9):G726–38. 65. Sathyanarayan, A., Mashek, M.T., and Mashek, D.G. ATGL promotes
41. Carr, R.M., Patel, R.T., Rao, V. et al. Reduction of TIP47 improves hepatic autophagy/lipophagy via SIRT1 to control hepatic lipid droplet catabolism.
steatosis and glucose homeostasis in mice. Am J Physiol Regul Integr Comp Cell Rep, 2017;19(1):1–9.
Physiol, 2012;302(8):R996–1003. 66. Tsai, T.‐H., Chen, E., Li, L. et  al. The constitutive lipid droplet protein
42. Wolins, N.E., Quaynor, B.K., Skinner, J.R. et al. OXPAT/PAT‐1 is a PPAR‐ PLIN2 regulates autophagy in liver. Autophagy, 2017;13(7):1130–44.
induced lipid droplet protein that promotes fatty acid utilization. Diabetes, 67. Romeo, S., Kozlitina, J., Xing, C. et al. Genetic variation in PNPLA3 confers
2006;55(12):3418–28. susceptibility to nonalcoholic fatty liver disease. Nat Genet, 2008;40(12):
43. Trevino, M.B., Mazur‐Hart, D., Machida, Y. et al. Liver perilipin 5 expres­ 1461–5.
sion worsens hepatosteatosis but not insulin resistance in high fat‐fed mice. 68. Davis, J.N., Lê, K.‐A., Walker, R.W. et  al. Increased hepatic fat in over­
Mol Endocrinol, 2015;29(10):1414–25. weight Hispanic youth influenced by interaction between genetic variation in
21:  Hepatic Lipid Droplets in Liver Function and Disease 253

PNPLA3 and high dietary carbohydrate and sugar consumption. Am J Clin 92. Lizarazo, D., Zabala, V., Tong, M., Longato, L., and de la Monte, S.M.
Nutr, 2010;92(6):1522–7. Ceramide inhibitor myriocin restores insulin/insulin growth factor signal­
69. Smagris, E., BasuRay, S., Li, J. et al. Pnpla3I148M knockin mice accumu­ ing for liver remodeling in experimental alcohol‐related steatohepatitis.
late PNPLA3 on lipid droplets and develop hepatic steatosis. Hepatology, J Gastroenterol Hepatol, 2013;28(10):1660–8.
2015;61(1):108–11. 93. Tong, M., Longato, L., Ramirez, T. et al. Therapeutic reversal of chronic
70. He, S., McPhaul, C., Li, J.Z. et al. A sequence variation (I148M) in PNPLA3 alcohol‐related steatohepatitis with the ceramide inhibitor myriocin. Int J Exp
associated with nonalcoholic fatty liver disease disrupts triglyceride hydroly­ Pathol, 2014;95(1):49–63.
sis. J Biol Chem, 2010;285(9):6706–15. 94. Senkal, C.E., Salama, M.F., Snider, A.J. et al. Ceramide is metabolized to
71. Kumari, M., Schoiswohl, G., Chitraju, C. et al. Adiponutrin functions as a acylceramide and stored in lipid droplets. Cell Metab, 2017;25(3):686–97.
nutritionally regulated lysophosphatidic acid acyltransferase. Cell Metab, 95. Bozza, P.T., Bakker‐Abreu, I., Navarro‐Xavier, R.A., and Bandeira‐Melo,
2012;15(5):691–702. C. Lipid body function in eicosanoid synthesis: an update. Prostaglandins
72. Bruschi, F.V., Tardelli, M., Claudel, T., and Trauner, M. PNPLA3 expression Leukot Essent Fatty Acids, 2011;85(5):205–13.
and its impact on the liver: current perspectives. Hepatic Med Evid Res, 96. Singal, A.G., Manjunath, H., Yopp, A.C. et al. The effect of PNPLA3 on
2017;9:55–66. fibrosis progression and development of hepatocellular carcinoma: a meta‐
73. Subramanian, V., Rothenberg, A., Gomez, C. et al. Perilipin A mediates the analysis. Am J Gastroenterol, 2014;109(3):325–34.
reversible binding of CGI‐58 to lipid droplets in 3T3‐L1 adipocytes. J Biol 97. Valenti, L., Alisi, A., Galmozzi, E. et al. I148M patatin‐like phospholipase
Chem, 2004;279(40):42062–71. domain‐containing 3 gene variant and severity of pediatric nonalcoholic
74. Pawella, L.M., Hashani, M., Eiteneuer, E. et al. Perilipin discerns chronic fatty liver disease. Hepatology, 2010;52(4):1274–80.
from acute hepatocellular steatosis. J Hepatol, 2014;60(3):633–42. 98. Carr, R.M., Dhir, R., Mahadev, K. et  al. Perilipin staining distinguishes
75. Kramer, D.A., Quiroga, A.D., Lian, J., Fahlman, R.P., and Lehner, R. Fasting between steatosis and non‐alcoholic steatohepatitis in adults and children.
and refeeding induces changes in the mouse hepatic lipid droplet proteome. Clin Gastroenterol, 2017;15(1):145–7.
J Proteomics, 2018;181:213–24. 99. Yamaguchi, K., Yang, L., McCall, S. et al. Diacylglycerol acyltranferase 1
76. Khan, S.A., Wollaston‐Hayden, E.E., Markowski, T.W., Higgins, L., and anti‐sense oligonucleotides reduce hepatic fibrosis in mice with nonalco­
Mashek, D.G. Quantitative analysis of the murine lipid droplet‐associated holic steatohepatitis. Hepatology, 2008;47(2):625–35.
proteome during diet‐induced hepatic steatosis. J Lipid Res, 2015;56(12): 100. Yamaguchi, K., Yang, L., McCall, S. et al. Inhibiting triglyceride synthesis
2260–72. improves hepatic steatosis but exacerbates liver damage and fibrosis in
77. Su, W., Wang, Y., Jia, X. et al. Comparative proteomic study reveals 17β‐ obese mice with nonalcoholic steatohepatitis. Hepatology, 2007;45(6):
HSD13 as a pathogenic protein in nonalcoholic fatty liver disease. Proc Natl 1366–74.
Acad Sci U S A, 2014;111(31):11437–42. 101. Amrutkar, M., Cansby, E., Nuñez‐Durán, E. et al. Protein kinase STK25
78. Liu, M., Ge, R., Liu, W. et  al. Differential proteomics profiling identifies regulates hepatic lipid partitioning and progression of liver steatosis and
LDPs and biological functions in high‐fat diet‐induced fatty livers. J Lipid NASH. FASEB J, 2015;29(4):1564–76.
Res, 2017;58(4):681–94. 102. Ioannou, G.N., Haigh, W.G., Thorning, D., and Savard, C. Hepatic choles­
79. Zhang, X., Wang, Y., and Liu, P. Omic studies reveal the pathogenic lipid terol crystals and crown‐like structures distinguish NASH from simple
droplet proteins in non‐alcoholic fatty liver disease. Protein Cell, steatosis. J Lipid Res, 2013;54(5):1326–34.
2017;8(1):4–13. 103. Ioannou, G.N., Subramanian, S., Chait, A. et al. Cholesterol crystallization
80. Mancina, R.M., Dongiovanni, P., Petta, S. et al. The MBOAT7‐TMC4 vari­ within hepatocyte lipid droplets and its role in murine NASH. J Lipid Res,
ant rs641738 increases risk of nonalcoholic fatty liver disease in individuals 2017;58(6):1067–79.
of European descent. Gastroenterology, 2016;150(5):1219–30.e6. 104. Savard, C., Tartaglione, E.V., Kuver, R. et  al. Synergistic interaction of
81. Luukkonen, P.K., Zhou, Y., Hyötyläinen, T. et  al. The MBOAT7 variant dietary cholesterol and dietary fat in inducing experimental steatohepatitis.
rs641738 alters hepatic phosphatidylinositols and increases severity of non‐ Hepatology, 2013;57(1):81–92.
alcoholic fatty liver disease in humans. J Hepatol, 2016;65(6):1263–5. 105. Yasutake, K., Nakamuta, M., Shima, Y. et al. Nutritional investigation of
82. Kozlitina, J., Smagris, E., Stender, S. et al. Exome‐wide association study non‐obese patients with non‐alcoholic fatty liver disease: the significance
identifies a TM6SF2 variant that confers susceptibility to nonalcoholic fatty of dietary cholesterol. Scand J Gastroenterol, 2009;44(4):471–7.
liver disease. Nat Genet, 2014;46(4):352–6. 106. Puri, P., Baillie, R.A., Wiest, M.M. et al. A lipidomic analysis of nonalco­
83. Mahdessian, H., Taxiarchis, A., Popov, S. et al. TM6SF2 is a regulator of holic fatty liver disease. Hepatology, 2007;46(4):1081–90.
liver fat metabolism influencing triglyceride secretion and hepatic lipid 107. Sentinelli, F., Capoccia, D., Incani, M. et al. The perilipin 2 (PLIN2) gene
droplet content. Proc Natl Acad Sci U S A, 2014;111(24):8913–18. Ser251Pro missense mutation is associated with reduced insulin secretion
84. Luukkonen, P.K., Zhou, Y., Nidhina Haridas, P.A. et  al. Impaired hepatic and increased insulin sensitivity in Italian obese subjects. Diabetes Metab
lipid synthesis from polyunsaturated fatty acids in TM6SF2 E167K variant Res Rev, 2016;32(6):550–6.
carriers with NAFLD. J Hepatol, 2017;67(1):128–36. 108. Bosma, M., Sparks, L.M., Hooiveld, G.J. et al. Overexpression of PLIN5 in
85. Tian, C., Stokowski, R.P., Kershenobich, D., Ballinger, D.G., and Hinds, skeletal muscle promotes oxidative gene expression and intramyocellular
D.A. Variant in PNPLA3 is associated with alcoholic liver disease. Nat lipid content without compromising insulin sensitivity. Biochim Biophys
Genet, 2010;42(1):21–3. Acta, 2013;1831(4):844–52.
86. Chamorro, A.‐J., Torres, J.‐L., Mirón‐Canelo, J.‐A. et al. Systematic review 109. Kuramoto, K., Okamura, T., Yamaguchi, T. et al. Perilipin 5, a lipid droplet‐
with meta‐analysis: the I148M variant of patatin‐like phospholipase domain‐ binding protein, protects heart from oxidative burden by sequestering fatty
containing 3 gene (PNPLA3) is significantly associated with alcoholic liver acid from excessive oxidation. J Biol Chem, 2012;287(28):23852–63.
cirrhosis. Aliment Pharmacol Ther, 2014;40(6):571–81. 110. Ong, K.T., Mashek, M.T., Bu, S.Y., and Mashek, D.G. Hepatic ATGL
87. Buch, S., Stickel, F., Trépo, E. et al. A genome‐wide association study con­ knockdown uncouples glucose intolerance from liver TAG accumulation.
firms PNPLA3 and identifies TM6SF2 and MBOAT7 as risk loci for alcohol‐ FASEB J, 2013;27(1):313–21.
related cirrhosis. Nat Genet, 2015;47(12):1443–8. 111. Zhang, W., Bu, S.Y., Mashek, M.T. et al. Integrated regulation of hepatic
88. Mak, K.M., Ren, C., Ponomarenko, A., Cao, Q., and Lieber, C.S. Adipose lipid and glucose metabolism by adipose triacylglycerol lipase and FoxO
differentiation‐related protein is a reliable lipid droplet marker in alcoholic proteins. Cell Rep, 2016;15(2):349–59.
fatty liver of rats. Alcohol Clin Exp Res, 2008;32(4):683–9. 112. Brown, J.M., Betters, J.L., Lord, C. et  al. CGI‐58 knockdown in mice
89. Schott, M.B., Rasineni, K., Weller, S.G. et  al. β‐Adrenergic induction of causes hepatic steatosis but prevents diet‐induced obesity and glucose intol­
lipolysis in hepatocytes is inhibited by ethanol exposure. J Biol Chem, erance. J Lipid Res, 2010;51(11):3306–15.
2017;292(28):11815–28. 113. DiStefano, M.T., Danai, L.V., Roth Flach, R.J. et al. The lipid droplet pro­
90. Carr, R.M., Dhir, R., Yin, X., Agarwal, B., and Ahima, R.S. Temporal effects tein hypoxia‐inducible gene 2 promotes hepatic triglyceride deposition by
of ethanol consumption on energy homeostasis, hepatic steatosis, and insulin inhibiting lipolysis. J Biol Chem, 2015;290(24):15175–84.
sensitivity in mice. Alcohol Clin Exp Res, 2013;37(7):1091–9. 114. Monetti, M., Levin, M.C., Watt, M.J. et al. Dissociation of hepatic steatosis
91. Williams, B., Correnti, J., Oranu, A. et al. A novel role for ceramide synthase and insulin resistance in mice overexpressing DGAT in the liver. Cell
6 in mouse and human alcoholic steatosis. FASEB J, 2018;32(1):130–42. Metab, 2007;6(1):69–78.
254 THE LIVER:  REFERENCES

115. Luukkonen, P.K., Zhou, Y., Sädevirta, S. et al. Ceramides dissociate steato­ lycerol stores, VLDL assembly and HCV production. J Hepatol, 2013;
sis and insulin resistance in the human liver in non‐alcoholic fatty liver 59(2):336–43.
disease. J Hepatol, 2016;64(5):1167–75. 126. Vieyres, G., Welsch, K., Gerold, G. et al. ABHD5/CGI‐58, the Chanarin‐
116. Speliotes, E.K., Butler, J.L., Palmer, C.D. et al. PNPLA3 variants specifi­ Dorfman syndrome protein, mobilises lipid stores for hepatitis C virus
cally confer increased risk for histologic nonalcoholic fatty liver disease but production. PLoS Pathog, 2016;12(4):e1005568.
not metabolic disease. Hepatology, 2010;52(3):904–12. 127. Cai, H., Yao, W., Li, L. et  al. Cell‐death‐inducing DFFA‐like effector B
117. Sookoian, S. and Pirola, C.J. Meta‐analysis of the influence of I148M vari­ contributes to the assembly of hepatitis C virus (HCV) particles and inter­
ant of patatin‐like phospholipase domain containing 3 gene (PNPLA3) on acts with HCV NS5A. Sci Rep, 2016;6:27778.
the susceptibility and histological severity of nonalcoholic fatty liver 128. Liu, Y.‐L., Patman, G.L., Leathart, J.B.S. et al. Carriage of the PNPLA3
disease. Hepatology, 2011;53(6):1883–94. rs738409 C >G polymorphism confers an increased risk of non‐alcoholic
118. Pirola, C.J. and Sookoian, S. The dual and opposite role of the TM6SF2‐ fatty liver disease associated hepatocellular carcinoma. J Hepatol,
rs58542926 variant in protecting against cardiovascular disease and confer­ 2014;61(1):75–81.
ring risk for nonalcoholic fatty liver: a meta‐analysis. Hepatology, 129. Krawczyk, M., Stokes, C.S., Romeo, S., and Lammert, F. HCC and liver
2015;62(6):1742–56. disease risks in homozygous PNPLA3 pI148M carriers approach mono­
119. Adinolfi, L.E., Gambardella, M., Andreana, A. et al. Steatosis accelerates genic inheritance. J Hepatol, 2015;62(4):980–1.
the progression of liver damage of chronic hepatitis C patients and corre­ 130. Cholankeril, G., Patel, R., Khurana, S., and Satapathy, S.K. Hepatocellular
lates with specific HCV genotype and visceral obesity. Hepatology, carcinoma in non‐alcoholic steatohepatitis: current knowledge and implica­
2001;33(6):1358–64. tions for management. World J Hepatol, 2017;9(11):533–43.
120. Herker, E., Harris, C., Hernandez, C. et al. Efficient hepatitis C virus parti­ 131. Stickel, F., Buch, S., Nischalke, H.D. et al. Genetic variants in PNPLA3
cle formation requires diacylglycerol acyltransferase‐1. Nat Med, and TM6SF2 predispose to the development of hepatocellular carcinoma in
2010;16(11):1295–8. individuals with alcohol‐related cirrhosis. Am J Gastroenterol, 2018;
121. Camus, G., Herker, E., Modi, A.A. et  al. Diacylglycerol acyltransferase‐1 113(10):1475–83.
localizes hepatitis C virus NS5A protein to lipid droplets and enhances NS5A 132. Straub, B.K., Herpel, E., Singer, S. et al. Lipid droplet‐associated PAT‐
interaction with the viral capsid core. J Biol Chem, 2013; 288(14):9915–23. proteins show frequent and differential expression in neoplastic steatogen­
122. Vogt, D.A., Camus, G., Herker, E. et  al. Lipid droplet‐binding protein esis. Mod Pathol, 2010;23(3):480–92.
TIP47 regulates hepatitis C Virus RNA replication through interaction with 133. Zhu, W., Zhao, Y., Zhou, J. et al. Monoacylglycerol lipase promotes progression
the viral NS5A protein. PLoS Pathog, 2013;9(4):e1003302. of hepatocellular carcinoma via NF‐κB‐mediated epithelial‐mesenchymal
123. Salloum, S., Wang, H., Ferguson, C., Parton, R.G., and Tai, A.W. Rab18 transition. J Hematol Oncol, 2016;9(1):127.
binds to hepatitis C virus NS5A and promotes interaction between sites of 134. Zhang, J., Liu, Z., Lian, Z. et al. Monoacylglycerol lipase: a novel potential
viral replication and lipid droplets. PLoS Pathog, 2013;9(8):e1003513. therapeutic target and prognostic indicator for hepatocellular carcinoma.
124. Camus, G., Schweiger, M., Herker, E. et al. The hepatitis C virus core protein Sci Rep, 2016;6:35784.
inhibits adipose triglyceride lipase (ATGL)‐mediated lipid mobilization and 135. Kochan, K., Maslak, E., Krafft, C. et al. Raman spectroscopy analysis of
enhances the ATGL interaction with comparative gene identification 58 lipid droplets content, distribution and saturation level in non‐alcoholic
(CGI‐58) and lipid droplets. J Biol Chem, 2014; 289(52):35770–80. fatty liver disease in mice. J Biophotonics, 2015;8(7):597–609.
125. Nourbakhsh, M., Douglas, D.N., Pu, C.H. et al. Arylacetamide deacetylase: 136. Guzmán, M. and Castro, J. Zonation of fatty acid metabolism in rat liver.
a novel host factor with important roles in the lipolysis of cellular triacylg­ Biochem J, 1989;264(1):107–13.
Lipoprotein Metabolism
22 and Cholesterol Balance
Mariana Acuña‐Aravena and David E. Cohen
Division of Gastroenterology and Hepatology, Joan & Sanford I. Weill Department of Medicine, Weill Cornell
Medical College, New York, NY, USA

INTRODUCTION and protein content (Table 22.1). As a general rule, larger, less


dense lipoproteins have a greater percentage composition of
Lipids are insoluble or sparingly soluble molecules that are lipids; chylomicrons are the largest and least dense lipoprotein
essential for membrane biogenesis and maintenance of mem- subclass, whereas HDL are the smallest lipoproteins, containing
brane integrity. They also serve as energy sources, hormone the lowest lipid content and the highest proportion of protein.
precursors, and signaling molecules. In order to facilitate trans- Structurally, lipoproteins are microscopic spherical particles
port through the relatively aqueous blood, nonpolar lipids, ranging from 7 to 100 nm in diameter. Lipoprotein particles
such as cholesteryl esters or triglycerides, are packaged within consist of a monolayer of polar, amphipathic lipids surrounding
lipoproteins. a hydrophobic core. Each lipoprotein particle also contains one
Increased concentrations of certain lipoproteins in the circula- or more types of apolipoprotein (Table 22.1). The polar lipids
tion are associated strongly with atherosclerosis. Much of the that comprise the surface coat are unesterified cholesterol and
prevalence of cardiovascular disease, the leading cause of death phospholipid molecules, arranged in a monolayer. The hydro-
in the United States and most Western countries, can be attributed phobic core of a lipoprotein contains the cholesteryl esters
to elevated plasma concentrations of cholesterol‐rich low‐density (cholesterol molecules linked by an ester bond to a fatty acid)
lipoprotein (LDL) particles, as well as lipoproteins that are rich and triglycerides (three fatty acids esterified to a glycerol mol-
in triglycerides. Epidemiologically, decreased concentrations of ecule). The apolipoproteins (also referred to as apoproteins) are
high‐density lipoprotein (HDL) cholesterol also predispose to amphipathic proteins that intercalate into the lipid membrane of
atherosclerotic disease. This chapter highlights the biochemistry lipoproteins. In addition to stabilizing the structure of lipopro-
and ­physiology of cholesterol and lipoproteins. Because abun- teins, apolipoproteins engage in biological functions. They may
dant clinical outcomes data have proven that morbidity and mor- act as receptor ligands for lipoprotein particles or may activate
tality from cardiovascular disease can be reduced by the use of enzymatic activities in the plasma. As will be discussed, the
lipid‐­lowering drugs, mechanisms of pharmacologic interven- apolipoprotein composition determines the metabolic fate of
tions that can ameliorate hyperlipidemia will be discussed. the lipoprotein.
From a metabolic perspective, lipoprotein particles can be
divided into lipoproteins that participate in the delivery of tri-
glyceride molecules to muscle and fat tissue (the apolipoprotein
BIOCHEMISTRY AND PHYSIOLOGY B (apoB)‐containing lipoproteins: chylomicrons and very‐low‐
OF CHOLESTEROL AND LIPOPROTEIN density lipoprotein, VLDL) and lipoproteins that are involved
METABOLISM primarily in cholesterol transport (HDL and the remnants of
apoB‐containing lipoproteins). HDL also serves as a reservoir
Lipoproteins are macromolecular aggregates that transport for exchangeable apolipoproteins in the plasma, including
triglycerides and cholesterol through the blood. Circulating
­ apoA‐I, apoC‐II, and apoE. The following discussion presents
lipoproteins can be differentiated on the basis of density, size, each lipoprotein class in the context of its function.

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
256 THE LIVER:  BIOCHEMISTRY AND PHYSIOLOGY OF CHOLESTEROL AND LIPOPROTEIN METABOLISM

Table 22.1  Characteristics of plasma lipoproteins


CM VLDL IDL LDL HDL
Density (g mL ) −1
<0.95 0.95–1.006 1.006–1.019 1.019–1.063 1.063–1.210
Diameter (nm) 75–1200 30–80 25–35 18–25 5–12
Total lipid (% wt) 98 90 82 75 40
Composition, % dry weight
Protein 2 10 18 25 33
Triglycerides 83 50 31 9 8
Unesterified cholesterol + 8 22 29 45 30
cholesterol ester
Phospholipids (% wt lipid) 7 18 22 21 29
Electrophoretic mobilitya None Pre‐β β β α or Pre‐β
Major apolipoproteins B48, A‐I, A‐IV, E, B100, E, C‐I, B100, E, C‐I, B100 A‐I, A‐II, C‐I,
C‐I, C‐II, C‐III C‐II, C‐III C‐II, C‐III C‐II, C‐III, E
CM, chylomicron; VLDL, very‐low‐density lipoprotein; IDL, intermediate‐density lipoprotein; LDL, low‐density lipoprotein; HDL, high‐density lipoprotein.
a
 Electrophoretic mobility of lipoprotein particles is designated relative to migration of plasma α‐ and β‐globulins.
Adapted from [103] with permission of Elsevier.

Metabolism of apoB‐containing lipoproteins Once apoB has been fully translated, the nascent lipoprotein is
enlarged in the Golgi apparatus, during which MTP adds addi-
The primary function of apoB‐containing lipoproteins is to deliver tional triglycerides to the core of the particle [4]. MTP also
fatty acids in the form of triglycerides to muscle tissue to be used enhances rates of cholesterol esterification by removing choles-
for ATP biogenesis and to adipose tissue for long‐term storage. teryl esters from their site of synthesis in the endoplasmic retic-
Chylomicrons are formed in the intestine and transport dietary ulum and transferring them into nascent apoB lipoproteins [5].
triglycerides, whereas VLDL particles are formed by the liver and This assembly process produces the lipoprotein particles, each
transport triglycerides that are synthesized endogenously. For of which contains a single molecule of apoB.
conceptual purposes, the metabolic lifespan of apoB‐containing The diet is the main source of triglycerides in chylomicrons
lipoproteins can be divided into three phases: assembly, intravas- (Figure 22.1b), so their assembly, secretion, and metabolism are
cular metabolism, and receptor‐mediated clearance. This is also a collectively referred to as the exogenous pathway of lipoprotein
convenient categorization because pharmacologic agents are metabolism. By contrast, cholesteryl esters in chylomicrons are
available that influence these different phases. derived mainly from biliary cholesterol (approximately 75%),
with the remainder contributed by dietary sources. During
Assembly of apoB‐containing lipoproteins
digestion, cholesteryl esters and triglycerides in food are hydro-
The cellular mechanisms by which chylomicrons and VLDL are lyzed by lipases secreted from the pancreas to form unesterified
assembled are quite similar. Regulation of the assembly process cholesterol, free fatty acids, and monoglycerides [3]. Bile salts,
depends upon the availability of apoB and triglycerides, as well as phospholipids, and cholesterol are secreted by the liver into bile
the activity of microsomal triglyceride transfer protein (MTP) [1]. and stored in the gallbladder during fasting as micelles and vesi-
The gene that encodes apoB is transcribed principally in the cles, which are macromolecular lipid aggregates that form due
intestine and the liver. Apart from this tissue‐specific expres- to the detergent properties of bile salt molecules. The stimulus
sion, there is little transcriptional regulation of the apoB gene. of eating a meal promotes emptying of gallbladder bile into the
By contrast, a key regulatory event that differentiates chylomi- small intestine, where micelles and vesicles solubilize the
cron from VLDL metabolism is editing of apoB mRNA. Within digested lipids. Lipid absorption into enterocytes of the duode-
enterocytes but not hepatocytes, a protein named apoB‐editing num and jejunum is facilitated mainly by micelles [6]. Long‐
complex 1 (apobec‐1) is expressed. This protein constitutes the chain fatty acids and monoglycerides are taken up separately
catalytic subunit of the apoB‐editing complex, which deami- into the enterocyte by carrier‐mediated transport and then re‐
nates a cytosine at position 6666 of the apoB mRNA molecule esterified to form triglycerides by the enzyme diacylglycerol
[2]. This converts the cytosine to a uridine base. As a result, the acyltransferase (DGAT) [3]. By contrast, medium‐chain fatty
codon containing this nucleotide is converted from a glutamine acids are absorbed directly into the portal blood and metabo-
to a premature stop. When translated, the intestinal form apoB48 lized by the liver. The steps of cholesterol absorption are mecha-
is 48% as long as the full‐length protein that is expressed in the nistically defined [7]. Dietary and biliary cholesterol from
liver and referred to as apoB100. As a consequence, chylomi- micelles enter the enterocyte via a protein channel named
crons, the apoB‐containing lipoprotein produced by the intes- Niemann-Pick C1-like 1 protein (NPC1L1). Some of this cho-
tine, contain apoB48. By contrast, VLDL particles produced by lesterol is immediately pumped back into the intestinal lumen
the liver contain apoB100. by the ATP‐dependent action of a heterodimeric protein, ATP‐
Figure  22.1a illustrates the cellular mechanisms for the binding cassette (ABC) G5/ABCG8 (ABCG5/G8) expressed
assembly and secretion of apoB‐containing lipoproteins. As in enterocytes and hepatocytes [8]. The fraction of cholesterol
the  apoB protein is translated by ribosomes, it crosses into that remains is esterified to a long‐chain fatty acid by acyl‐
the ­endoplasmic reticulum (ER) [3, 4]. Within the ER, triglycer- CoA:cholesterol acyltransferase (ACAT). Once triglycerides
ide molecules are added co‐translationally to the elongating and cholesteryl esters are packaged together with apoB48,
apoB protein (i.e. apoB is lipidated) by the action of MTP [3]. apoA‐IV is added to stabilize the surface, allowing additional
22:  Lipoprotein Metabolism and Cholesterol Balance 257

(a)
Enterocyte or hepatocyte

Cytosol
Presence of triglycerides

Lipidation
Ribosome CM

MTP MTP
VLDL
ApoB

Absence of triglycerides

Degradation

ApoB Endoplasmic reticulum

(b)
Cholesterol absorption Triglyceride absorption
LYMPH ENTEROCYTE INTESTINAL LYMPH ENTEROCYTE INTESTINAL
LUMEN LUMEN

Cholesterol 2 Fatty acid


+
NPC1L1 Monoglyceride
ACAT

DGAT
Cholesteryl
ABCG5/G8 Exogenous
ester
triglyceride
Bile salt

(c)
Chylomicron formation
LYMPH ENTEROCYTE INTESTINAL
LUMEN
Exogenous
CM triglycerides
apoB48

Cholesteryl
ester

Figure 22.1  Assembly and secretion of apolipoprotein B‐containing lipoproteins. (a) Chylomicron (CM) and VLDL particles are assembled and
secreted by similar mechanisms in the enterocyte and hepatocyte, respectively. The apoB protein (i.e. apoB48 or apoB100) is translated by ribo-
somes and enters the lumen of the endoplasmic reticulum. If triglycerides are available, the apoB protein is lipidated by the action of microsomal
triglyceride transfer protein (MTP) in two distinct steps, accumulating triglyceride as well as cholesteryl ester molecules. The resulting CM or
VLDL particle is secreted by exocytosis into the lymphatics by enterocytes or into the plasma by hepatocytes. In the absence of triglycerides, the
apoB protein is degraded. Modified from [101] with permission of John Wiley & Sons. (b) Exogenous triglycerides and cholesterol absorption are
simultaneously absorbed from the intestinal lumen by different mechanisms. Cholesterol is taken up from micelles across a regulatory channel
named NPC1L1. A fraction of the cholesterol is pumped back into the lumen by ABCG5/G8, a heterodimeric ATP‐dependent plasma membrane
protein. The remainder of the cholesterol is converted to cholesteryl esters by ACAT. Triglycerides are taken up as fatty acids and monoglycerides,
which are re‐esterified by DGAT. Modified from [102] with permission of Wolters Kluwer. (c) Exogenous triglycerides plus cholesteryl esters are
assembled into CM within the enterocyte.
258 THE LIVER:  BIOCHEMISTRY AND PHYSIOLOGY OF CHOLESTEROL AND LIPOPROTEIN METABOLISM

core lipidation, and subsequently apoA‐I is attached [9]. The part by aqueous transfer of apoC‐II from HDL particles [19].
assembled chylomicron particle is then exocytosed into the lym- Because there is an inherent delay in the transfer of apoC‐II to
phatics for transport into the circulation via the thoracic duct chylomicrons and VLDL particles, this allows time for
(Figure  22.1c). The plasma concentration of triglyceride‐rich ­widespread circulation of triglyceride‐rich particles throughout
chylomicrons varies in proportion to dietary fat intake. the body.
VLDL particles comprise triglycerides that are assembled by Lipoprotein lipase (LPL) is a lipolytic enzyme that is primar-
the liver using plasma fatty acids derived from adipose tissue or ily synthesized and secreted by a variety of parenchymal cells,
synthesized by de novo lipogenesis. For this reason, the assem- including myocytes, adipocytes, skeletal muscle cells, mac-
bly, secretion, and metabolism of VLDL particles are often rophages, and mammary gland cells [20, 21]. This glycoprotein
referred to as the endogenous pathway of lipoprotein metabo- is transported from the interstitial space to endothelial cell sur-
lism. Hepatocytes synthesize triglycerides in response to face of the luminal capillary in muscle and fat tissues by the
increased free fatty acid flux to the liver. This typically occurs in glycosylphosphatidylinositol (GPI)‐linked protein (GPIHBP1)
response to fasting, thereby insuring a continuous supply of [22]. GPIHBP1 also serves to anchor LPL to the endothelial cell
fatty acids for delivery to muscle in the absence of triglycerides plasma membrane. Once chylomicrons and VLDL particles
from the diet. Dietary saturated fats as well as carbohydrates acquire apoC‐II, they can bind to LPL, which hydrolyzes tri-
also stimulate the ­synthesis of triglycerides within the liver [10, glycerides from the core of the lipoprotein (Figure 22.2b). LPL‐
11]. By cellular mechanisms that are similar to those that pro- mediated lipolysis liberates free fatty acids and glycerol, which
duce chylomicrons, MTP in hepatocytes lipidates apoB100 to are then taken up by the neighboring parenchymal cells. Because
form nascent VLDL particles. Under the continued influence of the expression level and intrinsic activity of LPL in muscle
MTP, the nascent VLDL particles coalesce with larger triglycer- ­versus fat tissue is regulated according to the fed/fasting state,
ide droplets and are secreted directly into the circulation. VLDL this allows the body to direct the delivery of fatty acids prefer-
particles may also acquire apoE, apoC‐I, apoC‐II, and apoC‐III entially to muscle during fasting and to fat post prandially [21].
within the hepatocyte prior to secretion. However, these apoli-
poproteins may also be transferred to VLDL from HDL during
(a)
their circulation through the bloodstream.
The translation of apoB48 in the intestine and apoB100 in the VLDL CM
liver is constitutive. This permits the immediate production of LIVER apoB100 apoB48 INTESTINE
chylomicrons and VLDL particles when triglyceride molecules
are available. In the absence of triglycerides, such as in entero-
cytes during fasting, apoB is degraded by a variety of cellular
mechanisms [12].
Recent evidence from whole‐exome sequencing, exome chip, apoE apoC-II
and genome‐wide association studies (GWAS) have identified
new genes that regulate apoB‐containing lipoprotein metabo- α-HDL apoAI
lism [13]. One of these is SORT1, which encodes the lysosomal
trafficking protein sortilin‐1. A single nucleotide polymorphism
(SNP) which increases the hepatic expression of sortilin‐1 is (b) Capillary in
associated with reduced plasma levels of LDL cholesterol. This
muscle or adipose tissue
is partially explained because sortilin‐1 facilitates the post‐
transcriptional degradation of apoB by a lysosomal‐dependent
VLDL CM
pathway, leading to decreased VLDL secretion [14]. Another apoB48
apoB100
SNP found in the gene encoding transmembrane 6 superfamily
member 2 (TM6SF2), which reduces the protein level, is associ-
ated with low plasma triglyceride and cholesterol concentra-
tions, but also non‐alcoholic fatty liver disease (NAFLD). This
is because TM6SF2 decreases VLDL concentrations in the Fatty Fatty
acids Lipoprotein acids
plasma, thereby increasing hepatic triglyceride contents [15, 16].
lipase
Additional studies have revealed that the lowered TM6SF2 pro-
tein levels impair apoB lipidation and VLDL assembly, but not
its secretion [17]. Similar to TM6SF2, a loss‐of‐function SNP in Endothelium
patatin‐like phospholipase domain‐containing 3 (PNPLA3) is Figure 22.2  Intravascular metabolism of apoB‐containing lipoproteins.
also associated with NAFLD, and impairs VLDL secretion [18]. (a) Following secretion, chylomicron (CM) and VLDL particles are
­activated for lipolysis when they encounter HDL particles in the plasma
and acquire the exchangeable apolipoprotein apoC‐II. (b) When CM and
Intravascular metabolism of apoB‐containing lipoproteins VLDL particles circulate into capillaries of muscle or fat tissue, apoC‐II
promotes binding of the particle to lipoprotein lipase, which is bound to
Within the circulation, chylomicrons and VLDL particles must
the surface of endothelial cells. Lipoprotein lipase mediates hydrolysis of
be activated in order to target triglyceride delivery to muscle and triglycerides, but not cholesteryl esters, from the core of the lipoprotein
fat tissues (Figure 22.2a). This requires the addition of an opti- particle. The resulting fatty acids are taken up into muscle or fat tissue.
mal complement of apoC‐II molecules, which occurs at least in Modified from [101] with permission of John Wiley & Sons.
22:  Lipoprotein Metabolism and Cholesterol Balance 259

The rate of lipolysis of chylomicrons and VLDL triglycerides is (a)


VLDL
VLDL CM
CM
also controlled by apoC‐III, which is an inhibitor of LPL a­ ctivity remnant remnant
apoB100 apoB48
[23]. Antagonism of LPL activity by apoC‐III may be an apoB100 apoB48
additional mechanism promoting widespread distribution of
­
­triglyceride‐rich particles in the circulation.
apoCII

Receptor‐mediated clearance of apoB‐containing apoE


lipoproteins
α-HDL apoAI
As LPL continues to hydrolyze triglycerides from chylomi-
crons and VLDL, the particles become progressively depleted
of ­triglycerides and relatively enriched with cholesterol. Once (b) Sinusoidal endothelium CM or VLDL remnant
approximately 50% of triglycerides have been removed, the
particles lose affinity for LPL and dissociate. The exchangeable HSPG
apolipoproteins apoA‐I and apoC‐II (as well as apoC‐I and Space
apoC‐III) are then transferred to HDL in exchange for apoE Sequestration of
(Figure 22.3a) [24], which serves as a high‐affinity ligand for Disse
Hepatic lipase
receptor‐mediated clearance [25]. Upon acquiring apoE, the
particles are termed chylomicron or VLDL remnants [26]. As
Lipolysis
explained below, intermediate‐density lipoproteins (IDLs) are
also remnant particles.
Remnants of chylomicrons and VLDL particles, as well as
some IDL particles, are taken up by the liver in a three‐step
Uptake LDLr LRP
process (Figure  22.3b) [26]. The first step is sequestration HSPG
Hepatocyte
of  the particles within the space of Disse between the fenes-
trated endothelium of the liver sinusoids and the sinusoidal Figure 22.3  Formation and hepatic uptake of remnant particles. (a)
(­basolateral) plasma membrane of hepatocytes. Sequestration Upon completion of hydrolysis, chylomicron (CM) and VLDL particles
requires that the remnant particles become small enough dur- lose affinity for lipoprotein lipase. When an HDL particle is encoun-
ing the process of lipolysis to fit in between the endothelial tered, apoC‐II is transferred back to HDL particles in exchange for
cells. Once in the space of Disse, remnants are bound and apoE. The resulting particles are CM and VLDL remnants and interme-
sequestered by large heparan sulfate proteoglycans (HSPGs). diate‐density lipoproteins (IDLs), which are VLDL particles that inter-
act for prolonged periods with LPL to become more dense. (b) The
The next step is remodeling within the space of Disse by the activity of lipoprotein lipase results in remnant lipoprotein particles that
action of hepatic lipase, a lipolytic enzyme that is similar to are small enough in size to enter the space of Disse. Remnant lipopro-
LPL, but is expressed by hepatocytes. Hepatic lipase appears to teins are sequestered in the space of Disse by binding to high molecular
optimize the triglyceride content of remnant particles, so they weight heparin sulfate proteoglycan (HSPG) molecules. This is fol-
are efficiently cleared by receptor‐mediated mechanisms. The lowed by binding of hepatic lipase, which promotes lipolysis of some
final phase of remnant clearance is receptor‐mediated uptake residual triglycerides in the core of the remnant lipoproteins and the
release of fatty acids (as indicated by the dashed arrow). Uptake of rem-
[27]. This is accomplished by one of four pathways. At the
nant lipoprotein particles into hepatocytes is mediated by the LDL
sinusoidal hepatocyte plasma membrane, remnant particles receptor (LDLr), the LDL‐related protein (LRP), a complex formed
may be bound and taken up by the LDL receptor, the LDL between LRP and HSPG, or HSPG alone. Modified from [101] with
receptor‐related protein (LRP), or by HSPGs. A separate path- permission of John Wiley & Sons.
way is mediated by the combined activities of LRP and HSPGs.
These efficient and redundant mechanisms allow for efficient
particle clearance, so that the half‐life of remnants in the remainder are converted to LDL by hepatic lipase, which further
plasma is approximately 30 minutes. hydrolyzes triglycerides in the core of IDL. The further reduc-
tion in size of the particles results in the transfer of apoE to
HDL. As a result, LDLs are distinct, cholesteryl ester‐enriched
Formation and clearance of LDL particles
lipoproteins with apoB100 as their only apolipoprotein [29].
Whereas apoB48‐containing chylomicron remnants are com- The LDL receptor is the primary receptor capable of clearing
pletely cleared from the plasma, the presence of apoB100 alters significant amounts of LDL from the plasma [30]. It is concen-
the metabolism of VLDL remnants so that only approximately trated on the surface of hepatocytes, macrophages, lympho-
50% are cleared by the pathways for remnant particles. The dif- cytes, adrenocortical cells, gonadal cells, and smooth muscle
ference begins when the other 50% are metabolized to a greater cells. Due to the lack of apoE, the LDL particles are relatively
extent by LPL, becoming smaller and relatively deficient in tri- weak ligands for the LDL receptor [30]. As a result, the half‐life
glycerides and enriched in cholesteryl esters. When converted to of LDL in the circulation is markedly prolonged (2–4 days).
remnants following exchange of apolipoproteins with HDL, This explains why LDL cholesterol accounts for approximately
these more dense particles become IDL. Because IDL particles 65–75% of total plasma cholesterol.
contain apoE, some of them may be cleared into the liver by Interaction of apoB100 with the LDL receptor facilitates
remnant receptor pathways (Figure  22.3b) [28]. However, the receptor‐mediated endocytosis of LDL particles and subsequent
260 THE LIVER:  BIOCHEMISTRY AND PHYSIOLOGY OF CHOLESTEROL AND LIPOPROTEIN METABOLISM

vesicle fusion with a lysosome [30]. The LDL receptor is recy- destabilize atherosclerotic plaques, in part due to the liberation
cled to the cell surface, while the LDL particle is hydrolyzed of matrix metalloproteinases which may precipitate acute
within the lysosome by lysosomal acid lipase to release unest- ischemic cardiovascular events in the form of myocardial infarc-
erified cholesterol and free fatty acids. LDL‐derived cholesterol tions and strokes. Moreover, Lp(a) exhibits prothrombotic/anti‐
is released from lysosomes through the action of two proteins, fibrinolytic effects, which can accelerate atherosclerosis and
the membrane‐bound Niemann–Pick C1 (NPC1) and the solu- lead to the intimal deposition of Lp(a) cholesterol [39].
ble (NPC2) [31]. Recent studies suggest that cholesterol is
transported first to replenish the plasma membrane in order to
maintain optimal levels, and the excess is then trafficked to the
HDL metabolism and reverse cholesterol
ER, where it impacts three major homeostatic pathways [32]. transport
First, intracellular cholesterol inhibits 3‐hydroxy‐3‐methylglu- Virtually all cells in the body are capable of synthesizing all of
taryl‐coenzyme A reductase (HMG‐CoA reductase), the enzyme the cholesterol they require. However, only the liver has the
that catalyzes the rate‐limiting step in de novo cholesterol capacity to eliminate cholesterol by secretion into bile in its unes-
­synthesis. Second, cholesterol activates ACAT to increase ester- terified form or following its conversion to bile salts. In addition
ification and storage of cholesterol in the cell. Third, LDL to serving as a reservoir for exchangeable apolipoproteins for the
receptor expression is downregulated, reducing further uptake metabolism of apoB‐containing lipoproteins, HDLs are lipopro-
of cholesterol into the cells. The majority of LDL receptors teins that play a key role in cholesterol homeostasis by removing
(approximately 70%) are expressed on the surface of hepato- excess cholesterol from cells and transporting it through plasma
cytes. As a result, the liver is primarily responsible for the to the liver. This process is often referred to as reverse cholesterol
removal of LDL particles from the circulation. Reductions in transport (RCT) (Figure 22.4a) [40]. The process of RCT protects
intracellular cholesterol levels, such as occur during therapy against the development of atherosclerosis by removing excess
with statin drugs, lead to LDL receptor upregulation. This is due cholesterol from plaque macrophages [41]. Inflammatory and
to the proteolytic processing and activity of sterol regulatory oxidative damage also play pivotal roles in atherogenesis, and
element‐binding protein 2 (SREBP‐2) [33]. certain HDL particle populations have been shown to be pro­
Sortilin‐1 also promotes the uptake of LDL by LDL receptor‐ tective by preventing oxidation of LDL particles and inhibiting
dependent and ‐independent pathways. In the independent path- endothelial expression of adhesion molecules [42]. Moreover,
way, sortilin‐1 serves as a cell surface receptor for LDL, facilitating epidemiological studies have shown that low plasma levels of
its cellular uptake and lysosomal degradation [14]. Another HDL cholesterol are associated with an increased risk in cardio-
important regulator of LDL receptor clearance is the proprotein vascular disease. Therefore, high plasma concentrations of HDL
convertase subtilisin/kexin type 9 (PCSK9). PCSK9 is a serine cholesterol in plasma are considered atheroprotective [43].
protease that binds LDL receptor and induces its degradation in HDL particles are heterogeneous in structure and biological
lysosomes via a mechanism that requires neither ubiquitination activity, and have been classified by physical properties into a
nor autophagy [34]. Loss‐of‐function mutations in PCSK9 result variety of subclasses [44]. These particles contain 14 different
in marked decreased plasma LDL cholesterol concentrations. apolipoproteins, the major ones being apoA‐I and apoA‐II [45].
A variant of LDL is produced when the glycoprotein apo(a) is ApoA‐I is the main structural determinant of HDL, and partici-
covalently attached to apoB100 by a disulfide bond, generating pates in the formation, as well as interaction of the particle with
lipoprotein(a) or Lp(a) [35]. Apo(a) contains from 3 to more its receptor, scavenger receptor class B type I (SR‐BI) [40, 46].
than 50 plasminogen‐like domains, which give rise to a hetero- ApoA‐I is also important for the antioxidant properties of HDL
geneous population of Lp(a) particles. The physiological func- [47]. ApoA‐II is more hydrophobic than apoA‐I, which helps
tions and catabolism of Lp(a) remain under study [36]. to confer the lipid surface curvature in HDL maturation and to
High plasma concentrations of LDL and Lp(a) particles stabilize the particle [48].
­constitute risk factors for the development of atherosclerotic
cardiovascular disease [37]. LDL particles that are not taken up
HDL formation
by LDL receptor may penetrate the intima of blood vessels and
bind to proteoglycans. There, they are subject to oxidation, HDL formation occurs mainly in the liver, although a percent-
which results in lipid peroxidation and may create reactive alde- age is contributed by the small intestine [43]. Unlike other lipo-
hyde intermediates that fragment apoB100. The modified LDL proteins, HDL particles are not assembled in cells; instead they
is internalized by scavenger receptors (e.g. SR‐A), which are are formed in the extracellular space and mature in the plasma
expressed predominantly by mononuclear phagocytic cells [38]. as they are remodeled. The earliest events occur when lipid‐poor
Unlike the LDL receptor, scavenger receptors are not downreg- apoA‐I is secreted by the liver or intestine [49], or dissociates
ulated when the phagocytic immune cells begin to accumulate from lipoprotein particles in the plasma [50]. These amphip-
cholesterol. As a result, the continued accumulation of oxidized athic apoA‐I molecules interact with ABCA1, which is imbed-
LDL in macrophages can lead to foam cell formation (choles- ded in the sinusoidal membrane of hepatocytes or basolateral
terol‐rich macrophages). Oxidized LDL also causes upregula- membrane of enterocytes [51, 52]. ABCA1 incorporates a small
tion of cytokine production, impairs endothelial function, and amount of membrane phospholipids and unesterified choles-
increases expression of endothelial adhesion molecules. All of terol into the apoA‐I molecule [52]. The resulting small disk‐
these effects increase the local inflammatory response and shaped HDL particle consists mainly of phospholipid and
­promote atherosclerosis. Foam cells are a major constituent of apoA‐I. The particle that is formed is referred to as nascent or
atherosclerotic lesions, and excessive foam cell death can pre‐β‐HDL, due to its characteristic migration on agarose gels.
22:  Lipoprotein Metabolism and Cholesterol Balance 261

(a)
LIVER BLOOD TISSUES

Cholesterol LCAT
α-HDL
pre-β-HDL
Phospholipid
apoA-I
ABCA1

Cholesterol
SR-BI
LCAT
PLTP
Hepatic CETP
lipase

(b)

CM or VLDL
Albumin
remnant

Lyso-PC
PLTP
LCAT

Cell
CETP

α -HDL

Figure 22.4  Reverse cholesterol transport. (a) The process of reverse cholesterol transport begins when apoA‐I is secreted from the liver. ApoA‐I
in plasma interacts with ATP‐binding cassette protein AI (ABCA1), which incorporates a small amount of phospholipid and unesterified cholesterol
from hepatocyte plasma membranes to form a discoidal‐shaped pre‐β‐HDL particle. Due to the activity of lecithin:cholesterol acyltransferase
(LCAT) in plasma, pre‐β‐HDL particles mature to form spherical α‐HDL. Spherical α‐migrating HDL particles function to accept excess unesterified
cholesterol from the plasma membranes of cells in a wide variety of tissues. The unesterified cholesterol is transferred from the cell to nearby HDL
particles by diffusion through the plasma. As shown in (b), LCAT and phospholipid transfer protein (PLTP) increase the capacity of HDL to accept
unesterified cholesterol molecules from cells by allowing for expansion of the core and the surface coat of the particle. Cholesteryl ester transfer
protein (CETP) exchanges some of the cholesteryl ester molecules from the core of HDL for triglycerides from the core of remnant particles. HDL
particles interact with scavenger receptor, class B type I (SR‐BI), which mediates selective hepatic uptake of cholesteryl esters, but not apoA‐I. This
process is facilitated when hepatic lipase hydrolyzes triglycerides from the core of the particle. The remaining apoA‐I molecules may begin the cycle
of reverse cholesterol transport again. In (a), solid arrows indicate metabolic events in HDL metabolism, whereas dashed arrows denote transfer of
molecules. (b) LCAT, PLTP, and CETP promote the removal of excess cholesterol from the plasma membranes of cells. LCAT removes a fatty acid
from a phosphatidylcholine molecule in the surface coat of α‐HDL (or pre‐β‐HDL) and esterifies an unesterified cholesterol molecule on the surface
of the particle. The resulting lysophosphatidylcholine (lyso‐PC) becomes bound to albumin in the plasma, whereas the cholesteryl ester migrates
spontaneously into the core of the lipoprotein particle. The unesterified cholesterol molecules that are consumed by LCAT are replaced by unesteri-
fied cholesterol from cells. HDL phospholipids that are consumed by LCAT action are replaced with excess phospholipids from remnant particles
by the activity of PLTP. As described in (a), CETP increases the efficiency of cholesterol movement to the liver by exchanging cholesteryl ester
molecules from the core of α‐HDL for triglycerides from the core of remnant particles. In (b), solid arrows denote protein‐mediated lipid transfer,
whereas dashed arrows indicate that lipids move by diffusion through the plasma. Modified from [101] with permission of John Wiley & Sons.

Intravascular maturation of HDL


occurs due to the activity of two distinct circulating proteins
Because disk‐shaped pre‐β‐HDL particles are relatively ineffi- (Figure 22.4b). Lecithin:cholesterol acyltransferase (LCAT) binds
cient at removing excess cholesterol from cell membranes, they preferentially to disk‐shaped HDL particles and converts choles-
must mature into spherical particles within the plasma. This terol molecules within the particle to cholesteryl esters [53].
262 THE LIVER:  BIOCHEMISTRY AND PHYSIOLOGY OF CHOLESTEROL AND LIPOPROTEIN METABOLISM

This is accomplished by transesterification of a fatty acid from a PLTP, which prevent the surface coat of the particle from
phosphatidylcholine molecule on the surface of the HDL to the becoming saturated with unesterified cholesterol from cells.
hydroxyl group of a cholesterol molecule. This two‐step reaction
also creates a lysophosphatidylcholine molecule, which departs Delivery of HDL cholesterol to the liver
from the particle upon binding to serum albumin [53]. Because
they are highly insoluble, cholesteryl esters produced by LCAT When mature HDL particles circulate to the liver, they interact
spontaneously migrate into the core of the HDL particle. The with SR‐BI, the principal HDL receptor [57], which is expressed
development of a hydrophobic core converts the pre‐β‐HDL to a on the sinusoidal plasma membranes of hepatocytes. Although
spherical HDL particle, which exhibits α migration on agarose it can mediate cholesterol efflux from cells with excess choles-
gels [45]. terol, SR‐BI in the liver promotes selective uptake of lipids, a
The second important protein that contributes to HDL matura- process whereby the cholesterol and cholesteryl esters of HDL
tion in the plasma is phospholipid transfer protein (PLTP), which particles are taken up into the cell, in the absence of uptake of
is mainly expressed in adipose tissue, lung and liver [54]. PLTP apolipoproteins [46]. During SR‐BI‐mediated selective lipid
transfers phospholipids from the surface coat of apoB‐contain- uptake, apoA‐I is liberated to participate in pre‐β‐HDL forma-
ing remnant particles to the surface coat of HDL by penetrating tion. As a result, the “lifespan” of an HDL particle is 2–5 days,
into the surface of both lipoproteins and forming a ternary com- suggesting that each apoA‐I molecule can participate in many
plex [55]. During LPL‐mediated lipolysis of apoB‐containing cycles of RCT. Among the tissues that express high levels of
lipoproteins, the particles become smaller as triglycerides are SR‐BI are the adrenal glands and gonads, presumably reflecting
removed from the core. This leaves a relative excess of phospho- their requirement for cholesterol in order to support steroido-
lipids on the surface of the particle. Because phospholipids are genesis [58].
insoluble and cannot otherwise dissociate from a particle, PLTP Delivery of cholesterol from extrahepatic tissues to the liver
removes excess phospholipids and thereby maintains the appro- is optimized by two additional proteins, cholesteryl ester trans-
priate surface concentration for the shrinking core. By transfer- fer protein (CETP) and hepatic lipase. CETP is a plasma protein
ring phospholipids to the surface of HDL, PLTP also replaces the that transfers cholesteryl esters from mature spherical HDL par-
molecules that are consumed by the LCAT reaction. This allows ticles to the cores of remnant lipoproteins in exchange for a tri-
the core of HDL to continue to enlarge. Based on studies on mice glyceride molecule, which is inserted into the core of the HDL
lacking PLTP, it has been demonstrated that the majority of HDL particle (Figure  22.4b) [59]. This process allows the body to
phospholipids are transferred from the surface coat of remnant utilize remnant particles that have completed their function of
particles [54]. triglyceride transport for purposes of transporting cholesterol to
the liver. Removal of cholesteryl ester molecules from HDL
appears to serve two functions. First, it further increases the
HDL‐mediated cholesterol efflux from cells capacity of HDL to take on additional cholesterol molecules
Cellular cholesterol efflux is the mechanism by which excess from cells. Second, it makes the process of selective uptake by
insoluble cholesterol molecules are removed from cells. This SR‐BI more efficient [59]. This is because hydrolysis of triglyc-
occurs when unesterified cholesterol is transferred from the erides by hepatic lipase on the hepatocyte surface facilitates the
plasma membrane of cells to an HDL particle. The mechanism activity of SR‐BI (Figure 22.4a).
of cholesterol efflux varies depending upon the cell type and the There are conflicting data concerning the net effects of CETP
type of HDL particle [56]. Lipid‐poor pre‐β‐HDL particles can on lipid metabolism. Studies in humans have suggested that
promote cholesterol efflux by interacting with ABCA1. In addi- CETP may improve cholesterol efflux by increasing the capac-
tion to beginning the process of HDL formation by the liver, this ity of HDL particles to take up cholesterol from macrophages,
is also an active mechanism for removing excess cholesterol preventing lipid accumulation in these cells [60]. On the other
from cells that serves to protect macrophages within the suben- hand, the observation that deleterious gene mutations in CETP
dothelial space from cholesterol‐induced cytotoxicity. Besides gene had the effect of increasing plasma HDL cholesterol and
ABCA1‐mediated cholesterol efflux, spherical HDL particles apoA‐I concentrations in humans with no evidence for prema-
very efficiently stimulate cholesterol export by three other path- ture atherosclerosis suggests that CETP inhibition could repre-
ways [56]. First, a passive mechanism includes simple diffusion sent a promising therapeutic approach for cardiovascular disease
via the aqueous phase in the absence of binding to a specific cell [43]. This notion was supported by the selective beneficial
surface protein. Although cholesterol has very low monomeric effects of certain hypolipidemic drugs observed in in rodents,
solubility, it can dissociate in appreciable amounts and travel which are naturally deficient in CETP, but not in humans [61].
short distances through the plasma to acceptor particles that Although CETP inhibitors were advanced to phase 3 clinical
are enriched with phospholipids on their surfaces. The second trials, an apparent lack of benefit along with concerns around
is facilitated diffusion mediated by SR‐BI, whereby HDL parti- toxicity prevented any candidate drugs from coming to the
cles interact with SR‐BI on the plasma membrane. promoting market.
­cholesterol efflux. Finally, an active efflux of cholesterol to
spherical HDL particles is mediated by ABCG1, which is
expressed in macrophages. Quantitatively, efflux to spherical
Biliary lipid secretion
HDL particles accounts for most of the removal of excess cho- Once cholesterol is delivered to the liver, it is eliminated by
lesterol from cells. This capacity of HDL to remove cellular ­biliary secretion. A key essential step occurs  when a fraction
cholesterol is enhanced through the activities of LCAT and of  the cholesterol is converted to bile salts. The enzymatic
22:  Lipoprotein Metabolism and Cholesterol Balance 263

convertion into bile salt molecules catalyzed by CYP7A1 is fractional loss of bile salts amounts to about 0.4 g per day [71].
referred to as the classical pathway; an alternative pathway is Considering that cholesterol is the substrate for bile salt synthe-
initiated by the sterol 27‐hydroxylase (CYP27A) followed by sis, fecal bile salts represent a source of cholesterol loss from
oxysterol 7α‐hydroxylase (CYP7B1) [62]. In humans the alter- the body. Sensitive nuclear hormone receptors within the liver
native pathway contributes only a minor proportion of bile salts are capable of detecting the rate of loss of bile salts into the
under normal circumstances, whereas in rodents the classical feces [63]. These receptors tightly regulate transcription of bile
and alternative pathways contribute about equally to bile acid salt synthetic genes. As a result, the liver synthesizes precisely
synthesis [63]. Bile salts, unlike cholesterol, are highly soluble an amount of bile salts that is sufficient to replace what is lost in
in water. Moreover, bile salts are biological detergents that pro- the feces. The gut microbiota also regulates bile acids synthesis,
mote the formation of micelles. These macromolecular aggre- bile acid pool size and composition, and enterohepatic circula-
gates are rich in phospholipids, which are derived from tion of bile acids [72].
hepatocyte membranes and solubilize cholesterol in bile for On a daily basis, an average of 24 g of bile salts are secreted
transport from the liver to the small ­intestine, essentially func- [71], together with 11 g of phospholipids [73] and approxi-
tioning as counterparts to HDL particles in plasma [64]. mately 0.9 g of cholesterol. TICE accounts for about 0.7 g of
Bile formation begins when ABCB11, an ATP‐driven canali- cholesterol, and the average American diet contributes approxi-
cular membrane transporter, functions to pump bile salts into mately 0.4 g per day to intestinal cholesterol. Therefore, dietary
bile [65]. Bile salts within bile then activate two other ATP‐ cholesterol represents only a minor fraction (20%) compared to
dependent transporters on the canalicular membrane of the the endogenous (i.e. biliary) cholesterol that passes through the
hepatocyte: ABCB4 and the heterodimer ABCG5/G8 [66]. intestine [74]. The extent to which intestinal cholesterol is
These transporters promote the secretion of phospholipid and absorbed appears to be genetically regulated. Each individual
cholesterol molecules, respectively, which together with bile absorbs a fixed percentage of intestinal cholesterol. In the popu-
salts form mixed micelles. Biliary lipids are stored in the gall- lation, percentages range from as low as 20% to more than 80%
bladder during fasting. The stimulus of a fatty meal leads to [75, 76]. For example, when an average individual absorbs
gallbladder contraction, which propels its contents into the 50%, this will amount to half of the 2.0 g (i.e. 0.9 g of biliary
small intestine. As described above, bile facilitates the digestion cholesterol plus 0.7 g TICE cholesterol plus 0.4 g of dietary
and absorption of fats, in addition to promoting the elimination cholesterol). The other half (1.0 g) is lost in the feces. This,
of endogenous cholesterol. Hepatic hypersecretion of choles- combined with a loss of 0.4 g per day of cholesterol in the form
terol into bile and gallbladder hypomotility, among other genetic of fecal bile salts, yields a total cholesterol loss from the body
and environmental factors, promote the development of choles- of 1.4 g per day. Taking into account intestinal absorption of
terol gallstone disease (cholelithiasis) [67]. dietary cholesterol and reabsorption of biliary cholesterol, total
Although biliary secretion into the intestinal lumen was clas- body cholesterol synthesis is 1.0 g (i.e. cholesterol synthesis =
sically considered as the only significant route for cholesterol fecal loss of cholesterol + bile salts − dietary cholesterol intake)
elimination, more recent evidence has revealed that transintesti- [74]. This amounts to more than double what is consumed in the
nal cholesterol excretion (TICE) accounts for up to 35% [68, 69]. average diet.
TICE reflects the net effect of four cholesterol fluxes within the
enterocyte: cholesterol absorption from the intestine by the
action of NPC1L1 on the apical plasma membrane, cholesterol
excretion into the intestinal lumen mainly via ABCG5/8, also on PHARMACOLOGIC CLASSES
the apical membrane, the uptake of plasma lipoproteins at the AND AGENTS
basolateral membrane, and the basolateral excretion of choles-
terol in chylomicrons into the lymphatics [70]. The decision to treat dyslipidemia is largely dependent upon the
calculated cardiovascular risk. A number of clinical algorithms
exist for determining initiation of therapy. Goals for lipid lower-
Cholesterol balance ing were established in the 2001 National Cholesterol Education
Because cholesterol is converted by the liver to bile salts and is Program Adult Treatment Panel III (ATP III) guidelines [77],
secreted unmodified into bile, overall cholesterol balance which were updated in 2004 based upon the results of several
depends upon the disposition of both types of molecules. Rather additional large, randomized clinical trials [78]. These guidelines
than being lost in the feces after participating in cholesterol provide target LDL levels based on 10‐year risk of death from
transport and fat digestion, most bile salt molecules are recy- cardiovascular disease. More recent guidelines for the manage-
cled when they are taken up by high‐affinity transport proteins ment of dyslipidemia from the American Heart Association are
in the distal ileum. Bile salts enter the portal circulation and are no longer based on target LDL levels. Importantly, the guidelines
transported back to the liver, where they are cleared by hepato- stress that the first intervention should be therapeutic lifestyle
cytes from the blood with high first‐pass efficiency. Bile salts changes, including reduction of dietary saturated fat and choles-
are then re‐secreted into bile. This process of recycling bile salts terol intake, weight reduction, increased physical activity, and
between the liver and intestine is referred to as the enterohepatic possibly stress reduction.
circulation [71]. Whereas successful dietary therapy can reduce total choles-
The enterohepatic circulation is highly efficient, allowing terol by 5–25%, depending upon adherence and the metabolic
<5% of secreted bile salts to be lost in the feces. However, basis for elevated cholesterol concentrations. If this approach is
because bile salts are secreted in such large amounts, the small unsuccessful or insufficient to normalize lipid levels, drug
264 THE LIVER:  PHARMACOLOGIC CLASSES AND AGENTS

therapy is generally recommended. Eight classes of agents, displace cholesterol from micelles, increasing the loss of
including pharmacologic drugs and dietary supplements, are cholesterol from the feces [82]. The plant sterols and stanols
available for modification of lipid metabolism, and more are are themselves poorly absorbed. Based upon their mechanism
under study. Three of these classes (i.e. inhibitors of cholesterol of action, gram quantities of plant sterols and stanols are
synthesis, bile salt sequestrants, and cholesterol absorption required to reduce plasma LDL cholesterol concentrations by
inhibitors) have relatively well‐defined effects on lipid metabo- approximately 15%. Because an average diet contains 200–
lism. Whereas the overall effects of the other five classes are 400 mg of plant sterols and stanols, the molecules must be
clear, their molecular mechanisms of action are diverse and still enriched in dietary supplements (approximately 2 g) to be
subjects of active investigation. The following is a condensed effective [83].
summary of the mechanisms of these agents. Ezetimibe at very low concentrations reduces intestinal choles-
terol absorption by about 50%, without reducing the absorption
of triglycerides or fat‐soluble vitamins. It decreases cholesterol
Inhibitors of cholesterol synthesis movement from micelles into the enterocyte by inhibiting uptake
Cholesterol synthesis inhibitors, commonly known as statins, through the brush border protein NPC1L1 [84] (Figure  22.1b).
competitively inhibit the activity of HMG‐CoA reductase, the TICE is markedly increased by inhibition of NPC1L1 using
rate‐limiting enzyme in cholesterol synthesis [30]. This acti- ezetimibe [69], suggesting that modulation of TICE by other
vates a cellular signaling cascade culminating in the activation mechanisms could provide an additional strategy for reducing the
of SREBP‐2, which is a transcription factor that upregulates risk of atherosclerotic cardiovascular disease [68].
expression of the gene encoding the LDL receptor. Increased A reduction in cholesterol absorption that is achieved by
LDL receptor expression causes increased uptake of plasma either plant sterols and stanols or ezetimibe would be expected
LDL and, consequently, decreases plasma LDL cholesterol con- to decrease the cholesterol content of chylomicrons and there-
centrations [79]. fore the movement of cholesterol from the intestine to the liver.
This is because chylomicrons retain the intestinal cholesterol
that was originally incorporated into the particles after they are
Inhibitors of bile salt absorption metabolized to remnants by the liver. Within the liver, choles-
The bile salt sequestrants are cationic polymers that bind non- terol derived from chylomicron remnants contributes to the
covalently to negatively charged bile salt molecules in the small cholesterol that is packaged into VLDL particles. Therefore,
intestine. The resin–bile salt complex cannot be reabsorbed in inhibiting cholesterol absorption reduces cholesterol incorpo-
the distal ileum and is excreted into the stool. Decreased bile ration into VLDL particles, and decreases LDL cholesterol
salt reabsorption by the ileum partially interrupts enterohepatic concentrations in the plasma. Reduced hepatic cholesterol con-
bile salt circulation, causing hepatocytes to upregulate CYP7A1, tents lead to upregulation of the LDL receptor, which also con-
the rate‐limiting enzyme in bile salt synthesis from cholesterol tributes to the mechanism of LDL lowering by cholesterol
[80]. The increase in bile salt synthesis decreases hepatocyte absorption inhibitors [80].
cholesterol concentration, leading to increased expression of
the LDL receptor and enhanced LDL clearance from the circu-
lation [30]. The effectiveness of bile salt sequestrants in clear-
Fibrates
ing LDL from plasma is partially offset by concurrent Fibrates bind and activate peroxisome proliferator‐activated
upregulation of hepatic cholesterol and triglyceride synthesis, receptor α (PPARα), a nuclear receptor expressed in hepato-
which stimulates the production of VLDL particles by the liver cytes, skeletal muscle, macrophages, and the heart [85, 86].
[80]. As a result, bile salt sequestrants may also raise triglycer- Upon fibrate binding, PPARα heterodimerizes with the retinoid
ide levels and should be used with caution in patients with X receptor (RXR). This heterodimer binds to peroxisome
hypertriglyceridemia. proliferator response elements (PPREs) present in the pro-
­
moter region of specific genes and activates transcription of the
target genes [85].
Inhibitors of cholesterol absorption Activation of PPARα by fibrates results in numerous changes
Cholesterol absorption inhibitors reduce cholesterol absorption in lipid metabolism that act to decrease plasma triglyceride
by the small intestine. This includes dietary cholesterol, but ­levels and increase plasma HDL [87]. The decrease in plasma
more important they reduce the reabsorption of biliary choles- triglyceride levels is caused by increased muscle cell expression
terol, which comprises the majority of intestinal cholesterol of LPL, decreased hepatic expression of apoC‐III and increased
[74]. Like statins and bile salt‐binding resins, cholesterol hepatic oxidation of fatty acids. The increased expression of
absorption inhibitors reduce LDL cholesterol by increasing LPL in muscle results in increased uptake of triglyceride‐rich
clearance via the LDL receptor. In doing so, they enhance the lipoproteins, with a resultant decrease in plasma triglyceride
catabolism of VLDL and LDL apoB100, without affecting the levels. Because apoC‐III normally functions to inhibit interac-
production rate of these lipoproteins [80]. tion of triglyceride‐rich lipoproteins with their receptors, the
There are two available cholesterol absorption inhibitors: decrease in hepatic production of apoC‐III may potentiate the
plant sterols and ezetimibe. Plant sterols and stanols are increased LPL activity [23].
­naturally present in vegetables and fruits. They are similar in The mechanisms by which fibrates raise plasma HDL levels
molecular structure to cholesterol, but are substantially more are complex [88]. Whereas PPARα decreases hepatocyte
hydrophobic [81]. As a result, plant sterols and stanols ­production of apoA‐I in the mouse, the opposite appears be the
22:  Lipoprotein Metabolism and Cholesterol Balance 265

case in humans. This contributes directly to increased plasma gain‐of‐function mutations produce increase levels of LDL cho-
HDL. Upregulation of SR‐BI and ABCA1 in macrophages pre- lesterol in the plasma [93]. Monoclonal antibodies, genetic
sumably promotes cholesterol efflux from these cells in vivo. knockdown, and gene repair techniques have been used to inhibit
Hepatocytes also reduce expression of SR‐BI in response to PCSK9 [94]. The commercially available monoclonal antibodies
PPARα, which provides another mechanism for increased HDL alirocumab and evolocumab reduce LDL cholesterol levels from
levels in the plasma. 30 to 70%, even in severe familial hypercholesterolemia and sta-
Fibrates also lower LDL levels modestly. The lower LDL lev- tin‐resistant patients [95], and may produce other pleiotropic
els result from a PPARα‐mediated shift in hepatocyte metabo- anti‐atherogenic effects independent of LDL lowering [96].
lism toward fatty acid oxidation. PPARα increases the expression
of numerous enzymes involved in fatty acid transport and oxida-
tion [85]. This increases fatty acid catabolism, leading to
Inhibitors of VLDL secretion
decreased triglyceride synthesis and VLDL production. PPARα Therapies against dyslipidemia include the reduction of LDL
activation also results in LDL particles of larger size, which cholesterol and triglyceride levels using agents that reduce
appear to be taken up more efficiently by LDL receptor. Many VLDL secretion from the liver into the plasma. Mipomersen is
of these effects of PPARα on lipid metabolism are still the sub- an antisense oligonucleotide to apoB, whereas lomitapide is an
ject of basic and clinical investigation to develop of more selec- MTP inhibitor [97]. Both are effective at lowering plasma LDL
tive PPARα agonists that are capable of targeting selective concentration, even in patients with homozygous familial
aspects of lipid metabolism [89]. hypercholesterolemia, who lack both alleles of the LDL
­receptor. However, a mechanism‐based side‐effect is hepatic
steatosis, and steatorrhea in the case of lomitapide.
Niacin
Niacin (nicotinic acid, vitamin B3) is a water‐soluble vitamin. Agents in development
At physiologic concentrations, it is a substrate in the synthesis
of nicotinamide adenine dinucleotide (NAD) and nicotinamide Bempedoic acid, currently in phase 3 clinical trials, is an inhib-
adenine dinucleotide phosphate (NADP), which are important itor of adenosine triphosphate citrate lyase (ACLY), a cytosolic
cofactors in intermediary metabolism. enzyme upstream of HMG‐CoA reductase, which catalyzes the
The pharmacologic use of niacin necessitates large doses reaction that produces acetyl‐CoA from citrate, inhibiting the
(1500–3000 mg day−1) and is independent of the conversion of ­synthesis of cholesterol and ultimately leading to upregulation
nicotinic acid to NAD or NADP. Niacin increases HDL cho- of LDL receptor [98]. Other agents that increase LPL activity
lesterol and decreases plasma LDL cholesterol and triglycer- and thereby reduce plasma triglycerides concentrations are
ide concentrations, as well as Lp(a) levels [90]. The effects of currently under study, including antisense oligonucleotides to
niacin are partially mediated by a G protein‐coupled receptor acpoC-III, and gemcabene, which is a small molecule that
on adipocytes, which decreases the activity of adipose tissue reduces apoC‐III expression, also inhibits lipid synthesis, as
hormone‐sensitive lipase. Decreased hormone‐sensitive lipase well as enhancing VLDL clearance [97, 99]. Other approaches
reduces peripheral tissue triglyceride catabolism, and there- to increasing LPL activity are antisense oligonucleotides and
fore decreases the flux of free fatty acids to the liver. This, monoclonal antibodies against angiopoietin‐like protein 3,
together with local effects of niacin in liver, decreases the rate which is secreted by the liver and inhibits LPL [100].
of hepatic triglyceride synthesis and VLDL production. Niacin
also decreases the fractional catabolic rate of apoA‐I [90]. The
increased plasma apoA‐I increases plasma HDL concentra-
tions and presumably augments reverse cholesterol transport. CONCLUSIONS
The efficacy of available lipid‐lowering drugs that reduce LDL
Omega‐3 fatty acids represents an important advance in reducing cardiovascular
disease mortality. Future advances will build upon a rich
­
The omega‐3 fatty acids eicosopantaenoic acid (EPA) and doco-
­knowledge base of lipoprotein metabolism in order to identify
sahexaenoic acid (DHA), also referred to as fish oil, are effec-
new biochemical targets for the management of plasma lipids,
tive at reducing plasma triglycerides [91]. Although the
as well as related diseases such as type 2 diabetes and non‐­
molecular mechanisms are incompletely understood, the effects
alcoholic fatty liver disease.
are reduced hepatic triglyceride biosynthesis and increased fatty
acid oxidation in the liver [92].

Proprotein convertase subtilisin/kexin type 9 ACKNOWLEDGMENTS


inhibitors This work was supported in part by research grants DK056626,
PCSK9, which increases LDL cholesterol concentrations by pro- DK048873, and DK103046 from the National Institutes of
moting the degradation of the LDL receptor, was mapped as the Health (US Public Health Service) to DEC. MAA is the
third locus associated with autosomal dominant hypercholes- ­recipient of a 2017 NASH Fatty Liver Postdoctoral Research
terolemia together with LDL receptor and apoB, suggesting that Fellowship Award from the American Liver Foundation.
266 THE LIVER:  REFERENCES

REFERENCES 27. Foley, E.M., Gordts, P., Stanford, K.I. et  al. Hepatic remnant lipoprotein
clearance by heparan sulfate proteoglycans and low‐density lipoprotein
receptors depend on dietary conditions in mice. Arterioscler Thromb Vasc
1. Hussain, M.M., Rava, P., Walsh, M., Rana, M., and Iqbal, J. Multiple
Biol, 2013;33:2065–74.
­functions of microsomal triglyceride transfer protein. Nutr Metab (Lond),
28. Packard, C.J. and Shepherd, J. Lipoprotein heterogeneity and apolipoprotein
2012;9:14.
B metabolism. Arterioscler Thromb Vasc Biol, 1997;17:3542–56.
2. Smith, H.C. RNA binding to APOBEC deaminases: not simply a substrate
29. Ramasamy, I. Recent advances in physiological lipoprotein metabolism.
for C to U editing. RNA Biol, 2017;14:1153–65.
Clin Chem Lab Med, 2014;52:1695–727.
3. Hussain, M.M. Intestinal lipid absorption and lipoprotein formation. Curr
30. Brown, M.S. and Goldstein, J.L. A receptor‐mediated pathway for choles-
Opin Lipidol, 2014;25:200–6.
terol homeostasis. Science, 1986;232:34–47.
4. Sirwi, A. and Hussain, M.M. Lipid transfer proteins in the assembly of apoB‐
31. Kwon, H.J., Abi‐Mosleh, L., Wang, M.L. et  al. Structure of N‐terminal
containing lipoproteins. J Lipid Res, 2018;59:1094–102.
domain of NPC1 reveals distinct subdomains for binding and transfer of cho-
5. Iqbal, J., Rudel, L.L., and Hussain, M.M. Microsomal triglyceride transfer
lesterol. Cell, 2009;137:1213–24.
protein enhances cellular cholesteryl esterification by relieving product
32. Infante, R.E. and Radhakrishnan, A. Continuous transport of a small fraction
­inhibition. J Biol Chem, 2008;283:19967–80.
of plasma membrane cholesterol to endoplasmic reticulum regulates total
6. Carey, M.C., Small, D.M., and Bliss, C.M. Lipid digestion and absorption.
cellular cholesterol. Elife, 2017;6.
Annu Rev Physiol, 1983;45:651–77.
33. Horton, J.D., Goldstein, J.L., and Brown, M.S. SREBPs: activators of the
7. Wang, D.Q. and Cohen, D.E. Absorption and excretion of cholesterol and
complete program of cholesterol and fatty acid synthesis in the liver. J Clin
other sterols, in Clinical Lipidology: Companion to Braunwald’s Heart
Invest, 2002;109:1125–31.
Disease (ed. C. Ballantyne), Elsevier, Philadelphia, PA, 2014, pp. 25–42.
34. Wang, Y., Huang, Y., Hobbs, H.H., and Cohen, J.C. Molecular characteriza-
8. Yu, X.H., Qian, K., Jiang, N. et al. ABCG5/ABCG8 in cholesterol excretion
tion of proprotein convertase subtilisin/kexin type 9‐mediated degradation of
and atherosclerosis. Clin Chim Acta, 2014;428:82–8.
the LDLR. J Lipid Res, 2012;53:1932–43.
9. Hesse, D., Jaschke, A., Chung, B., and Schurmann, A. Trans‐Golgi proteins
35. Brown, M.S. and Goldstein, J.L. Plasma lipoproteins: teaching old dogmas
participate in the control of lipid droplet and chylomicron formation. Biosci
new tricks. Nature, 1987;330:113–14.
Rep, 2013;33:1–9.
36. Banach, M. Lipoprotein (a)‐We know so much yet still have much to learn.
10. Xu, X., So, J.S., Park, J.G., and Lee, A.H. Transcriptional control of
J Am Heart Assoc, 2016;5.
hepatic  lipid metabolism by SREBP and ChREBP. Semin Liver Dis,
37. Emdin, C.A., Khera, A.V., Natarajan, P. et al. Phenotypic characterization of
2013;33:301–11.
genetically lowered human lipoprotein(a) levels. J Am Coll Cardiol,
11. Wang, Y., Viscarra, J., Kim, S.J., and Sul, H.S. Transcriptional regulation of
2016;68:2761–72.
hepatic lipogenesis. Nat Rev Mol Cell Biol, 2015;16:678–89.
38. Boullier, A., Bird, D.A., Chang, M.K. et  al. Scavenger receptors, oxidized
12. Fisher, E.A. The degradation of apolipoprotein B100: multiple opportunities
LDL, and atherosclerosis. Ann N Y Acad Sci, 2001;947:214–22; discussion
to regulate VLDL triglyceride production by different proteolytic pathways.
22–3.
Biochim Biophys Acta, 2012;1821:778–81.
39. Nordestgaard, B.G., Chapman, M.J., Ray, K. et  al. Lipoprotein(a) as a
13. Christoffersen, M. and Tybjaerg‐Hansen, A. Novel genes in LDL metabo-
­cardiovascular risk factor: current status. Eur Heart J, 2010;31:2844–53.
lism – a comprehensive overview. Curr Opin Lipidol, 2015;26:179–87.
40. Santos‐Gallego, C.G., Badimon, J.J., and Rosenson, R.S. Beginning to
14. Strong, A., Patel, K., and Rader, D.J. Sortilin and lipoprotein metabolism:
understand high‐density lipoproteins. Endocrinol Metab Clin North Am,
making sense out of complexity. Curr Opin Lipidol, 2014;25:350–7.
2014;43:913–47.
15. Holmen, O.L., Zhang, H., Fan, Y. et al. Systematic evaluation of coding
41. Rye, K.A., Bursill, C.A., Lambert, G., Tabet, F., and Barter, P.J. The metabo-
­variation identifies a candidate causal variant in TM6SF2 influencing
lism and anti‐atherogenic properties of HDL. J Lipid Res, 2009;50(Suppl):
total cholesterol and myocardial infarction risk. Nat Genet, 2014;46:
S195–200.
345–51.
42. Tabet, F. and Rye, K.A. High‐density lipoproteins, inflammation and oxida-
16. Kozlitina, J., Smagris, E., Stender, S. et al. Exome‐wide association study
tive stress. Clin Sci (Lond), 2009;116:87–98.
identifies a TM6SF2 variant that confers susceptibility to nonalcoholic fatty
43. Kardassis, D., Mosialou, I., Kanaki, M., Tiniakou, I., and Thymiakou, E.
liver disease. Nat Genet, 2014;46:352–6.
Metabolism of HDL and its regulation. Curr Med Chem, 2014;21:2864–80.
17. Smagris, E., Gilyard, S., BasuRay, S., Cohen, J.C., and Hobbs, H.H.
44. Rosenson, R.S., Brewer, H.B., Jr., Chapman, M.J. et al. HDL measures, par-
Inactivation of Tm6sf2, a gene defective in fatty liver disease, impairs lipida-
ticle heterogeneity, proposed nomenclature, and relation to atherosclerotic
tion but not cecretion of very low density lipoproteins. J Biol Chem,
cardiovascular events. Clin Chem, 2011;57:392–410.
2016;291:10659–76.
45. Zannis, V.I., Fotakis, P., Koukos, G. et al. HDL biogenesis, remodeling, and
18. Pirazzi, C., Adiels, M., Burza, M.A. et  al. Patatin‐like phospholipase
catabolism. Handb Exp Pharmacol, 2015;224:53–111.
domain‐containing 3 (PNPLA3) I148M (rs738409) affects hepatic VLDL
46. Gillard, B.K., Bassett, G.R., Gotto, A.M., Jr, Rosales, C., and Pownall, H.J.
secretion in humans and in vitro. J Hepatol, 2012;57:1276–82.
Scavenger receptor B1 (SR‐B1) profoundly excludes high density lipopro-
19. Wolska, A., Dunbar, R.L., Freeman, L.A. et  al. Apolipoprotein C‐II: New
tein (HDL) apolipoprotein AII as it nibbles HDL‐cholesteryl ester. J Biol
findings related to genetics, biochemistry, and role in triglyceride metabo-
Chem, 2017;292:8864–73.
lism. Atherosclerosis, 2017;267:49–60.
47. Tabet, F., Remaley, A.T., Segaliny, A.I. et  al. The 5A apolipoprotein A‐I
20. Goldberg, I.J. 2017 George Lyman Duff Memorial Lecture: Fat in the blood,
mimetic peptide displays antiinflammatory and antioxidant properties in
fat in the artery, fat in the heart: triglyceride in physiology and disease.
vivo and in vitro. Arterioscler Thromb Vasc Biol, 2010;30:246–52.
Arterioscler Thromb Vasc Biol, 2018;38:700–6.
48. Gao, X., Yuan, S., Jayaraman, S., and Gursky, O. Role of apolipoprotein A‐II
21. He, P.P., Jiang, T., OuYang, X.P. et al. Lipoprotein lipase: biosynthesis, regu-
in the structure and remodeling of human high‐density lipoprotein (HDL):
latory factors, and its role in atherosclerosis and other diseases. Clin Chim
protein conformational ensemble on HDL. Biochemistry, 2012;51:4633–41.
Acta, 2018;480:126–37.
49. Miles, R.R., Perry, W., Haas, J.V. et al. Genome‐wide screen for modulation
22. Fong, L.G., Young, S.G., Beigneux, A.P. et al. GPIHBP1 and plasma triglyc-
of hepatic apolipoprotein A‐I (ApoA‐I) secretion. J Biol Chem, 2013;288:
eride metabolism. Trends Endocrinol Metab, 2016;27:455–69.
6386–96.
23. Rocha, N.A., East, C., Zhang, J., and McCullough, P.A. ApoCIII as a cardio-
50. Gursky, O. Structural stability and functional remodeling of high‐density
vascular risk factor and modulation by the novel lipid‐lowering agent volane-
lipoproteins. FEBS Lett, 2015;589:2627–39.
sorsen. Curr Atheroscler Rep, 2017;19:62.
51. Segrest, J.P., Jones, M.K., Catte, A. et al. Surface density‐induced pleating of
24. Cooper, A.D. Hepatic uptake of chylomicron remnants. J Lipid Res,
a lipid monolayer drives nascent high‐density lipoprotein assembly.
1997;38:2173–92.
Structure, 2015;23:1214–26.
25. Phillips, M.C. Apolipoprotein E isoforms and lipoprotein metabolism.
52. Wang, S. and Smith, J.D. ABCA1 and nascent HDL biogenesis. Biofactors,
IUBMB Life, 2014;66:616–23.
2014;40:547–54.
26. Mahley, R.W. and Ji, Z.S. Remnant lipoprotein metabolism: key pathways
53. Ossoli, A., Simonelli, S., Vitali, C., Franceschini, G., and Calabresi, L. Role
involving cell‐surface heparan sulfate proteoglycans and apolipoprotein E.
of LCAT in atherosclerosis. J Atheroscler Thromb, 2016;23:119–27.
J Lipid Res, 1999;40:1–16.
22:  Lipoprotein Metabolism and Cholesterol Balance 267

54. Jiang, X.C. Phospholipid transfer protein: its impact on lipoprotein homeo- 80. Couture, P. and Lamarche, B. Ezetimibe and bile acid sequestrants: impact
stasis and atherosclerosis. J Lipid Res, 2018;59:764–71. on lipoprotein metabolism and beyond. Curr Opin Lipidol, 2013;24:
55. Zhang, M., Zhai, X., Li, J. et al. Structural basis of the lipid transfer mecha- 227–32.
nism of phospholipid transfer protein (PLTP). Biochim Biophys Acta, 81. Armstrong, M.J. and Carey, M.C. Thermodynamic and molecular determi-
2018;1863:1082–94. nants of sterol solubilities in bile salt micelles. J Lipid Res, 1987;28:1144–55.
56. Phillips, M.C. Molecular mechanisms of cellular cholesterol efflux. J Biol 82. Taha, D.A., Wasan, E.K., Wasan, K.M., and Gershkovich, P. Lipid‐lowering
Chem, 2014;289:24020–9. activity of natural and semi‐synthetic sterols and stanols. J Pharm Pharm
57. Acton, S., Rigotti, A., Landschulz, K.T. et al. Identification of scavenger recep- Sci, 2015;18:344–67.
tor SR‐BI as a high density lipoprotein receptor. Science, 1996;271:518–20. 83. Gylling, H., Plat, J., Turley, S. et al. Plant sterols and plant stanols in the
58. Shen, W.J., Hu, J., Hu, Z., Kraemer, F.B., and Azhar, S. Scavenger receptor management of dyslipidaemia and prevention of cardiovascular disease.
class B type I (SR‐BI): a versatile receptor with multiple functions and Atherosclerosis, 2014;232:346–60.
actions. Metabolism, 2014;63:875–86. 84. Pirillo, A., Catapano, A.L., and Norata, G.D. Niemann‐Pick C1‐like 1
59. Lund-Katz, S. and Phillips, M.C. High density lipoprotein structure-function (NPC1L1) Inhibition and cardiovascular diseases. Curr Med Chem,
and role in reverse cholesterol transport. Subcell Biochem, 2010;51:183–227. 2016;23:983–99.
60. Villard, E.F., El Khoury, P., Duchene, E. et  al. Elevated CETP activity 85. Rakhshandehroo, M., Knoch, B., Muller, M., and Kersten, S. Peroxisome
improves plasma cholesterol efflux capacity from human macrophages in proliferator‐activated receptor alpha target genes. PPAR Res, 2010;2010.
women. Arterioscler Thromb Vasc Biol, 2012;32:2341–9. 86. Tyagi, S., Gupta, P., Saini, A.S., Kaushal, C., and Sharma, S. The peroxisome
61. Hogarth, C.A., Roy, A., and Ebert, D.L. Genomic evidence for the absence ­proliferator-activated receptor: A family of nuclear receptors role in various
of a functional cholesteryl ester transfer protein gene in mice and rats. Comp diseases. J Adv Pharm Technol Res, 2011;2(4):236–40.
Biochem Physiol B Biochem Mol Biol, 2003;135:219–29. 87. Shah, A., Rader, D.J., and Millar, J.S. The effect of PPAR‐alpha agonism on
62. Chiang, J.Y. Bile acid metabolism and signaling. Compr Physiol, 2013;3: apolipoprotein metabolism in humans. Atherosclerosis, 2010;210:35–40.
1191–212. 88. Botta, M., Audano, M., Sahebkar, A. et al. PPAR agonists and metabolic
63. Chiang, J.Y. Recent advances in understanding bile acid homeostasis. syndrome: an established role? Int J Mol Sci, 2018;19.
F1000Res, 2017;6:2029. 89. Ferri, N., Corsini, A., Sirtori, C., and Ruscica, M. PPAR‐alpha agonists are
64. Wang, D.Q., Cohen, D.E., and Carey, M.C. Biliary lipids and cholesterol still on the rise: an update on clinical and experimental findings. Expert
gallstone disease. J Lipid Res, 2009;509Suppl):S406–11. Opin Investig Drugs, 2017;26:593–602.
65. Reshetnyak, V.I. Physiological and molecular biochemical mechanisms of 90. Song, W.L. and FitzGerald, G.A. Niacin, an old drug with a new twist.
bile formation. World J Gastroenterol, 2013;19:7341–60. J Lipid Res, 2013;54:2586–94.
66. Wang, H.H., Garruti, G., Liu, M., Portincasa, P., and Wang, D.Q. Cholesterol 91. Del Gobbo, L.C., Imamura, F., Aslibekyan, S. et al. Omega‐3 polyunsatu-
and lipoprotein metabolism and atherosclerosis: Recent advances in reverse rated fatty acid biomarkers and coronary heart disease: pooling project of
cholesterol transport. Ann Hepatol, 2017;16:s27–s42. 19 cohort studies. JAMA Intern Med, 2016;176:1155–66.
67. Lammert, F., Gurusamy, K., Ko, C.W. et al. Gallstones. Nat Rev Dis Primers, 92. Ooi, E.M., Ng, T.W., Watts, G.F., and Barrett, P.H. Dietary fatty acids and
2016;2:16024. lipoprotein metabolism: new insights and updates. Curr Opin Lipidol,
68. Cohen, D.E. Of TICE in men. Cell Metab, 2016;24:773–4. 2013;24:192–7.
69. Jakulj, L., van Dijk, T.H., de Boer, J.F. et  al. Transintestinal cholesterol 93. Abifadel, M., Varret, M., Rabes, J.P. et al. Mutations in PCSK9 cause auto-
­transport is active in mice and humans and controls ezetimibe‐induced fecal somal dominant hypercholesterolemia. Nat Genet, 2003;34:154–6.
neutral sterol excretion. Cell Metab, 2016;24:783–94. 94. Banach, M., Aronow, W.S., Serban, M.C. et al. Lipids, blood pressure and
70. Reeskamp, L.F., Meessen, E.C.E., and Groen, A.K. Transintestinal choles- kidney update 2015. Lipids Health Dis, 2015;14:167.
terol excretion in humans. Curr Opin Lipidol, 2018;29:10–17. 95. Moriarty, P.M., Jacobson, T.A., Bruckert, E. et al. Efficacy and safety of
71. Carey, M.C. and Duane, W.C. Enterohepatic circulation, in The Liver: alirocumab, a monoclonal antibody to PCSK9, in statin‐intolerant patients:
Biology and Pathobiology (eds. I.M. Arias et al.), Raven Press, New York, design and rationale of ODYSSEY ALTERNATIVE, a randomized phase 3
1994, pp. 719–67. trial. J Clin Lipidol, 2014;8:554–61.
72. Wahlstrom, A., Sayin, S.I., Marschall, H.U., and Backhed, F. Intestinal 96. Karagiannis, A.D., Liu, M., Toth, P.P. et al. Pleiotropic anti‐atherosclerotic
crosstalk between bile acids and microbiota and its impact on host metabo- effects of PCSK9 inhibitorsfrom molecular biology to clinical translation.
lism. Cell Metab, 2016;24:41–50. Curr Atheroscler Rep, 2018;20:20.
73. Cohen, D.E. Hepatocellular transport and secretion of biliary phospholipids. 97. Ajufo, E. and Rader, D.J. New therapeutic approaches for familial hyper-
Semin Liver Dis, 1996;16:191–200. cholesterolemia. Annu Rev Med, 2018;69:113–31.
74. Cohen, D.E. Balancing cholesterol synthesis and absorption in the gastroin- 98. Saeed, A. and Ballantyne, C.M. Bempedoic acid (ETC‐1002): a current
testinal tract. J Clin Lipidol, 2008;2:S1–3. review. Cardiol Clin, 2018;36:257–64.
75. Bosner, M.S., Lange, L.G., Stenson, W.F., Ostlund, R.E., Jr. Percent 99. Gaudet, D., Alexander, V.J., Baker, B.F. et al. Antisense inhibition of apoli-
­cholesterol absorption in normal women and men quantified with dual stable poprotein C‐III in patients with hypertriglyceridemia. N Engl J Med,
isotopic tracers and negative ion mass spectrometry. J Lipid Res, 1999;40: 2015;373:438–47.
302–8. 100. Graham, M.J., Lee, R.G., Brandt, T.A. et al. Cardiovascular and metabolic
76. Sehayek, E., Nath, C., Heinemann, T. et  al. U‐shape relationship between effects of ANGPTL3 antisense oligonucleotides. N Engl J Med, 2017;377:
change in dietary cholesterol absorption and plasma lipoprotein responsive- 222–32.
ness and evidence for extreme interindividual variation in dietary cholesterol 101. Scappa, M.C., Kanno, K., and Cohen, D.E. Lipoprotein metabolism, in
absorption in humans. J Lipid Res, 1998;39:2415–22. The Textbook of Hepatology: From Basic Science to Clinical Practice, 3rd
77. Expert Panel on Detection, Evaluation, and Treatment of High Blood edn (eds. J. Rod’es, J.P. Benhamou, M.A. Rizzetto, J. Reichen and A. Blei),
Cholesterol in Adults. Executive summary of the third report of the national Blackwell, Oxford, 2007.
cholesterol education program (NCEP) expert panel on detection, evalua- 102. Cohen, D.E. and Armstrong, E.J. Pharmacology of cholesterol and
tion, and treatment of high blood cholesterol in adults (Adult Treatment lipoprotein metabolism, in Principles of Pharmacology: The Patho­
­
Panel III). JAMA, 2001;285:2486–97. physiologic Basis of Drug Therapy, 2nd edn (eds. D.E. Golan, A.H.
78. Grundy, S.M., Cleeman, J.I., Merz, C.N. et al. Implications of recent clinical Tashjian, E.J. Armstrong et  al.), Lippincott, Williams and Wilkins,
trials for the national cholesterol education program Adult Treatment Philadelphia, 2007.
Panel III guidelines. J Am Coll Cardiol, 2004;44:720–32. 103. Jonas, A. Lipoprotein structure, in Biochemistry of Lipids, Lipoproteins
79. Young, S.G. and Fong, L.G. Lowering plasma cholesterol by raising LDL and  Membranes, 4th edn (eds. D.E. Vance and J.E. Vance), Elsevier,
receptors – revisited. N Engl J Med, 2012;366:1154–5. Amsterdam, 2002, pp. 483–504.
SECTION C:
TRANSPORTERS,
BILE ACIDS, AND
CHOLESTASIS
Bile Acid Metabolism
23 in Health and Disease:
An Update
Tiangang Li1 and John Y.L. Chiang2
1
Department of Pharmacology, Toxicology and Therapeutics, University of Kansas Medical Center, Kansas
City, KS, USA
2
Department of Integrative Medical Sciences, Northeast Ohio Medical University, Rootstown, OH, USA

INTRODUCTION chapter, we will summarize the basic knowledge of bile acid


chemistry and biology, new mechanistic understanding of the
One of the major functions of the liver is to generate bile. Bile regulation of bile acid homeostasis, human diseases related to
consists of ~95% water plus organic solutes, inorganic electro- altered bile acid metabolism, and the recent advances in bile
lytes, and proteins. Bile acids are synthesized from cholesterol acid‐based therapies.
in the hepatocytes and are the major organic solutes in bile. Bile
acid synthesis accounts for a major fraction of hepatic choles-
terol turnover, and hepatic secretion of bile acids is a major driv-
ing force to generate bile flow. Following conjugation with BILE ACID SYNTHESIS, FUNCTION,
glycine or taurine, bile acids are efficiently reabsorbed in small AND REGULATION
intestine and circulated back to the liver in a process called the
enterohepatic circulation of bile acids. Many physiological
Bile acid biosynthesis
functions of bile acids depend on their physiochemical proper-
ties as amphipathic detergent molecules, while other functions Conversion of cholesterol to bile acids is a multistep reaction
are related to their signaling properties. In bile, bile acids pre- that involves enzymes located in the endoplasmic reticulum,
vent cholesterol precipitation by forming mixed micelles with mitochondria, cytosol, and peroxisomes of hepatocytes [1]. As
phospholipids. In the small intestine, bile acids facilitate the shown in Figure  23.1a, the conversion of cholesterol to bile
digestion of dietary fat by the pancreatic lipase and colipase, acids (cholic acid as an example) involves hydroxylation,
which forms mixed micelles with fatty acids and monoacylg- ­oxidation of the 3‐hydroxy group with isomerization of the
lycerols. These compounds are absorbed into enterocytes, ­double bond from C5 to C4, stereo‐specific reduction of the
re‐esterified to form triglycerides and then assembled into double bond at C4, and the oxidative cleavage of the side‐chain
­chylomicrons that are transported via the lymphatic system and to convert the isooctane side‐chain of cholesterol to an isopen-
delivered into circulation. In liver and intestine, bile acids serve tanoic acid of C24 bile acid molecules. In mammals, most bile
as endogenous ligands for nuclear receptors and cell surface acids have a 5β‐hydrogen group and a cis‐configuration along
G  protein‐coupled bile acid receptors that regulate metabolic the plane of the steroid nucleus. The space‐filling model
homeostasis and immune responses. (Figure 23.1a, right) of cholesterol shows that the carbon skel-
Disrupted bile flow due to biliary obstruction, genetic defects eton of cholesterol (black) is in a planar structure with one
of bile acid transporters, or acquired autoimmune destruction of hydroxy group at 3β‐position and a double bond at C5–C6 posi-
bile ducts causes cholestasis in humans. Pharmacological tion. In cholic acid, the A and B rings are kinked along the
approaches targeting bile acid metabolism, transport, and sign- C5–C10 bond and all three α‐hydroxyl groups and a ‐COOH
aling have been developed to treat hepatobiliary diseases. In this group are positioned on one side of the steroid nucleus

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
272 THE LIVER:  BILE ACID SYNTHESIS, FUNCTION, AND REGULATION

(a) 21
22

20 23
18 26
12 24
17 25
19 16 27
11 13 D
C
15
2 1 9 14
A
10
B 8 Cholesterol
3β HO 3 5
7
OH Hydrophobic face
4 6

12α OH

COOH
Cholic acid

3α HO OH 7α
OH COOH
Hydrophilic face

(b)
Major bile acid species in humans and mice.

Bile acids C-3 C-7 C-12 C-6


Chenodeoxycholic acid α-OH α-OH H H
Cholic acid α-OH α-OH α-OH H
Deoxycholic acid α-OH H α-OH H
Lithocholic acid α-OH H H H
α-muricholic acid α-OH α-OH H β-OH
β-muricholic acid α-OH β-OH H β-OH
ω-muricholic acid α-OH β-OH H α-OH
Ursodeoxycholic acid α-OH β-OH H H

Figure 23.1  Bile acid structure. (a) Structures of cholesterol and cholic acid are illustrated as examples. Bile acids are amphipathic molecules.
The 3α‐, 7α‐, and 12α‐HO groups of cholic acid face to one side of the steroid rings to form a hydrophilic face and the carbon skeleton forms a
hydrophobic face on the other side. Saturation of the C5–6 double bond creates a kink at A/B rings. (b) Major bile acids in humans and rodents have
hydroxy groups specifically located at C‐3, C‐7, C‐12, and C‐6 positions.

opposing the other hydrophobic side, which renders bile acids conjugates and taurine conjugates have a pKa value of ~4.0 and
amphipathic molecules with physiological detergent proper- ~2.0, respectively. Conjugated bile acids are soluble, mem-
ties. Figure 23.1b shows major bile acids in humans and rodents brane‐impermeable, and resistant to precipitation at physiolog-
with one‐ to three‐hydroxy groups located stereo‐specifically at ical pH. The amide bond is resistant to cleavage by the
C3, C7, C12, and C6 positions. Chenodeoxycholic acid (CDCA, pancreatic carboxypeptidases, and conjugated bile acids are
3α,7α‐dihydroxy bile acid) and cholic acid (CA, 3α,7α,12α‐tri- highly stable in the intestine lumen.
hydroxy bile acid) are major primary bile acids in humans, In humans, the bile acid pool consists of two primary bile
while cholic acid and α‐muricholic acid (α‐MCA, 3α,6β,7α‐tri- acids, CA and CDCA, and secondary bile acids including DCA
hydroxy bile  acid) and β‐muricholic acid (β‐MCA, 3α,6β,7β‐ and a very low concentration of LCA. Primary bile acids are the
trihydroxy bile acid)  are primary bile acids in rodents. The direct end‐products of bile acid biosynthesis in hepatocytes,
7α‐hydroxy group in CA and CDCA are removed by gut bacte- while secondary bile acids are formed from primary bile acids
rial 7α‐dehydroxylase to form secondary bile acids, deoxy- in the small and large intestine by bacterial enzymes. The human
cholic acid (DCA, 3α,12α‐dihydroxy bile acid) and lithocholic bile acid pool contains about 80% CDCA and CA, about 20%
acids (LCA, 3α‐monohydroxy bile acid) respectively. The gut DCA, trace amounts of LCA and UDCA, and other secondary
microbes also convert α‐ and β‐muricholic acids to ω‐muricholic bile acids in negligible amounts. In humans, CDCA and CA
acid (ω‐MCA, 3α,6α,7β‐trihydroxy bile acid) by epimerization generally present in roughly equal amounts but this varies by
for fecal excretion. Ursodeoxycholic acid (UDCA, 3α,7β‐ individual. TCA and Tα‐MCA and Tβ‐MCA are the predomi-
dihydroxy bile acid) is a minor primary in rodents and secondary nant bile acids in mice.
bile acid in humans formed by isomerization of the 7α‐hydroxy
group in CDCA to 7β‐position. Conjugated bile acids are ion-
ized at physiological pH. After synthesis, the side‐chain of bile
Bile acid biosynthesis in hepatocytes
acids is efficiently linked to the amino acids glycine (G) or tau- In hepatocytes, two major bile acid biosynthesis pathways,
rine (T) to form N‐acyl‐amidates, a process that is referred to as namely the classic pathway and the alternative pathway, synthe-
“conjugation” or “amidation.” The side‐chain of the glycine size primary bile acids (Figure 23.2a). Under normal physiology,
23:  BILE ACID METABOLISM IN HEALTH AND DISEASE 273

bile acid synthesis is more active in the pericentral hepatocytes of the steroid nucleus. In the alternative pathway, hydroxylation
and low in the periportal hepatocytes. This is because higher of the C‐7 position is catalyzed by the oxysterol 7α‐hydroxylase
sinusoidal bile acid uptake into the periportal hepatocytes inhib- (CYP7B1). CYP7B1 is expressed in liver for bile acid synthesis,
its bile acid synthesis. The relative contributions of the classic steroidogenic tissues for steroid synthesis, and macrophages
pathway and the alternative pathway to hepatic bile acid produc- for oxysterol synthesis. In the brain, sterol 24‐hydroxylase
tion vary in different species. In humans, the classic bile acid (CYP46A1) converts cholesterol to 24‐hydroxycholesterol,
synthesis pathway is the predominant pathway and accounts for which is hydroxylated at the 7α‐position by liver sterol 7α‐
about 90% of total hepatic bile acid production, while the alter- hydroxylase (CYP39A1). In the liver, sterol 25‐hydroxylase
native pathway accounts for less than 10% of total hepatic bile (non‐CYP enzyme) can convert cholesterol to 25‐hydroxycho-
acid production. In rodents, the classic and alternative pathways lesterol, a major oxysterol in circulation. Oxysterols generated
contribute about equally to the synthesis of primary bile acids. in extrahepatic tissues can be transported to the liver to be con-
The classic bile acid synthesis pathway is initiated by the rate‐ verted to bile acids. The enzyme cascade involved in bile acid
limiting enzyme cholesterol 7α‐hydroxylase (CYP7A1), a synthesis from oxysterols has not been identified.
cytochrome p450 enzyme located in the endoplasmic reticulum.
CYP7A1 catalyzes the stereo‐specific hydroxylation at the The classic versus alternative bile acid synthesis pathways
C‐7α ­position of cholesterol to produce 7α‐hydroxycholesterol
The classic pathway is also called “the neutral pathway” because
[2]. 3β‐Hydroxy‐Δ5‐C27‐steroid dehydrogenase (HSD3B7) con-
the steroid nucleus modifications occur before the side‐chain
verts 7α‐hydroxycholesterol to 7α‐hydroxy‐4‐cholesten‐3‐one
oxidation and thus most of the intermediates in this pathway do
(C4), which has been used as a surrogate serum marker for the
not have a carboxylic acid group. In contrast, the alternative
rate of hepatic bile acid synthesis in humans [3, 4]. If the C‐12
pathway is referred to as “the acidic pathway” because bile acid
position on C4 is left unmodified, CDCA will be produced.
synthesis in this pathway is initiated with cholesterol side‐chain
Alternatively, microsomal sterol 12α‐hydroxylase (CYP8B1)
oxidation to a C27‐carboxylic acid group. It is believed that
hydroxylates the C‐12 position on C4, leading to the production
hepatocytes are the only cells that express the entire set of
of CA. Therefore, C4 is a common precursor of CA and CDCA,
enzymes required for de novo bile acid synthesis. The classic
and CYP8B1 activity is one of the factors that affect the
bile acid synthesis pathway is tightly regulated, whereas the
CA:CDCA ratio in the bile acid pool. In this series of enzymatic
alternative pathway is constitutively active and is not regulated
reactions catalyzed by Δ4–3‐oxosteroid‐5β‐reductase (aldo‐keto
by bile acid feedback. In adult liver, the classic bile acid synthe-
reductase, AKR1D1) and 3α‐hydroxysteroid dehydrogenase
sis pathway is the predominant pathway for bile acid synthesis,
(AKR1C4), the 3β‐hydroxyl group of the steroid nucleus is
whereas in the newborn infant the alternative pathway produces
converted to a 3α‐hydroxyl group, the Δ5,6 double bond on
­
bile acids before CYP7A1 is expressed at weaning. There is a
the steroid nucleus is saturated, and the A ring and B ring of the
misconception that the alternative pathway only produces
steroid nucleus assume a cis‐configuration. After steroid ring
CDCA. In mice deficient of Cyp7a1, Cyp8b1 expression is
modifications, the mitochondrial enzyme sterol 27‐hydroxylase
increased and Cyp7b1 in the alternative pathway is doubled to
(CYP27A1) catalyzes the steroid side‐chain hydroxylation and
produce a smaller bile acid pool, but gallbladder bile still con-
oxidation of C27 (or C26) of the sterol intermediates to produce
tains a substantial amount of TCA (~30% compared to 45%
3α,7α,12α‐trihydroxycholestanoic acid (THCA), leading to cholic
TCA in wild‐type mice) [8]. Mutations of the CYP7A1 gene in
acid synthesis, and 3α,7α‐dihydroxycholestanic acid (DHCA),
a human patient only caused mild hypercholesterolemia and
leading to CDCA synthesis. Peroxisomal very‐long‐chain acyl‐
premature gallstone disease, indicating that the alternative bile
CoA synthase (VLACS, SLC27A5) then converts THCA and
acid synthesis pathway is activated to produce bile acids [9]. It
DHCA to acyl‐CoA thioesters, which are transported into
has been reported that in inborn errors of bile acid metabolism,
­peroxisomes via bile acid‐acyl transporter ABCD3 [5]. In the
C27‐acidic bile acid intermediates are accumulated, indicating
peroxisomes, α‐methylacyl‐CoA racemase (AMACR), enoyl‐
the alternative bile acid synthesis pathway may be stimulated.
CoA hydratase/3‐hydroxy acyl‐CoA dehydrogenase (EHHADH,
Therefore, bile acid synthesis can be switched from the classic
bifunctional enzyme) and acyl‐CoA oxidase 2 (ACOX2) catalyze a
pathway to the alternative pathways to synthesize both CA and
series of racemizations, hydrations, and dehydrogenations (β‐
CDCA. After synthesis, bile acids are conjugated to the amino
oxidation reaction), and then thiolase/sterol carrier protein X
acids glycine or taurine at C24‐COOH group immediately in
(SCPx) cleaves a propionyl‐CoA from THCA and DHCA to pro-
hepatocytes to form amidated bile acids (Figure 23.2b).
duce C24‐CoA thioesters, cholyl‐CoA, and chenodeoxycholyl‐
CA, respectively. Cholyl‐CoA and chenodeoxycholyl‐CoA
Species differences in bile acid synthesis
are subsequently conjugated to taurine or glycine by bile acid-
CoA:amino acid N‐acyltransferase (BAAT) [6, 7]. Unconjugated Bile acid synthesis pathways are remarkably similar between
bile acids returning to hepatocytes via p­ ortal circulation can be humans and other mammals. In mice and rats, the majority of
converted to their CoA thioesters by  microsomal bile acid- CDCA is converted to α‐MCA and β‐MCA (Figure 23.1b). Both
CoA synthase (BACS) and then r­econjugated to taurine or α‐MCA and β‐MCA have a hydroxyl group at the 6β‐position,
­glycine by BAAT. which increases solubility. A recent study reported that Cyp2c‐
The alternative pathway is initiated by mitochondrial cluster null mice lacked both α‐MCA and β‐MCA and had high
CYP27A1, which hydroxylates and oxidizes the cholesterol CDCA and UDCA content [10]. Further analysis showed that
side‐chain to produce 27‐hydroxycholesterol and then 3β‐ recombinant Cyp2c70 produced α‐MCA and β‐MCA using
hydroxy‐5‐cholestenoic acid prior to enzymatic modifications CDCA and UDCA, respectively, as a substrate (Figure 23.2c).
(a) (b)
Alternative pathway HO

Sterol 24-hydroxylase 24
12
(CYP46A1) (Brain) 12 C-NH-CH2-COO–
24
Classic pathway
O
3 7 Sterol 25-hydroxylase
HO 6
3 7
Cholesterol OH
Cholesterol 7α-hydroxylase Sterol 27-hydroxylase HO
6
(CYP7A1) (CYP27A1)
COOH Sterol 7α-hydroxylase Glyco-cholic acid
12 12
24 24 (CYP39A1)

Liver
3 7 3 7
Macrophage 24
HO OH HO 6
6
C-NH-CH2-CH2-SO3–
7α-hydroxycholesterol 3β-hydroxy-5-cholestanoic acid
Oxysterol 7α-hydroxylase O
3β-Hydroxy Δ5-C27 steroid
dehydrogenase (CYP7B1)
COOH
(HSD3B7) 12 12 Oxysterols 3 7
24 24 OH
HO
6
3 7
OH
3 7 Tauro-chenodeoxycholic acid
O 6 HO OH
6
7α-hydroxy-4-cholesten-3-one (C4) (c)
Sterol 12α-hydroxylase ? Hyocholic acid T/G-CA
T/G-CDCA
(CYP8B1) AKR1D1 (3α, 6α, 7α)
Aldo-keto reductases AKR1C4
Mitochondria
(AKR1D1, AKR1C4) 6α-hydroxylase Bile salt hydrolase
Sterol 27-hydroxylase (BSH)
(CYP27A1) α-MCA 6β-hydroxylase CDCA CA
3α, 7α, 12α-trihydroxycholestanoic acid 3α, 7α-dihydroxycholestanoic acid (3α, 6β, 7α)
(3α, 7α) (3α, 7α, 12α)
(THCA) (DHCA) (Cyp2c70)
Peroxisome Very long-chain acyl-CoA synthase CoA
(VLCAS), ABCD3, acyl-CoA oxidase, 7α/β-hydroxysteroid
CoA dehydrogenase/ 7α-dehydroxylase
Epimerase
bifunctional enzyme, thiolase/SCP2 Epimerase
HO Propionyl-CoA Cyp2c70
24 7β-hydroxylase DCA
24 β-MCA UDCA LCA
COOH (CoA) COOH (CoA) (3α, 12α)
12 (3α, 6β, 7β) (3α, 7β) (3α)
7β-dehydroxylase
3 7
3 7 Epimerase
HO OH 6α-hydroxylase 6β-hydroxylase
HO OH 6
6
Cholic acid (CA)(-CoA) Chenodeoxycholic acid (CDCA)(-CoA)
ω-MCA
Bile acid-CoA synthase (BACS) Hyodeoxycholic
(3α, 6α, 7β) Murideoxycholic
Bile acid-CoA amino N-acyltransferase acid (3α, 6α) acid (3α, 6β)
(BAAT)
T/G-CA T/G-CDCA

Figure 23.2  Bile acid synthesis pathways. (a) The classic and alternative bile acid synthesis pathways are shown. Enzymes involved in bile acid synthesis are described in the text. (b) Structure
of glycine‐ and taurine‐conjugated bile acids. Glycocholic acid and taurochenodeoxycholic acid are shown as examples. (c) Bile acid biotransformation by gut bacteria. Bacterial enzymes involved
in hydroxylation and dihydroxylation produce various secondary bile acids. Species differences in bile acid transformation are described in the text.
23:  BILE ACID METABOLISM IN HEALTH AND DISEASE 275

CDCA is isomerized to UDCA in liver and gut bacteria, which β‐MCA to ω‐MCA increased the solubility of MCA, leading to
may be the predominant substrate for Cyp2c70 to synthesize fecal excretion or circulating in portal blood. The fecal bile
β‐MCA in rodent liver. acids consist of predominantly unconjugated bile acids. Some
Another aspect of the species‐dependent difference in bile secondary bile acids are reabsorbed in the intestine and trans-
acid pool composition is the preferential use of glycine (G) or ported to the liver and bile by portal circulation. Secondary bile
taurine (T) for bile acid conjugation. The bile acid pool in many acids circulated to hepatocytes can be detoxified by sulfonation
mammalian species such as mice, rats, and dogs contains pre- for renal secretion, reconjugated to glycine or taurine, or glu-
dominantly taurine conjugates, while the bile acid pool in curonidated for secretion into bile.
humans, hamsters, and rabbits contains about two‐thirds or
more glycine conjugates, with the rest being taurine conjugates
[11, 12]. Taurine‐conjugated bile acids are less toxic than their
corresponding glycine‐conjugated and unconjugated bile acids.
From an evolutionary point of view, the use of glycine for bile ENTEROHEPATIC CIRCULATION
acid conjugation appears to be a more recent event. The spe- OF BILE ACIDS
cies‐dependent preference for using glycine or taurine for bile
acid conjugation is thought to be determined by BAAT substrate Overview of bile acid circulation
specificity and possibly the availability of taurine and glycine in and homeostasis
the peroxisomes [6, 13, 14].
Bile acids are actively transported across the hepatocyte cana-
licular membrane, which plays an important role in promoting
Bacterial modification of bile acids: synthesis water and other biliary lipid secretion and provides a major
osmotic force for bile salt‐dependent bile flow. Under normal
of secondary bile acids conditions, bile flows freely in intrahepatic bile ducts due to the
When bile is released into the small intestine, a portion of bile relatively low pressure in the biliary tree. Under fasting condi-
acids can further undergo complex modifications by bacterial tions, little bile enters the duodenum due to the high tone of the
enzymes in the ileum and large intestine [15]. Bacterial bile salt sphincter of Oddi, which therefore promotes bile entrance into the
hydrolases (BSH) deconjugate T/G‐conjugated bile acids to gallbladder via the cystic duct. In the gallbladder, water and elec-
form unconjugated bile acids (Figure 23.2c). In the colon, bac- trolytes are reabsorbed and bile is concentrated ~5‐ to 10‐fold.
terial 7α‐dehydroxylases convert CA to DCA, and CDCA to After food intake, dietary fat and amino acids stimulate the neu-
LCA [15–17]. DCA is the predominant bile acid in the colon. roendocrine cells of the duodenum to secrete a peptide hormone
Unconjugated primary and secondary bile acids are hydropho- called cholecystokinin (CCK), which stimulates gallbladder con-
bic and are passively absorbed in the ileum and large intestine. traction and sphincter of Oddi relaxation to release concentrated
DCA is delivered to the liver where it is reconjugated and joins bile into the duodenum [19]. CCK also stimulates pancreatic aci-
the circulating bile acids [18]. Bacterial 3α,7α,12α‐hydroxyster- nar cells to simultaneously secrete digestive juice into duodenum.
oid dehydrogenases (HSDHs) epimerize the α‐hydroxyl groups In the small intestine lumen, bile acids form mixed micelles with
of bile acids to carbonyl groups to form 3‐oxo, 7‐oxo‐, and 12‐ dietary lipids and cholesterol to facilitate digestion by pancreatic
oxo‐bile acids, respectively. Then the carbonyl groups in oxo‐ lipases and colipases. Bile acids are efficiently reabsorbed in the
bile acids are converted to β epimers: iso‐bile acids and epi‐bile ileum and transported back to the liver via portal circulation. In
acids by 3β‐, 7β‐, and 12β‐HSDHs. The 7α‐hydroxy‐bile acids humans, the bile acid pool circulates a few times a day between
have higher bactericidal activity than oxo‐bile acids and β‐ the liver and the intestine (Figure 23.3). Because bile acids are
epimers. In humans and mice, bacterial 7β‐HSDH epimerizes reabsorbed in the small intestine at ~95% efficiency, the bile acid
the 7α‐hydroxyl group of CDCA (3α,7α) to 7β, forming UDCA. pool is largely conserved. In humans, the bile acid pool contains
The 7β‐epimerization converts highly hydrophobic CDCA to about 2–4 g of bile acids. Roughly 0.2–0.6 g of bile acids is syn-
hydrophilic and nontoxic UDCA. thesized daily by the hepatocytes to replace the bile acids that
LCA is the most hydrophobic and toxic bile acid and is pre- are lost in the feces. The enterohepatic circula­tion of bile
sent in trace amounts in the human bile acid pool. In human acids is driven by a network of bile acid efflux and uptake trans-
hepatocytes, LCA is mainly detoxified by sulfonation at the C‐3 porters expressed in hepatocytes, cholangiocytes, and enterocytes
position by a family of sulfotransferases (SULT2A1, SULT2B8, briefly described here (Figure 23.3).
etc.). Sulfated LCA is poorly reabsorbed in the small intestine
and is excreted into feces. In mice, LCA is sulfated at the C‐7
Biliary secretion of bile acids
position but sulfonation is not a major detoxification pathway.
Mouse gut bacteria detoxify bile acids mainly by hydroxylation Hepatocytes are polarized cells. The apical membrane of two
of mono‐ and di‐hydroxyl bile acids to polyhydroxylated bile adjacent hepatocytes form a bile canaliculi surrounded by tight
acids for reabsorption or secretion. junctions. A network of bile canalicular collects bile generated
In pig, CDCA is 6α‐hydroxylated to hyocholic acid by hepatocytes and drains bile into small bile ducts that are
(3α,6α,7α) and LCA is converted to the hyodeoxycholic formed by biliary epithelial cholangiocytes at the portal triad.
acid (3α,6α). Only rodents can convert LCA back to UDCA by Cholangiocytes can further modify bile in the biliary tree by
7β‐hydroxylase and to murideoxycholic acid (3α,6β) by bacte- regulated secretion (e.g. water, HCO3−) and reabsorption (e.g. bile
rial 6α‐ and 6β‐hydroxylases (Figure 23.2c). Epimerization of acids, glucose, and other solutes). Bile acids, phospholipids, and
276 THE LIVER:  ENTEROHEPATIC CIRCULATION OF BILE ACIDS

ABCG5/G8
Cholesterol MDR3
Phospholipid

CYP7A1
Bile Bile

BSEP
MRP2

Bilirubin glucuronide
Bile acids Sulfated bile acids
OSTα/β
MRP3/4

NTCP OATPs
Systemic
circulation Cholehepatic shunt
ASBT
OSTα/β

Portal
blood
Hepatocyte
Bacteria-mediated
Bile acids biotransformation
Cholangiocyte of bile acids in gut
OSTα/β ASBT

Enterocyte

Figure 23.3  Enterohepatic circulation of bile acids. The major transporters involved in bile formation and enterohepatic circulation of bile acids
are illustrated in hepatocytes, cholangiocytes, and enterocytes.

cholesterol are the major organic solutes in bile. Once secreted small intestine where the heterodimer mediates the secretion of
by hepatocytes into bile, they form mixed micelles to increase cholesterol and other plant sterols back into the lumen. This pro-
cholesterol solubility. Genetic defects in either biliary bile acid cess is significant in that it prevents accumulation of toxic die-
secretion or phospholipid secretion are known to result in tary sitosterols in the blood and tissues, and it also limits
cholesterol monohydrate precipitation and gallstone disease intestinal cholesterol absorption efficiency at about 50%.
in pediatric patients. In addition, mixed micelle formation Multidrug‐resistant 3 (MDR3, ABCB4) mediates the canalicu-
decreases the biliary concentration of monohydroxy bile acids, lar secretion of phosphatidylcholine, which is the major phos-
which may damage the biliary tract upon chronic exposure at pholipid in bile [23]. Detailed description of bile acid
high concentrations. Canalicular bile acid secretion into the bile transporters can be found in Chapters 26 and 27.
against the concentration gradient is the rate‐limiting step in bile
formation [20]. The ATP‐binding cassette (ABC) transporter
bile salt export pump (BSEP, ABCB11) is the major canalicular
Intestinal reabsorption of bile acids
bile acid efflux transporter in hepatocytes [21]. Hydrolysis of Intestinal bile acids are reabsorbed via passive diffusion of
ATP provides the driving force for active transport of bile acids unconjugated bile acids and active transport of conjugated bile
against a high bile acid concentration gradient in bile. BSEP acids, with the latter being the quantitatively major mechanism
expression is hepatocyte‐specific, and BSEP has high substrate (Figure 23.3). Bile acid reabsorption mainly occurs at the termi-
specificity for conjugated bile acids. Multidrug resistance‐asso- nal ileum where apical sodium‐dependent bile salt transporter
ciated protein 2 (MRP2, ABCC2) mediates the secretion of (ASBT, SLC10A2) is highly expressed [24]. Humans and mice
bilirubin glucuronide, glutathione, and some sulfated and conju- lacking functional ASBT present with bile acid malabsorption
gated bile acids into bile. The ABCG5 and ABCG8 transporters [25, 26]. Once absorbed into enterocytes, bile acids bind to the
form a heterodimer to mediate free cholesterol secretion into bile acid‐binding protein (I‐BABP), which transports bile acids
bile [22]. Genetic variations of ABCG5 and ABCG8 are associ- to the basolateral membrane for secretion by the organic solute
ated with gallstone disease in humans because biliary choles- transporters OSTα (SLC51A) and OSTβ (SLC51B) [27]. OSTα
terol hypersecretion is the major risk factor for gallstone and OSTβ form a functional heterodimer to efflux bile acids
formation. ABCG5 and ABCG8 are also highly expressed in the across the basolateral membrane into the portal circulation [28].
23:  BILE ACID METABOLISM IN HEALTH AND DISEASE 277

Mice lacking a functional OSTα/OSTβ heterodimer had signifi- in the systemic blood is very low. It is uncertain if such low
cantly decreased intestinal bile acid absorption, lower plasma concentration of bile acids in the systemic blood can elicit sig-
bile acid concentration, and a small bile acid pool [29]. ASBT nificant physiological functions in peripheral organs that express
and OSTα/β are also expressed at the apical side and the baso- bile acid receptors. In addition, bile acid composition in the
lateral side, respectively, of cholangiocytes and renal proximal systemic blood may differ significantly from that of bile due to
tubular cells [28]. In the biliary tract, ASBT and OSTα/β func- the differential extraction rate of various bile acid species via
tion sequentially to transport a small flux of bile acids across the active transport or passive diffusion along the sinusoid. In chol-
biliary epithelium into the peribiliary plexus. From here, bile estasis, basolateral efflux of bile acids is increased and causes
acids are further transported to hepatic sinusoidal cells and are significantly elevated bile acid concentration in the systemic
reabsorbed by hepatocytes. circulation [28, 40, 41]. Several transporters including OSTα/β,
If unconjugated dihydroxy bile acids are secreted into the bil- MRP3, and MRP4 mediate basolateral efflux of bile acids from
iary ductules, they may be absorbed passively. This process is hepatocytes [18, 28, 42]. Plasma bile acids increased in choles-
called cholehepatic shunting, which generates a hypercholeresis tasis are sulfated and eliminated mainly via renal excretion.
effect by recycling biliary bile acids for hepatic apical resecre- Species differences in serum and bile acid concentration and
tion [30]. In the kidney proximal tubule cells, ASBT and OSTα/β composition have been reported [43].
coordinate the efflux of bile acids from kidney back to the
systemic circulation. Inhibition of ASBT is expected to increase
urinary bile acid excretion [31].
REGULATION OF BILE ACID SYNTHESIS,
TRANSPORT, AND HOMEOSTASIS
Hepatic uptake of bile acids
Hepatic basolateral uptake of bile acids is more active in the peri- Bile acid pool size is usually kept at a relatively constant level
portal hepatocytes rather than the pericentral hepatocytes. The because daily bile acid synthesis roughly equals the daily
first‐pass extraction rate is about 60–90%, depending on the bile amount of bile acids excreted in feces. Such a balance is main-
acid with little spillover of portal bile acids into the systemic cir- tained by bile acid sensing mechanisms that regulate hepatic
culation [32]. Na+‐taurocholate co‐transporting polypeptide bile acid synthesis and intestinal bile acid reabsorption.
(NTCP, SLC10A1) is the major basolateral conjugated When  bile acid concentration increases in the enterohepatic
bile acid uptake transporter in hepatocytes [33–35]. Coupling bile system, bile acids inhibit the transcription of genes in bile acid
acid transport to the Na+ gradient provides energy for active trans- synthesis and intestinal bile acid transport, leading to decreased
port of serum bile acids across the sinusoidal membrane to hepat- hepatic output and increased fecal loss. When bile acid con-
ocytes. The human isoforms of organic anion transporters centration decreases, these two pathways are upregulated. Bile
(OATP), OATP1A2, OATP1B1, and OATP1B3, mediate Na+‐ acid‐activated farnesoid X receptor (FXR) plays a key role in
independent basolateral transport of conjugated and unconju- mediating bile acid effects in the enterohepatic system. FXR
gated bile acids and other organic anions, such as steroids and belongs to the nuclear receptor superfamily, which consists of
drugs [18, 36]. So far there has been only one reported case of a group of ligand‐activated transcription factors [44]. Upon
human NTCP loss‐of‐function mutation that resulted in a hyper- ligand binding, FXR binds to its target gene promoters and
cholanemic phenotype with significantly elevated circulating bile activates gene transcription. Both conjugated and unconju-
acids of ~1500 μM (normally less than 10 μM) without clinical gated CDCA and CA are endogenous ligands for FXR, but CA
signs of cholestasis [37]. Experiments conducted in mice showed (EC50 = 586 μM) is a much weaker FXR agonist than CDCA
that the plasma clearance of taurocholate after intravenous admin- (EC50 = 17 μM). The hydrophilic bile acids T‐UDCA and T‐
istration was delayed but not abolished in mice lacking NTCP MCAs do not activate FXR but are FXR antagonists [45]. FXR
[38]. Only a small subset of the Ntcp knockout mice showed is expressed in hepatocytes and enterocytes that are routinely
marked elevation of plasma bile acid concentration, while the exposed to high concentrations of bile acids. The conjugated
majority of the Ntcp knockout mice had normal plasma bile acid and nonconjugated secondary bile acids LCA and DCA acti-
concentration [38]. A more recent study showed that pharmaco- vate membrane G protein‐coupled bile acid receptor 1
logical inhibition of NTCP in mice lacking OATP1A/1B achieved (Gpbar‐1) [46], aka Takeda G receptor 5 (TGR5) [47]. The
close to complete inhibition of plasma taurocholate clearance and roles of FXR and TGR5 in liver physiology are reviewed in
markedly elevated plasma bile acids [39]. These findings imply Chapters 24 and 25. The following sections will only focus on
that in mice both NTCP and OATPs mediate quantitative basolat- the role of bile acid receptors in the regulation of bile acid
eral bile acid uptake into hepatocytes. In humans, it is possible synthesis and transport.
that NTCP is the predominant basolateral bile acid uptake trans-
porter and its loss‐of‐function cannot be compensated to a Regulation of bile acid synthesis
­significant degree by the presence of OATPs.
The early evidence supporting the presence of bile acid feed-
back inhibition of bile acid synthesis came from experimental
Plasma bile acids observations that hepatic CYP7A1 enzyme activity markedly
The presence of bile acids in the systemic circulation is mainly decreased when rats were fed a diet containing bile acids, and
due to incomplete extraction of portal bile acids by the liver. that hepatic CYP7A1 enzyme activity was induced when intestine
Under normal physiological conditions, bile acid concentration bile acid absorption was disrupted by bile acid sequestrants.
278 THE LIVER:  REGULATION OF BILE ACID SYNTHESIS, TRANSPORT, AND HOMEOSTASIS

Hepatic CYP7A1 function is mainly controlled at the transcrip- injury, bile acids may activate FXR to induce nuclear receptor
tional level and evidence suggesting direct regulation of small heterodimer partner (SHP). SHP is an atypical nuclear
CYP7A1 protein stability or enzyme activity is scarce. The receptor without DNA binding activity and often acts as a
CYP7A1 gene promoter has been well characterized. It contains co‐repressor to inhibit other transcriptional factors via protein–
one or two bile acid response elements (BARE‐1 and BARE‐2). protein interactions. In this case, SHP inhibits the trans‐
Human, mouse, and rat CYP7A1 gene proximal promoters con- activating activity of HNF4α and LRH‐1, leading to the
tain a BARE‐1 that binds two nuclear receptors, hepatocyte inhibition of CYP7A1 gene transcription [48, 49]. However,
nuclear factor 4α (HNF4α) and liver‐related homolog 1 intrahepatic TCA and TCDCA concentrations may not be high
(LRH‐1). The putative endogenous ligands for HNF4α and enough to activate FXR under physiological conditions in mice
LRH‐1 are phospholipids, and they are believed to be constitu- and humans. It was reported in 1995 that intraduodenal infu-
tively active in hepatocytes. These two nuclear receptors are sion, but not intravenous infusion, of TCA inhibited CYP7A1
largely responsible for the basal expression of hepatic CYP7A1, expression in rats with bile fistula, suggesting that intestinal
because mutagenesis that abolishes their binding results in sig- factor(s) might be required to mediate bile acid feedback inhibi-
nificant reduction of CYP7A1 promoter activity. In addition, tion of bile acid synthesis [50]. The intestine is exposed to high
bile acids mainly inhibit bile acid synthesis by repressing concentrations of bile acids. Not only is trans‐enterocyte bile
HNF4α and LRH‐1 transactivation of CYP7A1 via BARE‐2. acid flux highly regulated, the intestine stores a significantly
Two major mechanisms are believed to mediate bile acid larger portion of the bile acid pool (~70–80%) compared to the
inhibition of CYP7A1 (and also CYP8B1) (Figure 23.4). One is gallbladder and the liver [51]. Bile acid activation of intestine
mediated by FXR in hepatocytes and the other is initiated by FXR induces the transcription of fibroblast growth factor 15
activation of FXR in the small intestine. When bile acid concen- (FGF15), which is subsequently released into the portal circula-
tration increases within hepatocytes, such as in cholestatic liver tion and acts as an endocrine signaling molecule to inhibit

Cholesterol

HNF4α
ERK CYP7A1
LRH-1

FGFR4 SHP

FXR FXR

Bile acids BSEP

NTCP

FGF15/19 Bile acids

Portal blood

OSTα/β
TCDCA
ASBT
IBABP
FXR

FGF15/19

Insulin
sensitivity TGR5 FXR
GLP-1

L cells

Figure 23.4  Mechanisms of bile acid feedback regulation of bile acid synthesis and homeostasis. Bile acid‐activated FXR regulates bile acid
synthesis and bile acid transport in hepatocytes and enterocytes. FXR and TGR5 are coexpressed in the intestinal L cells.
23:  BILE ACID METABOLISM IN HEALTH AND DISEASE 279

CYP7A1 gene transcription in hepatocytes [52]. FGF15 binds Regulation of bile acid transport
and activates membrane FGF receptor 4 (FGFR4), which is the
predominant FGF receptor expressed in hepatocytes. FGFR4 Biliary bile acid secretion and ileal bile acid uptake are two key
forms a complex with β‐Klotho to activate ERK1/2 to repress mechanisms driving the enterohepatic circulation of bile acids.
CYP7A1 gene expression. The downstream target of FGF15 In hepatocytes, FXR induces BSEP transcription to promote
signaling is still not fully clear, but FGF15‐activated ERK1/2 apical bile acid secretion [78]. On the basolateral side, FXR
may phosphorylate HNF4α to inhibit its binding to the CYP7A1 activation inhibits NTCP expression [79]. Similarly, in entero-
gene [53]. FGF15 is also required for gallbladder refilling [54], cytes, FXR activation induces intestinal bile acid binding pro-
and is an insulin‐independent postprandial regulator of hepatic tein (IBABP), OSTα, and OSTβ and inhibits ASBT [80–82].
protein and glucose metabolism [55]. FXR induces BSEP, I‐BABP, OSTα, and OSTβ gene transcrip-
FGF19 is the human ortholog of FGF15 and shares ~51% tion via direct binding to their gene promoters. FXR inhibition
amino acid sequence identity with mouse FGF15. FGF19 tran- of ASBT may be mediated by SHP. Therefore, FXR‐mediated
scription is also induced by FXR [53, 56]. Functionally, FGF19 regulation of bile acid transporter expression does not promote
activates FGFR4 and ERK1/2 and strongly represses CYP7A1 trans‐hepatocyte or trans‐enterocyte bile acid flux. Instead, the
gene transcription in human hepatocytes [53, 56]. Human pri- main purpose seems to be decreasing intracellular bile acid con-
mary hepatocytes express FGF19, while mouse hepatocytes do centration. This FXR‐regulated bile acid transport may serve as
not express FGF15 [53, 57]. Human patients with extrahepatic an adaptive response to protect against bile acid accumulation in
obstructive cholestasis had increased liver FGF19 mRNA and hepatocytes during cholestasis. Pharmacological inhibition of
plasma FGF19 concentration [57], while mouse intestine FGF15 NTCP has hepato‐protective effects in mouse models of choles-
expression significantly decreased in a bile duct ligation mouse tasis [83]. Isoforms of the basolateral transporters OSTα/β and
model of obstructive cholestasis [52]. These findings suggest an MRP 3/4 are induced during cholestasis to efflux bile acids into
autocrine model by which hepatocyte FGF19 inhibits CYP7A1 the systemic circulation [28, 40, 41].
in response to intrahepatic bile acid concentration.
The physiological role of FXR‐mediated regulation of hepatic
Gut microbiota and bile acid metabolism
bile acid synthesis is best demonstrated in genetically modified
mice with disrupted FXR, SHP, or FGF15 [58–62]. Another line High‐fat diets, circadian disruption, time of feeding/fasting,
of information came from mice with disrupted enterohepatic drugs, alcohol, and hormones all can shape the gut microbiota
bile acid transport. In mice lacking OSTα, hepatic CYP7A1 was to alter bile acid metabolism, homeostasis and host metabo-
increased, despite reduced bile acids returning to the liver and a lism [84]. Biotransformation of primary bile acids to second-
significantly smaller bile acid pool size [29, 63]. These observa- ary bile acids occurs in the small intestine and the colon. The
tions indicate that an enlarged bile acid pool in the enterocytes small intestine has a low bacterial population (a gradient of
can exert a dominant effect on inhibition of hepatic bile acid 103–108 g−1 wet weight in feces) but is the major site for bile
synthesis and the intestine‐to‐liver axis via FXR/FGF15 signal- acid reabsorption, and nutrient digestion and absorption. The
ing plays a critical role in bile acid feedback regulation. colon has a high bacterial population (1011 g−1 wet weight).
Several studies have shown that CYP7A1 mRNA was strongly DCA is a potent antimicrobial agent that controls bacterial
repressed in hepatocytes and in mice within 2–6 hours post overgrowth and maintains intestinal barrier function. BSH and
FGF15 or FGFG19 administration [58, 64, 65]. Such rapid 7α‐HSDH activities determine the extent of secondary bile
CYP7A1 downregulation is consistent with the reported short acid synthesis, the bile acid species present, and the hydropho-
CYP7A1 mRNA half‐life of about 30 minutes in cultured cells bicity of bile acids in enterohepatic circulation. Antibiotic‐
[66]. In a recent study, FXR activation was shown to induce treated mice and germ‐free mice have increased bile acid
RNA‐binding protein ZFP36L1 to decrease the stability of synthesis and pool size (~30%) compared to conventionally
CYP7A1 mRNA [67]. Liver CYP7A1 mRNA was markedly raised mice [45]. Thus, the gut microbiota determines the total
decreased in bile duct‐ligated mice [68, 69]. It is expected that bile acid pool size in the liver, gallbladder, bile, and intestine.
intestine FGF15 production is significantly decreased under Gut bacteria utilize bile acids, polysaccharides, cellulose, and
obstructive cholestasis. Both hepatic FXR activation and proin- starch to produce short‐chain fatty acids such as acetate,
flammatory cytokines can cause CYP7A1 repression in obstruc- butyrate, and propionate, and amino acids for energy metabo-
tive cholestasis [70]. These findings demonstrate that significant lism and growth. In the small intestine and colon, BSH activity
redundancy exists regarding bile acid feedback inhibition of bile is  high in the Gram‐positive bacteria genera Clostridium,
acid synthesis in both physiological and pathological conditions. Enterococcus, Bifidobacterium, and Lactobacillus, and Gram‐
Bile acid synthesis is the major pathway for cholesterol negative genus Bacteroides. These anaerobic bacteria contain
catabolism in hepatocytes. In mice and rats, but not in humans, high HSDH activity and are involved in multiple steps of 7α‐
cholesterol accumulation induces CYP7A1 via activation of the dehydroxylation during secondary bile acid synthesis by gut
nuclear receptor liver X receptor (LXR) [71, 72]. LXR in hepat- bacteria. The bile acid‐inducible operon (bai) in Clostridium
ocytes also induces the transcription of ABCG5/G8 and ABCA1 species has been elucidated [15]. The baiE gene encodes bile
and ABCG1, which mediate cholesterol efflux, alleviating intra- acid 7α‐HSDH, and the baiI gene may encode 7β‐HSDH.
hepatic cholesterol accumulation [73–75]. LXRα binds BARE‐1 Antibiotic‐resistant Clostridium difficile (C. dif) outbreaks in
in the mouse and rat Cyp7a1 gene promoter, but does not bind hospitals are a growing concern for infection and patient mor-
BARE‐1 in the human CYP7A1 gene promoter due to sequence tality. Studies show that 7α‐HSDH in Clostridium scindens
differences [76, 77]. may protect from C. dif. infection [85].
280 THE LIVER:  BILE ACID METABOLIC DISEASE AND THERAPY

Recent studies of human and mouse gut microbiomes by 16S Cyp8b1 gene expression, whereas fasting suppressed Cyp7a1
ribosomal RNA sequencing have identified gut microbe‐associ- and stimulated Cyp8b1 expression [101]. The gut microbiome
ated diseases, such as non‐alcoholic fatty liver disease (NAFLD), also exhibits a circadian rhythm in composition, which is damp-
type 2 diabetes, and hepatocellular carcinoma (HCC). In ened by HFD and circadian disruption, causing dysbiosis and
human feces, 90% of the bacteria belong to two phyla: Firmicutes impairment of  barrier function (leaky gut), and is associated
and Bacteroidetes. The relative abundance of Firmicutes and with the pathogenesis of inflammatory bowel diseases, diabetes,
Bacteroidetes is associated with obesity in mouse models and obesity [102]. The feeding and fasting cycle modulates cir-
and  human volunteers [86]. A higher ratio of Firmicutes to cadian rhythmicity and bile acid metabolism. Interestingly,
Bacteroidetes enables the gut microbiota to extract energy more intermittent ­fasting shapes the gut microbiome by increasing the
efficiently from high‐fat diets, thus increasing adiposity. Cholic ratio of gut Firmicutes to Bacteroidetes, which in turn increases
acid feeding is known to increase the ratio of Firmicutes to acetate and lactate. It also promotes white adipose tissue brown-
Bacteroidetes in mice. An animal‐based diet promotes dysbiosis ing and decreases obesity in mice [103].
by increasing the abundance of bile‐tolerant bacteria and by
decreasing Firmicutes, which metabolizes plant polysaccharides
[87]. A high‐fat diet significantly induced TCA, which expanded
Bacteroidetes and Bilophila wadsworthia populations to promote BILE ACID METABOLIC DISEASE
colitis in IL10−/− mice [88]. On the other hand, the antioxidant AND THERAPY
tempol reduced Lactobacillus and decreased BSH activity to
increase Tβ‐MCA, which antagonized intestinal FXR and resulted
Bile acid synthesis defects
in increased bile acid synthesis and improved diet‐induced obe-
sity and diabetes [89]. Interestingly, a recent study reports that Analysis of serum and urine bile acid intermediates has been
bile acids induce uncoupling protein 1 (UCP‐1) in brown adipose used to identify 13 human genes in inborn errors of bile acid
tissue to stimulate energy metabolism in mice housed at thermo- synthesis [104]. Deficiency of primary bile acid synthesis
neutrality, and reducing ambient temperature increases brown causes malabsorption of fat, steroids, and fat‐soluble vitamins,
adipose tissue thermogenesis to reduce weight in diet‐induced leading to steatorrhea and growth retardation. Reduced bile acid
obese mice [90]. It was found that hepatic cholesterol was synthesis and feedback inhibition causes stimulation of CYP7A1
increased during cold‐induced thermogenesis to stimulate bile and CYP8B1 expression and accumulation of bile acid interme-
acid synthesis by inducing Cyp7b1 of the alternative bile acid diates at the points of metabolic block [104]. Deficiency of
synthesis pathway. Increased bile acid synthesis via the alterna- CYP7A1 does not cause severe metabolic disorder in humans.
tive pathway shaped the gut microbiota to stimulate adaptive ther- Homozygous CYP7A1 mutation in humans resulted in decreased
mogenesis in brown adipose tissues in mice [91]. It is not clear bile acids levels, hypercholesterolemia and premature gallstone
why the alternative pathway was preferentially stimulated in disease, supporting the key role of bile acid synthesis in choles-
cold‐induced thermogenesis. Stimulating the alternative bile acid terol metabolism [9]. Individuals who were heterozygous for
synthesis pathway alters bile acid composition by reducing tauro- the CYP7A1 mutation also showed a hyperlipidemic phenotype.
cholic acid synthesis and increasing CDCA, which may be con- CYP7A1 polymorphisms have also been associated with risk of
verted to LCA by gut microbiota to stimulate TGR5‐mediated gallstones, hyperlipidemia, and cardiovascular events [105–
glucagon‐like peptide‐1 (GLP‐1) secretion to promote adipocyte 107]. In contrast, mutation of the CYP7B1 gene in an infant
browning [92]. caused neonatal cholestasis and fibrosis [108, 109], and pro-
gressive spastic paraplegia [110]. In CYP7B1‐deficient patients,
Circadian and nutrient interaction 3β‐monohydroxy‐Δ5 bile acids are accumulated, which are
toxic and cause cholestatic liver injury, giant cell hepatitis,
in the control of bile acid homeostasis severe fibrosis, and cirrhosis. Numerous CYP27A1 mutations
Circadian metabolic homeostasis is maintained by the mamma- identified in humans result in the rare lipid storage disorder
lian biological clock located in the hypothalamic suprachias- ­cerebrotendinous xanthomatosis (CTX) [111], which is charac-
matic nucleus (SCN). The central clock in the SCN is entrained terized by accumulation of cholesterol and cholestanol in xan-
by the daily environmental light/dark cycle and directs the phys- thomas and in the brain. CTX patients have xanthomatosis,
iological timing of peripheral clocks located in nearly all organs premature atherosclerosis and progressive neurological disorders.
and tissues. In the liver, lipid, glucose, cholesterol, and bile acid Deficiency of CYP27A1 prevented both classic and alternative
metabolisms are under circadian control. Disruption in circa- bile acid synthesis, and 5β‐cholestane‐3α,7α,12α‐triol is converted
dian timing (e.g. sleep deprivation, jet leg, shift work) uncou- to bile alcohol and cholestenol, which accumulate in tissues to
ples the central and peripheral clocks to cause dysbiosis and can form xanthomas. CTX can be effectively managed by CDCA
alter metabolic homeostasis and contribute to the pathogenesis therapy, which inhibits CYP7A1 to alleviate accumulation of
of cardiovascular disease, metabolic syndromes, gastrointesti- 7α‐hydroxylated cholesterol metabolites. Several mutations in
nal disorders, non‐alcoholic steatohepatitis (NASH), and HCC 3β‐hydroxy‐Δ5 ‐C27‐steroid oxidoreductase (HSD3B7) and
[93–98]. The Cyp7a1 gene expression and bile acid synthesis Δ4–3‐oxidoreductase (AKR1D1) have been identified. Mutations
exhibit a distinct circadian rhythm, peaking in day in humans of peroxisomal very‐long‐chain acyl‐CoA synthase (SLC27A5),
and night in rodents [99, 100]. This rhythm is altered by high‐fat bile acid‐acyl transporter (ABCD3), α‐methylacyl‐CoA race-
diets, alcohol consumption, and sleep disruption. Feeding mase (AMACR), acyl‐CoA oxidase 2 (ACOX2), D‐bifunctional
­rapidly stimulated Cyp7a1 gene expression but suppressed protein (HSD17B4), bile acyl‐CoA:amino acid N‐acyltransferase
23:  BILE ACID METABOLISM IN HEALTH AND DISEASE 281

(BAAT), and sterol carrier protein X (SCPx) cause Zellweger‐ cholestasis associated with autoimmune destruction of the small
related peroxisome diseases. Zellweger disorders are caused by bile ducts causing portal infiltration and fibrosis. About 95% of
deficiency of PEX genes involved in peroxisome biogenesis and the PBC patients are middle‐aged females. Primary sclerosing
abnormal C27‐bile acid intermediates and C29‐dicarboxylic cholangitis (PSC) is associated with injury and fibrosis of both
acid are accumulated in serum. intrahepatic and extrahepatic bile ducts, resulting in biliary
strictures and obstructed bile flow. PSC is a male‐dominant dis-
ease with a male:female ratio of ~2:1.
Cholestatic liver diseases
Cholestasis is a chronic liver condition resulting from obstructed
Bile acids as therapeutic agents
hepatic bile flow, leading to accumulation of bile acids in the
liver and increased bile acids in the systemic circulation. Non‐alcoholic fatty liver diseases
Chronic cholestasis leads to fibrosis, cirrhosis, liver failure, and
Diabetes and fatty liver are inflammatory diseases associated
higher risk of hepatocellular or cholangiocellular carcinomas.
with dyslipidemia and hepatic insulin resistance. NAFLD is the
Genetic mutations of bile acid transporter genes and autoim-
most common chronic liver disease with prevalence in ~30% of
mune destruction of small bile ductules result in intrahepatic
the US population, and is a risk factor for type 2 diabetes, obe-
cholestasis, while obstruction of extrahepatic bile ducts by com-
sity, and cardiovascular disease [119–121]. NASH is a progres-
mon bile duct stones or tumors of the bile duct or pancreas can
sive form of NAFLD that may lead to cirrhosis and liver cancer.
cause extrahepatic cholestasis [112, 113].
The molecular mechanism of progression from simple steatosis
Congenital cholestasis is usually early‐onset and is associated
to fibrosis in NASH is not completely understood, and there is no
with jaundice, pruritus, and growth failure. Progressive familial
effective drug therapy for NASH. Recent studies indicate NASH
intrahepatic cholestasis (PFIC) and benign recurrent intrahepatic
patients have increased circulating conjugated primary bile acids
cholestasis (BRIC) are autosomal recessive diseases associated
(especially G/TCA and TCDCA), an increased ratio of conju-
with genetic mutations in ATP8B1 (type 1, PFIC1), BSEP (type
gated CA to CDCA, and decreased secondary bile acids [122].
2, PFIC2), and MDR3 (type 3, PFIC3) [112]. The ATP8B1 gene
The cause of altered serum bile acid composition in NASH is not
encodes a p‐type cation transporter and phospholipid flippase
understood. It is likely that the gut microbiota plays a critical role
that transports phosphatidylserine from the outer to the inner
in the control of bile acid biotransformation, bile acid pool size,
leaflet to maintain an asymmetric membrane with higher phos-
and bile acid feedback regulation in NASH and type 2 diabetes.
phatidylcholine on the outer layer of the membrane. The current
Recently, bile acid‐based therapies have been developed to treat
hypothesis is that ATP8B1 loss of function alters membrane
cholestasis and NASH. Activation of FXR and TGR5 signaling
structure, resulting in impaired bile acid transporter function. In
protects against inflammation of liver and intestine.
PFIC2, lack of functional BSEP at the apical hepatocyte mem-
brane causes hepatic bile acid accumulation, giant cell hepatitis,
Ursodeoxycholic acid
and hepatocellular necrosis [114]. PFIC2 is also associated with
higher incidence of gallstones. PFIC3 is associated with muta- Ursodeoxycholic acid (UDCA) (trade name, ursodiol) is a 7β‐
tions of the apical phospholipid transporter MDR3. The resulting epimer of CDCA and a hydrophilic bile acid present in small
biliary injury is due to chronic exposure of cholangiocytes to quantities in human bile. Pharmacological doses of UDCA do
high concentrations of non‐micellar bile acids. Decreased phos- not produce a toxic effect in humans. UDCA has been used for
pholipid secretion also results in unstable micelles that favor cholesterol gallstone dissolution [123]. UDCA is more effective
cholesterol crystallization and small bile duct obstruction. High towards smaller stones and complete dissolution may take
serum levels of γ‐glutamyl transpeptidase (GGT), a marker of ~6–24 months. UDCA can be beneficial in the prevention of
ductular damage, is a characteristic feature of PFIC3 and distin- cholelithiasis in certain high‐risk conditions such as during
guishes it from PFIC1 and PFIC2. Mutation of the TJP2 gene, pregnancy or following bariatric surgery. Currently, UDCA is
which encodes tight junction protein 2 localized to hepatic tight the first‐line therapy for PBC [124]. It significantly improves
junctions, causes progressive cholestasis in humans and has been liver function tests and delays the need for liver transplantation
referred to as PFIC4 [115]. More recently, children with neonatal in PBC patients. UDCA can also provide beneficial effects in
cholestasis associated with FXR mutation have also been ICP and PFIC3. UDCA can increase bile acid pool hydrophilic-
reported [116]. These patients had undetectable hepatic expres- ity and decrease bile acid toxicity [125]. It stimulates bile flow
sion of BSEP, normal or near‐normal serum GGT, and rapidly and promotes biliary HCO3− secretion [126, 127]. and it also
progressed to end‐stage liver disease. exhibits anti‐inflammatory and anti‐apoptosis effects [127,
Among the acquired forms of cholestasis, intrahepatic choles- 128]. However, about 40% of PBC patients do not adequately
tasis of pregnancy (ICP) is a common pregnancy‐related liver respond to UDCA therapy [129].
disease that affects ~1% of all pregnant women. ICP is primarily Many clinical trials have been conducted to test UDCA ther-
diagnosed in the third trimester and pruritus is a major initial apy for treating PSC, but its use in PSC has not been recom-
clinical presentation of ICP. ICP is reversible and rapidly resolves mended in clinical practice guidelines due to the lack of
after delivery. Environmental, hormonal, and genetic factors are consistently perceived benefits [130, 131]. Nor‐ursodeoxy-
thought to be involved in the pathogenesis. Genetic variations of cholic acid (norUDCA) is a side‐chain‐shortened C23 homolog
the BSEP gene may be associated with ICP [117, 118]. of UDCA [132, 133]. In enterohepatic circulation, norUDCA
Primary biliary cholangitis (PBC), previously known as pri- does not undergo significant amidation but is partially glucuro-
mary biliary cirrhosis, is another form of acquired chronic nidated. NorUDCA undergoes cholehepatic shunting and
282 THE LIVER:  REFERENCES

promotes HCO3− secretion [127] and exhibits potent anti‐­ the detergent and signaling properties of bile acids. Nuclear
cholestasis and anti‐inflammatory effects in experimental receptors, the gut microbiota, nutrients, and circadian rhythms
­models [134]. A recent randomized controlled phase 2 trial control bile acid metabolism and homeostasis. New findings
showed that norUDCA significantly and dose‐dependently in  these areas demonstrated the high complexity of bile acid
reduced alkaline phosphatase in PSC patients in a 12‐week metabolism, regulation, and function in physiology and diseases.
treatment [135], which warrants further clinical investigation. The Inborn errors of bile acid metabolism are rare but are usually
use of norUDCA appeared to be safe in PSC patients, though associated with severe hepatic complications. Significant pro-
norUDCA did not improve pruritus. gress has been made in obtaining new knowledge of bile acid
biology and in translating these basic research findings to clinical
Farnesoid X receptor agonists applications. Current studies support the concept of targeting
different steps in the enterohepatic circulation of bile acids to
Several potent FXR agonists have been developed, among which
achieve therapeutic benefits in various liver and metabolic diseases.
is obeticholic acid (OCA), a 6α‐ethyl CDCA derivative that selec-
Further developments in bile acid‐based therapies are anticipated
tively activates FXR with an EC50 of ~100 nM [136]. Activation
in the years to come.
of FXR decreases bile acid synthesis, increases intracellular bile
acid accumulation, and alleviates inflammation, which underlie
the beneficial effects of OCA in experimental cholestasis [136]
and human PBC [137–139]. OCA has recently been approved by ACKNOWLEDGMENTS
the US Food and Drug Administration (FDA) to treat PBC
patients who do not respond to or cannot tolerate UDCA. In phase This work was supported in part by NIH grants, DK44442 and
2 trials, OCA also improved NASH scores and is under further DK58379 to JYLC, and 1R01DK102487‐01 to TL. We thank
testing for the treatment of fatty liver disease [140]. Dr. Alan Hofmann for his critical review of the manuscript.
Bile acid sequestrants
Bile acid sequestrants are a class of drugs that bind negatively
charged bile acids in the intestine and prevent bile acid reab- REFERENCES
sorption. This action leads to stimulated hepatic bile acid syn-
1. Russell, D.W. The enzymes, regulation, and genetics of bile acid synthesis.
thesis, increased hepatic LDL uptake, and decreased plasma
Annu Rev Biochem, 2003;72:1370174.
LDL cholesterol. Cholestyramine and colestipol are classic bile 2. Myant, N.B. and Mitropoulos, K.A. Cholesterol 7 alpha‐hydroxylase. J Lipid
acid sequestrants used to lower plasma cholesterol in humans, Res, 1977;18(2):135–53.
but their use is less common given the availability of other lipid‐ 3. Axelson, M., Mork, B., and Sjovall, J. Occurrence of 3 beta‐hydroxy‐5‐
lowering drugs. Bile acid sequestrants are used to treat pruritus cholestenoic acid, 3 beta,7 alpha‐dihydroxy‐5‐cholestenoic acid, and 7
alpha‐hydroxy‐3‐oxo‐4‐cholestenoic acid as normal constituents in human
in cholestasis. Colesevelam [141], a second‐generation bile acid blood. J Lipid Res, 1988;29(5):629–41.
sequestrant, improves glycemic control in type 2 diabetes. The 4. Matsuzawa, N., Takamura, T., Kurita, S. et al. Lipid‐induced oxidative stress
mechanism of action of bile acid sequestrants is not completely causes steatohepatitis in mice fed an atherogenic diet. Hepatology,
understood, but may be attributed to increased GLP‐1 secretion 2007;46(5):1392–403.
[142] and reduced hepatic glucose production [143]. 5. Mihalik, S.J., Steinberg, S.J., Pei, Z. et al. Participation of two members of
the very long‐chain acyl‐CoA synthetase family in bile acid synthesis and
recycling. J Biol Chem, 2002;277(27):24771–9.
Bariatric surgery 6. Falany, C.N., Johnson, M.R., Barnes, S., and Diasio, R.B. Glycine and tau-
rine conjugation of bile acids by a single enzyme. Molecular cloning and
Bariatric surgeries for weight reduction result in increased serum expression of human liver bile acid CoA:amino acid N‐acyltransferase. J
conjugated bile acids, which is positively correlated to improved Biol Chem, 1994;269(30):19375–9.
insulin resistance in obese patients shortly after bariatric surgery 7. Pellicoro, A., van den Heuvel, F.A., Geuken, M. et al. Human and rat bile
[144, 145]. The early increase of serum bile acids and insulin acid‐CoA:amino acid N‐acyltransferase are liver‐specific peroxisomal
enzymes: implications for intracellular bile salt transport. Hepatology, 2007;
sensitivity after Roux‐en‐Y gastric bypass may be linked to
45(2):340–8.
increased secondary bile acids and GLP‐1 [146]. Vertical sleeve 8. Ferrell, J.M., Boehme, S., Li, F., and Chiang, J.Y. Cholesterol 7α‐hydroxy-
gastrectomy improved insulin sensitivity in high‐fat diet‐fed lase‐deficient mice are protected from high fat/high cholesterol diet‐induced
wild‐type mice but not in Fxr−/− mice or Tgr5−/− mice [147, 148], metabolic disorders. J Lipid Res, 2016;57:1144–54.
suggesting that FXR and TGR5 may be involved. However, the 9. Pullinger, C.R., Eng, C., Salen, G. et al. Human cholesterol 7alpha‐hydroxy-
lase (CYP7A1) deficiency has a hypercholesterolemic phenotype. J Clin
underlying molecular mechanism of bile acid receptor signaling
Invest, 2002;110(1):109–17.
in improving diabetes after bariatric surgery is not clear. 10. Takahashi, S., Fukami, T., Masuo, Y. et al. Cyp2c70 is responsible for the
species difference in bile acid metabolism between mice and humans. J Lipid
Res, 2016;57(12):2130–7.
11. Hofmann, A.F., Hagey, L.R., and Krasowski, M.D. Bile salts of vertebrates:
CONCLUSION structural variation and possible evolutionary significance. J Lipid Res,
2010;51(2):226–46.
This chapter summarizes the physiochemical properties of bile 12. Moschetta, A., Xu, F., Hagey, L.R. et al. A phylogenetic survey of biliary
lipids in vertebrates. J Lipid Res, 2005;46(10):2221–32.
acids and updated the mechanisms of bile acid synthesis, trans- 13. Falany, C.N., Fortinberry, H., Leiter, E.H., and Barnes, S. Cloning, expres-
port, and regulation of homeostasis. The major physiological sion, and chromosomal localization of mouse liver bile acid CoA:amino acid
functions of bile acids in the digestive system depend on both N‐acyltransferase. J Lipid Res, 1997;38(6):1139–48.
23:  BILE ACID METABOLISM IN HEALTH AND DISEASE 283

14. Hardison, W.G. Hepatic taurine concentration and dietary taurine as regula- 39. Slijepcevic, D., Roscam Abbing, R.L.P., Katafuchi, T. et al. Hepatic uptake
tors of bile acid conjugation with taurine. Gastroenterology, 1978; of conjugated bile acids is mediated by both sodium taurocholate cotrans-
75(1):71–5. porting polypeptide and organic anion transporting polypeptides and modulated
15. Ridlon, J.M., Kang, D.J., and Hylemon, P.B. Bile salt biotransformations by by intestinal sensing of plasma bile acid levels in mice. Hepatology, 2017;
human intestinal bacteria. J Lipid Res, 2006;47(2):241–59. 66(5):1631–43.
16. Bjorkhem, I., Einarsson, K., Melone, P., and Hylemon, P. Mechanism of 40. Boyer, J.L., Trauner, M., Mennone, A. et al. Upregulation of a basolateral
intestinal formation of deoxycholic acid from cholic acid in humans: evidence FXR‐dependent bile acid efflux transporter OSTalpha‐OSTbeta in cholesta-
for a 3‐oxo‐delta 4‐steroid intermediate. J Lipid Res, 1989;30(7):1033–9. sis in humans and rodents. Am J Physiol Gastrointest Liver Physiol, 2006;
17. Hylemon, P.B., Melone, P.D., Franklund, C.V., Lund, E., and Bjorkhem, I. 290(6):G1124–30.
Mechanism of intestinal 7 alpha‐dehydroxylation of cholic acid: evidence 41. Cui, Y.J., Aleksunes, L.M., Tanaka, Y., Goedken, M.J., and Klaassen, C.D.
that allo‐deoxycholic acid is an inducible side‐product. J Lipid Res, 1991; Compensatory induction of liver efflux transporters in response to ANIT‐
32(1):89–96. induced liver injury is impaired in FXR‐null mice. Toxicol Sci, 2009;
18. Trauner, M. and Boyer, J.L. Bile salt transporters: molecular characterization, 110(1):47–60.
function, and regulation. Physiol Rev, 2003;83(2):633–71. 42. Kullak‐Ublick, G.A., Stieger, B., Hagenbuch, B., and Meier, P.J. Hepatic
19. Otsuki, M. Pathophysiological role of cholecystokinin in humans. J transport of bile salts. Semin Liver Dis, 2000;20(3):273–92.
Gastroenterol Hepatology, 2000;15(Suppl):D71–83. 43. Thakare, R., Alamoudi, J.A., Gautam, N., Rodrigues, A.D., and Alnouti, Y.
20. Boyer, J.L. Bile formation and secretion. Compr Physiol, 2013;3(3): Species differences in bile acids I. Plasma and urine bile acid composition. J
1035–78. Appl Toxicol, 2018;38(10):1323–35.
21. Childs, S., Yeh, R.L., Georges, E., and Ling, V. Identification of a sister gene 44. Mangelsdorf, D.J., Thummel, C., Beato, M. et  al. The nuclear receptor
to P‐glycoprotein. Cancer Res, 1995;55(10):2029–34. superfamily: the second decade. Cell, 1995;83(6):835–9.
22. Berge, K.E., Tian, H., Graf, G.A. et al. Accumulation of dietary cholesterol 45. Sayin, S.I., Wahlstrom, A., Felin, J. et al. Gut microbiota regulates bile acid
in sitosterolemia caused by mutations in adjacent ABC transporters. Science, metabolism by reducing the levels of tauro‐beta‐muricholic acid, a naturally
2000;290(5497):1771–5. occurring FXR antagonist. Cell Metab, 2013;17(2):225–35.
23. Smit, J.J., Schinkel, A.H., Oude Elferink, R.P. et al. Homozygous disruption 46. Maruyama, T., Miyamoto, Y., Nakamura, T. et  al. Identification of mem-
of the murine mdr2 P‐glycoprotein gene leads to a complete absence of brane‐type receptor for bile acids (M‐BAR). Biochem Biophys Res Commun,
phospholipid from bile and to liver disease. Cell, 1993;75(3):451–62. 2002;298(5):714–19.
24. Shneider, B.L., Dawson, P.A., Christie, D.M. et al. Cloning and molecular 47. Kawamata, Y., Fujii, R., Hosoya, M. et  al. A G protein‐coupled receptor
characterization of the ontogeny of a rat ileal sodium‐dependent bile acid responsive to bile acids. J Biol Chem, 2003;278:9435–40.
transporter. J Clin Invest, 1995;95(2):745–54. 48. Goodwin, B., Jones, S.A., Price, R.R. et  al. A regulatory cascade of the
25. Dawson, P.A., Haywood, J., Craddock, A.L. et al. Targeted deletion of the nuclear receptors FXR, SHP‐1, and LRH‐1 represses bile acid biosynthesis.
ileal bile acid transporter eliminates enterohepatic cycling of bile acids in Mol Cell, 2000;6(3):517–26.
mice. J Biol Chem, 2003;278(36):33920–7. 49. Lu, T.T., Makishima, M., Repa, J.J. et al. Molecular basis for feedback regu-
26. Oelkers, P., Kirby, L.C., Heubi, J.E., and Dawson, P.A. Primary bile acid lation of bile acid synthesis by nuclear receptors. Mol Cell, 2000;6(3):
malabsorption caused by mutations in the ileal sodium‐dependent bile acid 507–15.
transporter gene (SLC10A2). J Clin Invest, 1997;99(8):1880–7. 50. Pandak, W.M., Heuman, D.M., Hylemon, P.B., Chiang, J.Y., and Vlahcevic,
27. Gong, Y.Z., Everett, E.T., Schwartz, D.A., Norris, J.S., and Wilson, F.A. Z.R. Failure of intravenous infusion of taurocholate to down‐regulate choles-
Molecular cloning, tissue distribution, and expression of a 14‐kDa bile acid‐ terol 7 alpha‐hydroxylase in rats with biliary fistulas. Gastroenterology,
binding protein from rat ileal cytosol. Proc Natl Acad Sci U S A, 1994; 1995;108(2):533–44.
91(11):4741–5. 51. Li, T., Francl, J.M., Boehme, S. et al. Glucose and insulin induction of bile
28. Ballatori, N., Christian, W.V., Lee, J.Y. et al. OSTalpha‐OSTbeta: a major acid synthesis: mechanisms and implication in diabetes and obesity. J Biol
basolateral bile acid and steroid transporter in human intestinal, renal, and Chem, 2012;287(3):1861–73.
biliary epithelia. Hepatology, 2005;42(6):1270–9. 52. Inagaki, T., Choi, M., Moschetta, A. et al. Fibroblast growth factor 15 func-
29. Rao, A., Haywood, J., Craddock, A.L. et al. The organic solute transporter tions as an enterohepatic signal to regulate bile acid homeostasis. Cell
alpha‐beta, Ostalpha‐Ostbeta, is essential for intestinal bile acid transport Metab, 2005;2(4):217–25.
and homeostasis. Proc Natl Acad Sci U S A, 2008;105(10):3891–6. 53. Song, K.H., Li, T., Owsley, E., Strom, S., and Chiang, J.Y. Bile acids activate
30. Xia, X., Francis, H., Glaser, S., Alpini, G., and LeSage, G. Bile acid interac- fibroblast growth factor 19 signaling in human hepatocytes to inhibit choles-
tions with cholangiocytes. World J Gastroenterol, 2006;12(22):3553–63. terol 7alpha‐hydroxylase gene expression. Hepatology, 2009;49(1):
31. Soroka, C.J., Mennone, A., Hagey, L.R., Ballatori, N., and Boyer, J.L. Mouse 297–305.
organic solute transporter alpha deficiency enhances renal excretion of bile 54. Choi, M., Moschetta, A., Bookout, A.L. et al. Identification of a hormonal
acids and attenuates cholestasis. Hepatology, 2010;51(1):181–90. basis for gallbladder filling. Nat Med, 2006;12(11):1253–5.
32. Meier, P.J. Molecular mechanisms of hepatic bile salt transport from sinusoi- 55. Kir, S., Beddow, S.A., Samuel, V.T. et al. FGF19 as a postprandial, insulin‐
dal blood into bile. Am J Physiol, 1995;269(6 Pt 1):G801–12. independent activator of hepatic protein and glycogen synthesis. Science,
33. Hagenbuch, B., Stieger, B., Foguet, M., Lubbert, H., and Meier, P.J. 2011;331(6024):1621–4.
Functional expression cloning and characterization of the hepatocyte Na+/ 56. Holt, J.A., Luo, G., Billin, A.N. et al. Definition of a novel growth factor‐
bile acid cotransport system. Proc Natl Acad Sci U S A, 1991;88(23): dependent signal cascade for the suppression of bile acid biosynthesis.
10629–33. Genes Dev, 2003;17(13):1581–91.
34. Hagenbuch, B. and Meier, P.J. Molecular cloning, chromosomal localization, 57. Schaap, F.G., van der Gaag, N.A., Gouma, D.J., and Jansen, P.L. High expres-
and functional characterization of a human liver Na+/bile acid cotransporter. sion of the bile salt‐homeostatic hormone fibroblast growth factor 19 in the liver
J Clin Invest, 1994;93(3):1326–31. of patients with extrahepatic cholestasis. Hepatology, 2009;49(4):1228–35.
35. Meier, P.J. and Stieger, B. Bile salt transporters. Annu Rev Physiol, 58. Kong, B., Wang, L., Chiang, J.Y. et  al. Mechanism of tissue‐specific
2002;64:635–61. farnesoid X receptor in suppressing the expression of genes in bile‐acid syn-
36. Suga, T., Yamaguchi, H., Sato, T. et al. Preference of conjugated bile acids thesis in mice. Hepatology, 2012;56(3):1034–43.
over unconjugated bile acids as substrates for OATP1B1 and OATP1B3. 59. Yu, C., Wang, F., Kan, M. et al. Elevated cholesterol metabolism and bile
PLoS One, 2017;12(1):e0169719. acid synthesis in mice lacking membrane tyrosine kinase receptor FGFR4. J
37. Vaz, F.M., Paulusma, C.C., Huidekoper, H. et  al. Sodium taurocholate Biol Chem, 2000;275(20):15482–9.
cotransporting polypeptide (SLC10A1) deficiency: conjugated hyper- 60. Li, S., Hsu, D.D., Li, B. et al. Cytoplasmic tyrosine phosphatase Shp2 coor-
cholanemia without a clear clinical phenotype. Hepatology, 2015; dinates hepatic regulation of bile acid and FGF15/19 signaling to repress bile
61(1):260–7. acid synthesis. Cell Metab, 2014;20(2):320–32.
38. Slijepcevic, D., Kaufman, C., Wichers, C.G. et al. Impaired uptake of conjugated 61. Lin, B.C., Wang, M., Blackmore, C., and Desnoyers, L.R. Liver‐specific
bile acids and Hepatitis B Virus preS1‐binding in Na ‐taurocholate cotrans- activities of FGF19 require Klotho beta. J Biol Chem, 2007;282(37):
porting polypeptide knockout mice. Hepatology, 2015;62(1):207–19. 27277–84.
284 THE LIVER:  REFERENCES

62. Ito, S., Fujimori, T., Furuya, A. et al. Impaired negative feedback suppres- 85. Studer, N., Desharnais, L., Beutler, M. et al. Functional Intestinal Bile Acid
sion of bile acid synthesis in mice lacking betaKlotho. J Clin Invest, 7alpha‐Dehydroxylation by Clostridium scindens associated with protec-
2005;115(8):2202–8. tion from Clostridium difficile infection in a gnotobiotic mouse model.
63. Lan, T., Rao, A., Haywood, J., Kock, N.D., and Dawson, P.A. Mouse organic Front Cell Infect Microbiol, 2016;6:191.
solute transporter alpha deficiency alters FGF15 expression and bile acid 86. Ley, R.E., Turnbaugh, P.J., Klein, S., and Gordon, J.I. Microbial ecology:
metabolism. J Hepatol, 2012;57(2):359–65. human gut microbes associated with obesity. Nature, 2006;
64. Kong, B. and Guo, G.L. Soluble expression of disulfide bond containing 444(7122):1022–3.
proteins FGF15 and FGF19 in the cytoplasm of Escherichia coli. PLoS One, 87. David, L.A., Maurice, C.F., Carmody, R.N. et al. Diet rapidly and repro-
2014;9(1):e85890. ducibly alters the human gut microbiome. Nature, 2014(505):559–63.
65. Potthoff, M.J., Boney‐Montoya, J., Choi, M. et  al. FGF15/19 regulates 88. Devkota, S., Wang, Y., Musch, M.W. et al. Dietary‐fat‐induced taurocholic
hepatic glucose metabolism by inhibiting the CREB‐PGC‐1alpha pathway. acid promotes pathobiont expansion and colitis in I110‐/‐ mice. Nature,
Cell Metab, 2011;13(6):729–38. 2012;487(7405):104–8.
66. Baker, D.M., Wang, S.L., Bell, D.J., Drevon, C.A., and Davis, R.A. One or 89. Li, F., Jiang, C., Krausz, K.W. et al. Microbiome remodelling leads to inhi-
more labile proteins regulate the stability of chimeric mRNAs containing the bition of intestinal farnesoid X receptor signalling and decreased obesity.
3’‐untranslated region of cholesterol‐7alpha ‐hydroxylase mRNA. J Biol Nat Commun, 2013;4:2384.
Chem, 2000;275(26):19985–91. 90. Zietak, M., Kovatcheva‐Datchary, P., Markiewicz, L.H. et  al. Altered
67. Tarling, E.J., Clifford, B.L., Cheng, J. et al. RNA‐binding protein ZFP36L1 microbiota contributes to reduced diet‐induced obesity upon cold exposure.
maintains posttranscriptional regulation of bile acid metabolism. J Clin Cell Metab, 2016;23(6):1216–23.
Invest, 2017;127(10):3741–54. 91. Worthmann, A., John, C., Ruhlemann, M.C. et al. Cold‐induced conversion
68. Wang, M., Tan, Y., Costa, R.H., and Holterman, A.X. In vivo regulation of of cholesterol to bile acids in mice shapes the gut microbiome and promotes
murine CYP7A1 by HNF‐6: a novel mechanism for diminished CYP7A1 adaptive thermogenesis. Nat Med, 2017;23(7):839–49.
expression in biliary obstruction. Hepatology, 2004;40(3):600–8. 92. Pathak, P., Cen, X., Nichols, R.G. et al. Intestine farnesoid X receptor ago-
69. Li, J., Woolbright, B.L., Zhao, W. et al. Sortilin 1 loss‐of‐function protects nist and the gut microbiota activate G‐protein bile acid receptor‐1 signaling
against cholestatic liver injury by attenuating hepatic bile acid accumulation to improve metabolism. Hepatology, 2018;68(4):1574–88.
in bile duct ligated mice. Toxicol Sci, 2018;161(1):34–47. 93. Turek, F.W., Joshu, C., Kohsaka, A. et al. Obesity and metabolic syndrome
70. Miyake, J.H., Wang, S.L., and Davis, R.A. Bile acid induction of cytokine in circadian Clock mutant mice. Science, 2005;308(5724):1043–5.
expression by macrophages correlates with repression of hepatic cholesterol 94. Morris, C.J., Purvis, T.E., Hu, K., and Scheer, F.A. Circadian misalignment
7alpha‐hydroxylase. J Biol Chem, 2000;275(29):21805–8. increases cardiovascular disease risk factors in humans. Proc Natl Acad Sci
71. Peet, D.J., Turley, S.D., Ma, W. et al. Cholesterol and bile acid metabolism U S A, 2016;113(10):E1402–11.
are impaired in mice lacking the nuclear oxysterol receptor LXR alpha. Cell, 95. Bass, J. and Takahashi, J.S. Circadian integration of metabolism and ener-
1998;93(5):693–704. getics. Science, 2010;330(6009):1349–54.
72. Lehmann, J.M., Kliewer, S.A., Moore, L.B. et al. Activation of the nuclear 96. Panda, S., Antoch, M.P., Miller, B.H. et al. Coordinated transcription of key
receptor LXR by oxysterols defines a new hormone response pathway. J Biol pathways in the mouse by the circadian clock. Cell, 2002;109(3):307–20.
Chem, 1997;272(6):3137–40. 97. Kettner, N.M., Voicu, H., Finegold, M.J. et  al. Circadian homeostasis of
73. Repa, J.J., Berge, K.E., Pomajzl, C. et al. Regulation of ATP‐binding cassette liver metabolism suppresses hepatocarcinogenesis. Cancer Cell,
sterol transporters ABCG5 and ABCG8 by the liver X receptors alpha and 2016;30(6):909–24.
beta. J Biol Chem, 2002;277(21):18793–800. 98. Maury, E., Hong, H.K., and Bass, J. Circadian disruption in the pathogen-
74. Repa, J.J., Turley, S.D., Lobaccaro, J.A. et al. Regulation of Absorption and esis of metabolic syndrome. Diabetes Metab, 2014;40(5):338–46.
ABC1‐mediated efflux of cholesterol by RXR heterodimers. Science, 99. Galman, C., Angelin, B., and Rudling, M. Bile acid synthesis in humans has
2000;289(5484):1524–9. a rapid diurnal variation that is asynchronous with cholesterol synthesis.
75. Venkateswaran, A., Laffitte, B.A., Joseph, S.B. et al. Control of cellular cho- Gastroenterology, 2005;129(5):1445–53.
lesterol efflux by the nuclear oxysterol receptor LXR alpha. Proc Natl Acad 100. Lundasen, T., Galman, C., Angelin, B., and Rudling, M. Circulating intes-
Sci U S A, 2000;97(22):12097–102. tinal fibroblast growth factor 19 has a pronounced diurnal variation and
76. Chiang, J.Y., Kimmel, R., and Stroup, D. Regulation of cholesterol 7alpha‐ modulates hepatic bile acid synthesis in man. J Intern Med,
hydroxylase gene (CYP7A1) transcription by the liver orphan receptor 2006;260(6):530–6.
(LXRalpha). Gene, 2001;262(1–2):257–65. 101. Pathak, P., Li, T., and Chiang, J.Y. Retinoic acid‐related orphan receptor
77. Agellon, L.B., Drover, V.A., Cheema, S.K., Gbaguidi, G.F., and Walsh, A. alpha regulates diurnal rhythm and fasting induction of sterol 12alpha‐
Dietary cholesterol fails to stimulate the human cholesterol 7alpha‐hydroxylase hydroxylase in bile acid synthesis. J Biol Chem, 2013;288(52):37154–65.
gene (CYP7A1) in transgenic mice. J Biol Chem, 2002;277(23):20131–4. 102. Hatori, M., Vollmers, C., Zarrinpar, A. et al. Time‐restricted feeding with-
78. Ananthanarayanan, M., Balasubramanian, N., Makishima, M., Mangelsdorf, out reducing caloric intake prevents metabolic diseases in mice fed a high‐
D.J., and Suchy, F.J. Human bile salt export pump promoter is transactivated fat diet. Cell Metab, 2012;15(6):848–60.
by the farnesoid X receptor/bile acid receptor. J Biol Chem, 103. Li, G., Xie, C., Lu, S. et  al. Intermittent fasting promotes white adipose
2001;276(31):28857–65. browning and decreases obesity by shaping the gut microbiota. Cell Metab,
79. Denson, L.A., Sturm, E., Echevarria, W. et al. The orphan nuclear receptor, 2017;26(4):672–85 e4.
shp, mediates bile acid‐induced inhibition of the rat bile acid transporter, 104. Vaz, F.M. and Ferdinandusse, S. Bile acid analysis in human disorders of
ntcp. Gastroenterology, 2001;121(1):140–7. bile acid biosynthesis. Mol Aspects Med, 2017;56:10–24.
80. Lee, H., Zhang, Y., Lee, F.Y. et al. FXR regulates organic solute transporters 105. Hofman, M.K., Princen, H.M., Zwinderman, A.H., and Jukema, J.W.
alpha and beta in the adrenal gland, kidney, and intestine. J Lipid Res, Genetic variation in the rate‐limiting enzyme in cholesterol catabolism
2006;47(1):201–14. (cholesterol 7alpha‐hydroxylase) influences the progression of atheroscle-
81. Grober, J., Zaghini, I., Fujii, H. et al. Identification of a bile acid‐responsive rosis and risk of new clinical events. Clin Sci, 2005;108(6):539–45.
element in the human ileal bile acid‐binding protein gene. Involvement of the 106. Kajinami, K., Brousseau, M.E., Ordovas, J.M., and Schaefer, E.J.
farnesoid x receptor/9‐cis‐retinoic acid receptor heterodimer. J Biol Chem, Interactions between common genetic polymorphisms in ABCG5/G8 and
1999;274(42):29749–54. CYP7A1 on LDL cholesterol‐lowering response to atorvastatin.
82. Neimark, E., Chen, F., Li, X., and Shneider, B.L. Bile acid‐induced negative Atherosclerosis, 2004;175(2):287–93.
feedback regulation of the human ileal bile acid transporter. Hepatology, 107. Jiang, Z.Y., Han, T.Q., Suo, G.J. et  al. Polymorphisms at cholesterol
2004;40(1):149–56. 7alpha‐hydroxylase, apolipoproteins B and E and low density lipoprotein
83. Slijepcevic, D., Roscam Abbing, R.L.P., Fuchs, C.D. et al. Na+‐taurocholate receptor genes in patients with gallbladder stone disease. World J
cotransporting polypeptide inhibition has hepatoprotective effects in choles- Gastroenterol, 2004;10(10):1508–12.
tasis in mice. Hepatology, 2018; 68(3):1057–69. 108. Setchell, K.D.R., Schwarz, M., O’Connell, N.C. et al. Identification of a
84. Wahlstrom, A., Sayin, S.I., Marschall, H.U., and Backhed, F. Intestinal new inborn error in bile acid synthesis: mutation of the oxysterol 7a‐
crosstalk between bile acids and microbiota and its impact on host metabo- hydroxylase gene causes severe neonatal liver disease. J Clin Invest,
lism. Cell Metab, 2016;24(1):41–50. 1998;102(9):1690–703.
23:  BILE ACID METABOLISM IN HEALTH AND DISEASE 285

109. Ueki, I., Kimura, A., Nishiyori, A. et al. Neonatal cholestatic liver disease 130. European Association for the Study of the Liver. EASL Clinical Practice
in an Asian patient with a homozygous mutation in the oxysterol 7alpha‐ Guidelines: management of cholestatic liver diseases. J Hepatol,
hydroxylase gene. J Pediatr Gastroenterol Nutr, 2008;46(4):465–9. 2009;51(2):237–67.
110. Tsaousidou, M.K., Ouahchi, K., Warner, T.T. et  al. Sequence alterations 131. Chapman, R., Fevery, J., Kalloo, A. et al. Diagnosis and management of
within CYP7B1 implicate defective cholesterol homeostasis in motor‐neu- primary sclerosing cholangitis. Hepatology, 2010;51(2):660–78.
ron degeneration. Am J Hum Genet, 2008;82(2):510–15. 132. Yoon, Y.B., Hagey, L.R., Hofmann, A.F. et al. Effect of side‐chain shorten-
111. Leitersdorf, E., Safadi, R., Meiner, V. et al. Cerebrotendinous xanthomato- ing on the physiologic properties of bile acids: hepatic transport and effect
sis in the Israeli Druze: molecular genetics and phenotypic characteristics. on biliary secretion of 23‐nor‐ursodeoxycholate in rodents. Gastroenterology,
Am J Hum Genet, 1994;55(5):907–15. 1986;90(4):837–52.
112. Srivastava, A. Progressive familial intrahepatic cholestasis. J Clin Exp 133. Yeh, H.Z., Schteingart, C.D., Hagey, L.R. et al. Effect of side chain length
Hepatol, 2014;4(1):25–36. on biotransformation, hepatic transport, and choleretic properties of cheno-
113. Zollner, G. and Trauner, M. Mechanisms of cholestasis. Clin Liver Dis, deoxycholyl homologues in the rodent: studies with dinorchenodeoxy-
2008;12(1):1–26, vii. cholic acid, norchenodeoxycholic acid, and chenodeoxycholic acid.
114. Strautnieks, S.S., Kagalwalla, A.F., Tanner, M.S. et al. Identification of a Hepatology, 1997;26(2):374–85.
locus for progressive familial intrahepatic cholestasis PFIC2 on chromo- 134. Halilbasic, E., Fiorotto, R., Fickert, P. et al. Side chain structure determines
some 2q24. Am J Hum Genet, 1997;61(3):630–3. unique physiologic and therapeutic properties of norursodeoxycholic acid
115. Carlton, V.E., Harris, B.Z., Puffenberger, E.G. et al. Complex inheritance of in Mdr2‐/‐ mice. Hepatology, 2009;49(6):1972–81.
familial hypercholanemia with associated mutations in TJP2 and BAAT. 135. Fickert, P., Hirschfield, G.M., Denk, G. et  al. norUrsodeoxycholic acid
Nat Genet, 2003;34(1):91–6. improves cholestasis in primary sclerosing cholangitis. J Hepatol,
116. Gomez‐Ospina, N., Potter, C.J., Xiao, R. et al. Mutations in the nuclear bile 2017;67(3):549–58.
acid receptor FXR cause progressive familial intrahepatic cholestasis. Nat 136. Pellicciari, R., Fiorucci, S., Camaioni, E. et al. 6alpha‐ethyl‐chenodeoxy-
Commun, 2016;7:10713. cholic acid (6‐ECDCA), a potent and selective FXR agonist endowed with
117. Pauli‐Magnus, C., Lang, T., Meier, Y. et al. Sequence analysis of bile salt anticholestatic activity. J Med Chem, 2002;45(17):3569–72.
export pump (ABCB11) and multidrug resistance p‐glycoprotein 3 137. Hirschfield, G.M., Mason, A., Luketic, V. et al. Efficacy of obeticholic acid
(ABCB4, MDR3) in patients with intrahepatic cholestasis of pregnancy. in patients with primary biliary cirrhosis and inadequate response to urso-
Pharmacogenetics, 2004;14(2):91–102. deoxycholic acid. Gastroenterology, 2015;148(4):751–61 e8.
118. Noe, J., Kullak‐Ublick, G.A., Jochum, W. et al. Impaired expression and 138. Kowdley, K.V., Luketic, V., Chapman, R. et al. A randomized trial of obet-
function of the bile salt export pump due to three novel ABCB11 mutations icholic acid monotherapy in patients with primary biliary cholangitis.
in intrahepatic cholestasis. J Hepatol, 2005;43(3):536–43. Hepatology, 2018;67(5):1890–902.
119. Farrell, G.C. and Larter, C.Z. Nonalcoholic fatty liver disease: from steato- 139. Nevens, F., Andreone, P., Mazzella, G. et al. A placebo‐controlled trial of
sis to cirrhosis. Hepatology, 2006;43(2 Suppl 1):S99–S112. obeticholic acid in primary biliary cholangitis. N Engl J Med,
120. Cohen, J.C., Horton, J.D., and Hobbs, H.H. Human fatty liver disease: old 2016;375(7):631–43.
questions and new insights. Science, 2011;332(6037):1519–23. 140. Neuschwander‐Tetri, B.A., Loomba, R., Sanyal, A.J. et  al. Farnesoid X
121. Brunt, E.M. Pathology of nonalcoholic fatty liver disease. Nat Rev nuclear receptor ligand obeticholic acid for non‐cirrhotic, non‐alcoholic
Gastroenterol Hepatol, 2010;7(4):195–203. steatohepatitis (FLINT): a multicentre, randomised, placebo‐controlled
122. Puri, P., Daita, K., Joyce, A. et al. The presence and severity of nonalco- trial. Lancet, 2015;385(9972):956–65.
holic steatohepatitis is associated with specific changes in circulating bile 141. Fonseca, V.A., Handelsman, Y., and Staels, B. Colesevelam lowers glucose
acids. Hepatology, 2017. and lipid levels in type 2 diabetes: the clinical evidence. Diabetes Obes
123. Lioudaki, E., Ganotakis, E.S., and Mikhailidis, D.P. Lipid lowering drugs Metab, 2010;12(5):384–92.
and gallstones: a therapeutic option? Curr Pharm Des, 2011;17(33): 142. Shang, Q., Saumoy, M., Holst, J.J., Salen, G., and Xu, G. Colesevelam
3622–31. improves insulin resistance in a diet‐induced obesity (F‐DIO) rat model by
124. Dyson, J.K., Hirschfield, G.M., Adams, D.H. et al. Novel therapeutic tar- increasing the release of GLP‐1. Am J Physiol Gastrointest Liver Physiol,
gets in primary biliary cirrhosis. Nat Rev Gastroenterol Hepatol, 2015; 2010;298(3):G419–24.
12(3):147–58. 143. Potthoff, M.J., Potts, A., He, T. et al. Colesevelam suppresses hepatic gly-
125. Dilger, K., Hohenester, S., Winkler‐Budenhofer, U. et al. Effect of ursode- cogenolysis by TGR5‐mediated induction of GLP‐1 action in DIO mice.
oxycholic acid on bile acid profiles and intestinal detoxification machinery Am J Physiol Gastrointest Liver Physiol, 2013;304(4):G371–80.
in primary biliary cirrhosis and health. J Hepatol, 2012;57(1):133–40. 144. Patti, M.E., Houten, S.M., Bianco, A.C. et al. Serum bile acids are higher in
126. Prieto, J., Garcia, N., Marti‐Climent, J.M. et  al. Assessment of biliary humans with prior gastric bypass: potential contribution to improved glu-
bicarbonate secretion in humans by positron emission tomography. cose and lipid metabolism. Obesity (Silver Spring), 2009;17(9):1671–7.
Gastroenterology, 1999;117(1):167–72. 145. Simonen, M., Dali‐Youcef, N., Kaminska, D. et al. Conjugated bile acids
127. Beuers, U. Drug insight: mechanisms and sites of action of ursodeoxy- associate with altered rates of glucose and lipid oxidation after Roux‐en‐Y
cholic acid in cholestasis. Nat Clin Pract Gastroenterol Hepatol, gastric bypass. Obes Surg, 2012;22(9):1473–80.
2006;3(6):318–28. 146. Albaugh, V.L., Flynn, C.R., Cai, S. et al. Early increases in bile acids post
128. Poupon, R. Ursodeoxycholic acid and bile‐acid mimetics as therapeu- Roux‐en‐Y gastric bypass are driven by insulin‐sensitizing, secondary bile
tic agents for cholestatic liver diseases: an overview of their mecha- acids. J Clin Endocrinol Metab, 2015;100(9):E1225–33.
nisms of action. Clin Res Hepatol Gastroenterol, 2012;36(Suppl 147. Ryan, K.K., Tremaroli, V., Clemmensen, C. et al. FXR is a molecular target
1):S3–12. for the effects of vertical sleeve gastrectomy. Nature, 2014;509:183–8.
129. Pares, A., Caballeria, L., and Rodes, J. Excellent long‐term survival in 148. McGavigan, A.K., Garibay, D., Henseler, Z.M. et al. TGR5 contributes to
patients with primary biliary cirrhosis and biochemical response to ursode- glucoregulatory improvements after vertical sleeve gastrectomy in mice.
oxycholic acid. Gastroenterology, 2006;130(3):715–20. Gut, 2015;66(2):226–34.
TGR5 (GPBAR1)
24 in the Liver
Verena Keitel1, Christoph G.W. Gertzen2, Lina Spomer1, Holger Gohlke2,
and Dieter Häussinger1
1
Clinic for Gastroenterology, Hepatology and Infectious Diseases, University Hospital Düsseldorf, Medical
Faculty at Heinrich‐Heine‐University, Düsseldorf, Germany
2
Institute of Pharmaceutical and Medicinal Chemistry, Heinrich‐Heine University Düsseldorf, Düsseldorf,
Germany

INTRODUCTION intestine [15–20]. Thus, FXR acts as important metabolic regulator


and the FXR ligand obeticholic acid (6α‐ethyl‐CDCA), which
Bile acids are synthesized in hepatocytes from cholesterol, lead- has already been approved for the treatment of primary biliary
ing to the formation of cholic acid (CA) and chenodeoxycholic cholangitis (PBC), significantly improved histological features
acid (CDCA) in humans, while in mice CDCA is mostly con- of non‐alcoholic steatohepatitis (NASH) in a phase 2 trial [21,
verted to α‐muricholic acid (αMCA) and β‐muricholic acid 22]. Further NRs activated by bile acids are the pregnane X
(βMCA) [1, 2]. These primary bile acids are subsequently con- receptor (PXR, NR1I2) and the vitamin D receptor (VDR,
jugated at carbon 24 to either taurine or glycine and secreted NR1I1) [23–25]. Several G protein‐coupled receptors (GPCRs)
into bile, forming mixed micelles with phospholipids and cho- are also responsive to bile acids, including different types of
lesterol, which are released into the small intestine following muscarinic (acetylcholine) receptors (e.g. M2 and M3 recep-
food intake, where they facilitate absorption of dietary lipids as tors) [26–29], formyl‐peptide receptors (FPR) [11, 30, 31], the
well as cholesterol excretion [1–3]. In the intestine, bile acids sphingosine‐1‐phosphate receptor 2 (S1PR2) [32–37] and the
are deconjugated by microbial bile salt hydrolase (BSH) to form Takeda G protein‐coupled receptor 5 (TGR5) [38–40]. TGR5 is
unconjugated bile acids, which subsequently can be converted also known as G protein‐coupled bile acid receptor 1 (GPBAR1)
into the secondary bile acids deoxycholic acid (DCA) and litho- or membrane‐type bile acid receptor (M‐BAR) and was the first
cholic acid (LCA) through 7α‐dehydroxylation [4, 5]. Bile acids GPCR identified as bile acid receptor in 2002/2003 [38, 40].
are actively reabsorbed in the terminal ileum and transported
back to the liver via the portal venous blood [6]. In the liver, bile
acids are taken up into hepatocytes, reconjugated with taurine or
glycine and again secreted across the canalicular hepatocyte TGR5 EXPRESSION AND LOCALIZATION
membrane into bile [6]. This enterohepatic circulation of bile IN RODENT AND HUMAN LIVER
acids takes place 6–10 times per day in humans [3, 6]. Only a
small amount of secondary bile acids is excreted with the feces The human TGR5 gene encompasses two exons and is located in
and hence replaced by de novo synthesis of bile acids from cho- the chromosomal region 2q35 [41]. Exon 2 contains the entire
lesterol in the liver [2]. coding region of 993 base pairs, which translates to 330 amino
Over the past decades, bile acids have emerged as important acids [40, 41]. In contrast, rat and murine Tgr5 genes are present
signaling molecules that activate different classes of receptors, on chromosome 9q33 and chromosome 1qC3, respectively, and
allowing for a bile acid‐ and cell type‐specific response and contain coding regions of 990 base pairs each, resulting in pro-
explaining the pleiotropic effects of bile acids in the organism teins of 329 amino acids [42]. The protein sequences of human,
[7–11]. Bile acids were first identified as ligands for the nuclear bovine, rabbit, rat, and mouse TGR5 are highly conserved, with
receptor (NR) farnesoid X receptor (FXR, NR1H4) in 1999 amino acid identities of 82–91% [38, 40].
[12–14]. FXR transcriptionally regulates key genes involved in TGR5 mRNA is expressed almost ubiquitously in rodent and
bile acid, glucose, and lipid metabolism in both liver and human tissues [10, 39–40, 43, 44]. High mRNA levels were

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
24:  TGR5 (GPBAR1) in the Liver 287

found in organs involved in the enterohepatic circulation and TGR5‐binding mode model
excretion of bile acids such as liver, gallbladder, small and large
intestine, and kidney [38, 40, 42, 44–46]. Moreover, TGR5 Bile acids have been predicted to bind to the orthosteric site in
mRNA was also present in heart, lung, spleen, CD14‐positive the GPCR TGR5 [61]. There, bile acids bridge the transmem-
monocytes, placenta, female and male reproductive organs, brane helices (TM) 3 and 6 to activate the receptor, as is common
adrenal glands, brain, and adipose tissue [38, 40, 42, 44–46]. for GPCR agonists [62], while addressing additional TMs to
Expression levels in whole liver tissue were lowest in mice, increase their efficacy [61] (Figure 24.3a,b). Taurine conjugation
higher in rats and highest in humans [39]. increases the size of a bile acid compared to unconjugated bile
On the protein level, TGR5 was detected in different non‐ acids, which allows the bridging of residues R79 (extracellular
parenchymal cells of human and rodent livers, such as liver loop EL1) and Y240 (TM6) in TGR5 [61]. The salt‐bridge inter-
sinusoidal endothelial cells (LSECs), liver‐resident mac- action between the negatively charged sulfonic acid moiety and
rophages (Kupffer cells), biliary epithelial cells of small and the positively charged R79 likely increases the affinity of those
large intrahepatic bile ducts (cholangiocytes, BECs), BECs of bile acids towards TGR5. The localization of the sulfonic acid
the extrahepatic bile duct, and activated hepatic stellate cells moiety of taurine‐conjugated bile acids high up in the b­ inding
(HSCs) (Figure  24.1) [10, 39, 47–52]. TGR5 immunofluores- pocket of TGR5 may also explain why TGR5 can be activated by
cence staining intensity was minimal or absent in hepatocytes cholestyramine‐bound bile acids [63]; it may still be possible for
and quiescent HSCs (Figure 24.1) [10, 39, 48, 53]. In the gall- bile acids to reach TM6 with their cholane ­scaffold even when
bladder, TGR5 was localized in the epithelium and in smooth bound to cholestyramine with their acidic moiety. All bile acids
muscle cells of the muscularis propria [54, 55]. employ the 3‐hydroxyl groups of their cholane scaffold to form a
On the subcellular level, TGR5 was predominantly found in hydrogen bond to Y240, and this interaction is further stabilized
the plasma membrane [38, 40]. In polarized cells, such as gall- by a hydrogen bond to E169 (TM5; Figure  24.3a,b) [61]. The
bladder epithelial cells and cholangiocytes, TGR5 was localized hydrogen bond interaction with Y240 is vital for the activation of
in the apical membrane and on the primary cilia, a sensory orga- the receptor because a Y240F variant, lacking the hydroxyl
nelle extending from the apical plasma membrane into the bile group in the phenylalanine ring, is unresponsive to bile acids
duct lumen [47, 49, 51, 54, 56–58]. Interestingly, TGR5 has also [61]. Agonistic neurosteroids such as pregnanediol also utilize
been detected in the nuclear membrane of cholangiocytes; how- their hydroxyl or carbonyl groups to interact with Y240 in TGR5
ever, it is unknown whether the receptor can be activated intra- [61]. Lacking acidic groups, they mainly form additional hydro-
cellularly [56]. phobic contacts with Y89 in TM3 to bind to and activate TGR5
at a reasonable EC50 (e.g. pregnanediol EC50 = 0.58 μM [59]
(Figure 24.2)), allowing them to activate TGR5 in the brain. The
epimeric selectivity has been explained by a hydrogen bond for-
NATURAL AND SYNTHETIC mation of CDCA’s 7α‐hydroxyl group to Y89 in TM3 of TGR5.
In contrast, due to the β‐configuration, UDCA cannot form such
TGR5 LIGANDS a hydrogen bond with its 7‐hydroxyl group.

Endogenous ligands of TGR5


A wide spectrum of hydrophobic endogenous ligands can acti-
Synthetic TGR5 agonists and antagonists
vate TGR5 [59]. Those ligands comprise all known bile acids To date, a range of TGR5 agonists with a nonsteroidal core are
and many hydrophobic neurosteroids, such as pregnanolone, known [64–68] (Figure 24.2). Here, an acidic or amide moiety is
allopregnanolone, pregnanediol, and estradiol [45, 59, 60]. In linked to a system of three to four variably interconnected aro-
contrast to other bile acid receptors, TGR5 is rather promiscuous matic and aliphatic rings. The ring furthest from the acid or
as it recognizes bile acids regardless of their substitution pattern amide moiety always contains a hetero‐atom. Although the
and conjugation state [59]. However, the potency of bile acids binding mode of nonsteroidal TGR5 agonists is unknown, it is pos-
increases with the hydrophobicity of their cholane scaffold and sible that the hetero‐atom is necessary to form a hydrogen bond
varies with the state and type of conjugation. As such, taurolitho- to Y240 (TM6), which is crucial for the activation of TGR5 [69].
cholic acid (TLCA; Figure 24.2), a secondary bile acid with only Intestinal specificity can be achieved either by addition of a qua-
a hydroxyl group in position 3 of the cholane scaffold and with ternary ammonia moiety to the synthetic TGR5 agonists (26a,
taurine conjugation, is the most potent natural agonist of TGR5 Figure 24.2) or by linking two agonists via polyethylene glycol
(EC50 = 0.29 μM, Figure 24.2); its unconjugated derivative, litho- (15c, Figure 24.2). Ligands with quaternary ammonia moieties
cholic acid (LCA), has a potency of EC50 = 0.58 μM (Figure 24.2) can usually not be absorbed due to the permanent charge, while
[59]. Addition of hydroxyl groups in position 12 in the secondary the interconnected agonists are likely too spacious to pass the
bile acid deoxycholic acid (DCA) or position 7 in the primary cell membrane in the intestine and enter the bloodstream.
bile acid CDCA increases the EC50 4‐fold and 23‐fold compared So far, despite the broad spectrum of ligands recognized by
to TLCA, respectively [59] (Figure 24.2). TGR5 also shows an TGR5, only one antagonist has been discovered (SBI‐115,
epimeric selectivity for bile acids with an α‐hydroxyl group in Figure  24.2) [70]. It is composed of two interconnected aro-
position 7 of the cholane scaffold, as seen in the 5‐fold higher matic rings and an ethyl sulfo substituent and is slightly smaller
efficacy of CDCA than ursodeoxycholic acid (UDCA) than the known agonists. Due to the lack of an acid moiety and
(Figure  24.2) [59]. The effects of structural modifications on a steroidal core, the binding mode of this antagonist cannot
agonistic potency are explained by a binding mode model. easily be inferred from the agonist binding mode.
288 THE LIVER:  TGR5‐DEPENDENT SIGNALING AND REGULATION OF TGR5 EXPRESSION AND LOCALIZATION

Figure 24.1  Localization of TGR5 in different liver cells. The receptor has been detected in sinusoidal endothelial cells (LSEC), in liver‐resident
macrophages (Kupffer cells, KCs), in biliary epithelial cells (BEC) and in activated hepatic stellate cells (HSCs). In contrast, little or no immunofluo-
rescence staining could be observed in hepatic parenchymal cells (HPCs, hepatocytes) and quiescent HSC. A–D Staining for TGR5 in rat liver is shown
in red. CD163, Reca‐1, CK‐19, α‐SMA, and GFAP served as marker proteins for KCs, LSECs, BECs, activated HSCs, and quiescent HSCs, respec-
tively. Panels (a–c) show livers from control animals. Panel (d) depicts a rat liver from an animal treated with carbon tetrachloride. Inset in (a): TGR5
staining is increased in KC after three days of common bile duct ligation. Inset in (d): Quiescent HSCs (GFAP, in green) showed no staining for TGR5
(red), which was present in LSECs (stained in blue). Panels (e–h) show immunofluorescence staining of TGR5 in human liver (e–g) and human gall-
bladder (h). CD163 served as marker protein for KCs, while CK‐7 was used to visualize BECs. MRP2 and Na/K‐ATPase were used to stain the apical
and basolateral membrane of gallbladder epithelial cells, respectively (h). Panels (i–m) show staining of TGR5 in BECs. TGR5 is localized in the pri-
mary cilium of cultured murine cholangiocytes (i, j) and of a human cholangiocarcinoma cell line (TFK‐1, (k)). Phalloidin was used to stain F‐actin fila-
ments (i) while acetylated α‐tubulin (α‐tub) served as marker protein for primary cilia (j, k). Panels (l) and (m) depict reduced fluorescence intensity
for TGR5 in BECs of a PSC liver (l) and an increase in fluorescence staining of TGR5 in tumor cells of human cholangiocarcinoma (m). Bars = 10 μm.

TGR5‐DEPENDENT SIGNALING can also be activated [10, 52, 57, 58]. In the esophageal
AND REGULATION OF TGR5 adenocarcinoma cell line FLO, a coupling of TGR5 with
EXPRESSION AND LOCALIZATION Gα q and Gα i3 upon activation by taurodeoxycholic acid
(TDCA) has been shown [71]. In ciliated and nonciliated
cholangiocytes, different effects of TGR5 agonists on prolif-
TGR5 downstream signaling eration were observed, which have been linked to TGR5
Bile acid binding to TGR5 leads to Gαs‐mediated activation coupling to either Gαs or Gαi [56]. TGR5 thus shows a func-
of the adenylyl cyclase/cyclic AMP (cAMP) signaling path- tional selectivity that may be related to the occurrence of
way [38]. In addition, other cell‐specific signaling pathways different active conformations and to subcellular
24:  TGR5 (GPBAR1) in the Liver 289

Figure 24.2  Important TGR5 ligands. The table shows the substitution pattern for selected natural bile acid agonists with their EC50 values [59].
In addition, a neurosteroid agonist, pregnanediol, an antagonist, SBI‐115, two intestinal agonists, 26a and 15c, and three synthetic agonists, 23g,
18, and INT‐777, are shown [64–68, 70].

localization [56, 72]. Yet, in transfected HEK293 cells or Regulation of TGR5 through di‐ and 
colonocytes, after activation by endogenous agonists, TGR5
interacts neither with GPCR kinases 2, 5, or 6 nor with β‐
oligomerization
arrestins 1 or 2 and, accordingly, shows no trafficking to It is now well established that GPCRs can form homo‐ and het-
endosomes and no desensitization even after repeated stimu- ero‐oligomers, and oligomer formation can affect a broad range
lation [73]. of biological functions [79]. For TGR5, a combined strategy
Ligand binding to TGR5 not only triggers an elevation of using cell biology, multiparameter fluorescence image spectros-
intracellular cAMP but also of intracellular calcium levels, pro- copy (MFIS) for quantitative fluorescence resonance energy
motes activation of different ion channels, regulates gene transfer (FRET) analysis, and integrative modeling was applied
expression, and induces mitochondrial fission [38, 45, 48, 50, to obtain structural information on dimerization and oligomeri-
55, 74–77]. Moreover, stimulation of TGR5 by bile acids acti- zation of TGR5, fused to fluorescent proteins, in live cells [80].
vates a broad range of kinase signaling pathways, including pro- The correlation of FRET parameters with increasing FRET
tein kinase A (PKA), protein kinase B (AKT), epidermal growth acceptor‐to‐donor ratio showed that TGR5 wild‐type forms
factor receptor (EGFR), mammalian target of rapamycin higher-order oligomers, a process disrupted in a TGR5 Y111A
(mTOR), Rho kinase, and extracellular signal‐regulated kinase variant; residue 111 is located in TM3 within the highly con-
(ERK) pathways [38, 52, 55, 56, 76, 78]. served D/ERY motif. From the ratio dependence it was
290 THE LIVER:  FUNCTION OF TGR5 IN DIFFERENT LIVER CELLS

concluded that higher-order oligomers of TGR5 wild‐type are PKA‐mediated serine phosphorylation of endothelial nitric oxide
formed from dimers interacting with other dimers. In contrast, synthase (eNOS), and increased generation and release of nitric
dimers are the largest unit suggested for the TGR5 Y111A vari- oxide (NO) [10, 39, 50]. Furthermore, TGR5 activation triggered
ant. Higher-order oligomers likely have a linear arrangement an AKT‐dependent serine phosphorylation of cystathionine‐γ‐
with interaction sites involving TM1 and helix 8 in the cytoplasm lyase (CSE), an enzyme essential for the generation of the vaso-
as well as TM5 (Figure 24.3c). The latter interaction was sug- dilatory molecule hydrogen sulfide (H2S) [39, 83, 84]. On the
gested to be disrupted by the Y111A mutation. The importance genomic level, ligand binding to TGR5 promoted upregulation
of helix 8 for TGR5 dimerization, together with reports that of eNOS and CSE mRNA levels, while it suppressed expression
GPCR dimerization in the endoplasmic reticulum is obligatory of endothelin 1 (ET‐1) (Figure  24.4) [50, 83–85]. Therefore,
for their membrane trafficking, might explain why a membrane‐ TGR5 activation within LSECs favors the generation and
proximal, C‐terminal helix is structurally required for plasma release of vasodilatory molecules (NO, H2S) while it inhibits the
membrane localization and function of TGR5 [69]. Finally, time‐ expression of ET‐1, thus reducing HSC contractility as well as
series MFIS–FRET analysis before, at the time point of, and portal pressure (Figure 24.4) [39, 83].
after stimulation of TGR5 with taurocholate (TCA, Figure 24.2),
a bile acid less cytotoxic than TLCA in live cells, indicated that
TCA, and subsequent G-protein coupling, does not influence the
TGR5 in hepatic stellate cells
oligomerization state of TGR5 wild‐type and Y111A [80], an TGR5 mRNA and protein expression were negligible in freshly
observation supported by the literature [81]. isolated, quiescent rat HSCs [53]. Immunofluorescent staining
for TGR5 was negative in glial fibrillary acidic protein (GFAP)‐
positive quiescent HSC in rodent livers [48, 53]. However,
Regulation of TGR5 expression and function TGR5 mRNA and protein levels increased over days in cul-
Mechanisms regulating TGR5 expression, localization, and tured, myofibroblast‐like activated HSCs and in damaged livers
function are mostly elusive to date. Reduced TGR5 mRNA and in vivo (Figures 24.1 and 24.4) [39, 48, 53]. Cyclic AMP was
protein levels were observed in cultured rat astrocytes after stim- shown to desensitize the ETA receptor towards ET‐1 in activated
ulation with ammonia (NH4Cl; 0.5–5 mM over 72 hours) [45]. HSCs through receptor internalization [86]. Therefore, TGR5
Moreover, TGR5 mRNA expression was significantly lower in stimulation in activated HSCs may counteract contraction of
cortical brain tissue obtained from patients with hepatic enceph- these cells and ameliorate portal hypertension.
alopathy (HE) as compared to samples from patients without HE
[45]. Incubation of cultured rat astrocytes or human macrophages
with the TGR5 ligands 5β‐pregnan‐3α‐ol‐20‐one (pregnanolone)
TGR5 in Kupffer cells and macrophages
(10 μM, 72 hours), 5α‐pregnan‐3α‐ol‐20‐one (allopregnanolone) TGR5 ligand activation exerts strong anti‐inflammatory effects in
and 5α‐pregnan‐3β‐ol‐20‐one (isopregnanolone) (25 μM, 12 or Kupffer cells as well as bone marrow‐derived macrophages
72 hours), respectively, resulted in a significant downregulation (BMDMs), where activation of the receptor inhibits the expres-
of TGR5 mRNA levels as compared to vehicle‐treated cells [45, sion and secretion of proinflammatory cytokines [48, 87, 88].
46]. This suggests that continuous ligand stimulation and the Activation of TGR5 through elevation of cAMP inhibits phos-
resulting downregulation of TGR5 mRNA expression represents phorylation of IκBα and thus prevents translocation of the nuclear
an important mechanism of receptor desensitization, especially factor (NF)‐κB p65 into the nucleus [39, 89]. Furthermore, via
since TGR5 is not retrieved from the plasma membrane in PKA‐mediated phosphorylation of nucleotide‐binding domain
response to repetitive ligand activation [45, 46, 73]. (NB) and leucine‐rich repeat (LRR)‐containing receptor protein 3
In contrast, an upregulation of TGR5 mRNA levels has been (NLRP3), TGR5 suppresses caspase 1‐dependent maturation of
described in the frontal brain cortex of mice following azoxym- proinflammatory cytokines, such as interleukin 1β (IL‐1β) and
ethane‐induced liver failure [82]. Moreover, an increase in IL‐18 [39, 88]. Ligand activation of TGR5 also inhibits chemokine
TGR5 immunofluorescence staining was observed in Kupffer expression via a PKB (AKT)–mTOR signaling pathway, result-
cells in rat liver following common bile duct ligation (CBDL) ing in the increased expression of the CCAAT/enhancer binding
for 72 hours [48]. The molecular mechanisms regulating TGR5 protein β (C/EBPβ) isoform liver inhibitory protein (LIP) [39, 76,
mRNA expression are yet unknown. 90]. Other anti‐inflammatory mechanisms triggered by TGR5
activation comprise reduced macrophage migration and phagocy-
tosis as well as preferential differentiation of monocytes into an
anti‐inflammatory, regulatory phenotype (Figure  24.5) [76, 89,
FUNCTION OF TGR5 IN DIFFERENT 91–94]. Similar to FXR, TGR5 promotes a variety of anti‐inflam-
LIVER CELLS matory effects [39].

TGR5 in liver sinusoidal endothelial cells TGR5 in the biliary epithelium


Stimulation of TGR5 in cultivated liver sinusoidal endothelial and the gallbladder
cells (LSECs) from rat liver resulted in activation of adeny- TGR5 is expressed in cholangiocytes of the small and large
lyl cyclase, elevation of intracellular cAMP, subsequent intrahepatic ducts as well as the extrahepatic bile duct, where
24:  TGR5 (GPBAR1) in the Liver 291

(a) (c)

(b)

Figure 24.3  Interactions of TGR5 with TLC and TGR5 oligomer formation. (a) Binding mode model of TLC (orange sticks and transparent
surface) in TGR5 (gray, cartoon) [61]. Important interacting residues are shown as sticks. TLC bridges TM3 and TM6 by interacting with Y240 and
Y89, which are important for TGR5 activation. E169 and R79 are important for the efficacy of TLC. (b) Two‐dimensional diagram showing interac-
tions of TLC within TGR5. Hydrogen bonding and salt‐bridge interactions are shown as dashed lines, hydrophobic interactions are shown in green.
(c) Schematic of TGR5 oligomerization [80]. TGR5 forms dimers involving TM1 and helix 8 (1–8). Then, oligomerization can occur from dimers
either via the TM4–TM5 interface (4–5) or the TM5–TM6 interface (5–6).
292 THE LIVER:  FUNCTION OF TGR5 IN DIFFERENT LIVER CELLS

the receptor is localized in the apical plasma membrane and the molecule A (JAM‐A), thereby reducing paracellular permeabil-
primary cilium [39, 43, 47–49, 51, 52, 56–58]. Ligand binding ity and protecting liver tissue from bile leakage (Figure  24.6)
to TGR5 activates adenylyl cyclase, triggers cAMP elevation, [98]. TGR5 can exert opposing effects on cholangiocyte prolif-
and subsequent stimulation of the cystic fibrosis transmembrane eration dependent on the receptor’s subcellular localization at
conductance regulator (CFTR, ABCC7) resulting in increased time of activation [39, 47]. Ligand binding to TGR5 on the pri-
chloride secretion (Figure  24.6) [39, 49, 54, 57]. Chloride is mary cilia of simian virus 40‐transformed human cholangio-
subsequently exchanged against bicarbonate via the anion cytes (H69 cells) leads to coupling to an inhibitory Gαi protein,
exchanger 2 (AE2, SLC4A2), leading to bicarbonate‐rich reduces intracellular cAMP concentrations and impairs cell pro-
choleresis and formation of a protective bicarbonate layer liferation [39, 47, 56]. Stimulation of TGR5 in non‐ciliated H69
known as the bicarbonate umbrella [39, 47, 49, 54, 67, 95–97]. cells promotes cell proliferation through elevation of cAMP [47,
Cyclic AMP not only enhances transport activity but also trig- 56]. In contrast, incubation of murine cholangiocytes with bile
gers the insertion of CFTR and AE2 into the apical plasma acids or synthetic TGR5 agonists results in generation of reac-
membrane from intracellular vesicles, resulting in increased bil- tive oxygen species (ROS), stimulation of Src kinase and matrix
iary secretion [47, 49, 54, 58]. Absence of TGR5 renders chol- metalloproteinases, transactivation of the EGFR, phosphoryla-
angiocytes more susceptible towards bile acid‐induced toxicity tion of ERK1/2, and significantly increased cell proliferation
[43, 47, 52, 96]. Activation of TGR5 also triggers increased [39, 43, 47, 52]. This process is independent of adenylyl cyclase
expression and phosphorylation of the junctional adhesion activation [52]. In contrast, TGR5 attenuates CD95 ligand‐
induced apoptosis via stimulation of adenylyl cyclase, elevation
of cAMP and PKA activation (Figure 24.6) [47, 52].
In the gallbladder epithelium, TGR5 activation also enhances
CFTR‐mediated chloride secretion, while in gallbladder smooth
muscle cells TGR5 stimulation through a cAMP–PKA pathway
results in the opening of ATP‐sensitive potassium channels
(KATP) and subsequent muscle cell relaxation [54, 55]. Studies
with full‐thickness gallbladder tissue preparations and exposure
of the epithelial cell layer to the bile acids also resulted in
smooth muscle cell relaxation [55]. Thus, absorption of bile
acids through the epithelium and subsequent TGR5 activation

Figure 24.4  Function of TGR5 in liver sinusoidal endothelial cells


(LSECs) and activated hepatic stellate cells (HSCs). Ligand binding to
TGR5 stimulates adenylyl cyclase and the generation of cyclic AMP
(cAMP), which in turn triggers activation of further downstream targets Figure 24.5  TGR5 exerts anti‐inflammatory effects in Kupffer cells
comprising cAMP response element binding protein (CREB) and pro- and macrophages. TGR5‐dependent elevation of cAMP impairs IκB
tein kinase A (PKA). CREB promotes increased expression of endothe- kinase activity and phosphorylation of IκB, thereby stabilizing IκB as
lial nitric oxide synthase (eNOS) and cystathionine‐γ‐lyase (CSE), well as NF‐κB p65 within the cytoplasm. This results in decreased tran-
resulting in increased generation of NO and hydrogen sulfide (H2S). scriptional activity of NF‐κB and, thus, reduced expression of proin-
This is further enhanced through PKA‐ and AKT‐mediated activation flammatory cytokines. TGR5 activation impairs caspase 1‐dependent
of eNOS and CSE. Additionally, TGR5 suppresses expression of cleavage of proinflammatory cytokines through PKA‐dependent phos-
endothelin‐1 (ET‐1). In activated HSCs, TGR5 inhibits ET‐1 signaling phorylation of NLRP3. Moreover, TGR5 activation reduces chemokine
and, thus, contraction through cAMP‐dependent internalization of the expression and secretion via an AKT–mTOR–LIP signaling pathway.
endothelin receptor. In summary, TGR5 activation promotes secretion In summary, TGR5 stimulation attenuates expression of proinflamma-
of vasodilators and inhibits release of vasoconstrictors, leading to tory cytokines and chemokines and stabilizes or increases the transcrip-
reduced portal pressure. Modified from [39] with permission of Georg tion of anti‐inflammatory molecules. Modified from [39] with
Thieme Verlag KG Stuttgart. permission of Georg Thieme Verlag KG Stuttgart.
24:  TGR5 (GPBAR1) in the Liver 293

Role of TGR5 in gallbladder and biliary


diseases: lessons from mouse models
As described above, TGR5‐knockout mice not only have
reduced bile flow but also a very small gallbladder volume,
which is explained by impaired smooth muscle cell relaxation in
absence of TGR5 [43, 55, 95]. Oral administration of TGR5
ligands to wild‐type mice (6α‐ethyl‐23(S)‐methyl‐cholic acid
(INT‐777, EMCA, EC50 = 0.82 μM), or the highly potent small
molecules compound 18 (EC50 = 25 nM) or compound 23g
(EC50 = 0.72 nM)) (Figure 24.2) induced biliary bile flow and
promoted gallbladder filling with an increase in gallbladder size
to up to 231% (compound 23g, Figure 24.2) [43, 66, 67, 95].
The latter compound (23g), which is a 4‐phenoxy‐nicotinamide
derivative and thus a non‐bile acid TGR5 ligand, accumulated in
plasma, bile, and gallbladder tissue in mice, underscoring a
direct, local effect of this compound on TGR5 in gallbladder
smooth muscle cells [67]. Intestinal‐specific TGR5 ligands with
low systemic availability (such as compound 15c, Figure 24.2)
Figure 24.6  Function of TGR5 in biliary epithelial cells. Ligand bind- failed to increase gallbladder volume significantly [65]. This
ing to TGR5 triggers activation of a stimulatory G protein and adenylyl
suggests that TGR5 ligands need to accumulate in gallbladder
cyclase (AC). Through elevation of intracellular cyclic AMP (cAMP)
TGR5 promotes chloride and subsequent bicarbonate excretion via tissue either through absorption from the gallbladder lumen or
CFTR and the anion exchanger 2 (AE2), resulting in the formation of the through accumulation in plasma in order to facilitate gallbladder
bicarbonate umbrella. Furthermore, cAMP‐dependent activation of pro- filling. Whether systemic application of a TGR5 ligand, which
tein kinase A (PKA) leads to serine/threonine phosphorylation of the is not secreted into bile, is sufficient to trigger gallbladder
CD95 receptor and inhibition of apoptosis. Stimulation of TGR5 induces smooth muscle cell relaxation needs to be determined.
expression and phosphorylation of the junctional adhesion molecule A Interestingly, mice with targeted deletion of TGR5 are pro-
(JAM‐A) and regulates tight junction (TJ) integrity and paracellular per-
tected from cholesterol gallstone formation when administered
meability. Ligand binding to TGR5 can increase the levels of reactive
oxygen species (ROS) independent of AC activation, resulting in Src a lithogenic diet [42, 43]. Absence of TGR5 from gallbladder
kinase activation, matrix metalloproteinase (MMP)‐dependent shedding smooth muscle cells may prevent bile acid‐dependent reduction
of the epidermal growth factor (EGF), transactivation of the EGF receptor in contractility, while overexpression of TGR5, as observed in
(EGFR), followed by phosphorylation of MAP kinase ERK1/2 and human tissue from patients with gallstone disease, may promote
increased cell proliferation, while activation of TGR5 in primary cilia gallbladder hypomotility, a prerequisite for gallstone formation
inhibited cell proliferation. Modified from [47, 49, 52]. [54, 55]. Whether this mechanism plays a role in human gall-
stone disease needs further investigation. However, monitoring
of gallbladder function is warranted if TGR5 agonists are evalu-
on gallbladder smooth muscle cells is a rapid and local mecha- ated for clinical applications [39].
nism for gallbladder filling [55, 66]. Moreover, bile acids medi- Targeted deletion of TGR5 renders mice more susceptible
ate gallbladder filling not only through TGR5 but also through towards different types of cholestatic liver injury [39, 43, 47, 52].
the ileal FXR downstream effector fibroblast growth factor 19 TGR5‐knockout mice display increased liver damage as deter-
(FGF19, which is FGF15 in mice) [55, 95, 99]. mined by elevated liver enzymes or by histology in response to
cholic acid feeding (CA, 0.5% for 7 days or 1% for 5 days) or
common bile duct ligation (CBDL) for up to 7 days [39, 43, 47,
52, 63]. Gene deletion of TGR5 impaired cholangiocyte and
INSIGHTS FROM TGR5‐KNOCKOUT hepatocyte proliferation in response to cholestasis in vivo [47, 52,
MICE AND APPLICATION OF TGR5 63]. In isolated, cultured murine cholangiocytes, TGR5 was
LIGANDS TO RODENTS essential for bile acid‐induced proliferation, which was mediated
by a TGR5–ROS–Src–MMP–EGFR–ERK1/2 signaling cascade
Targeted deletion of TGR5 in mice is not associated with an (Figure  24.6) [52]. In contrast, the mechanisms responsible for
obvious phenotype or development of spontaneous liver disease reduced hepatocyte proliferation in TGR5‐knockout mice are as
[42, 44]. However, TGR5‐knockout mice have a smaller bile yet unknown [43]. However, impaired hepatocyte proliferation
acid pool size with reduced levels of the FXR‐antagonistic bile was also detected after partial hepatectomy (PHx) in TGR5‐
acid TβMCA and an increase in the fraction of TCA and TDCA, knockout mice and may be attributed in part to higher hepatic bile
which may be attributed to lower expression levels of Cyp7b1 acid concentrations as well as a more hydrophobic bile acid pool
and suppression of bile acid synthesis via the alternative path- composition in the absence of TGR5 [43, 63]. Liver injury after
way [43, 95, 100, 101]. This further supports an overlap of PHx could be attenuated in TGR5-knockout through bile acid
TGR5 and FXR functions, since reduced concentrations of depletion with cholestyramine (2%), underscoring the hypothesis
TβMCA may disinhibit FXR signaling in the absence of TGR5 that elevated bile acid levels as well as altered composition of the
[43, 101]. bile acid pool underlie the observed phenotype [43, 63].
294 THE LIVER:  HUMAN LIVER DISEASE

Role of TGR5 in inflammatory and metabolic Role of TGR5 in progressive liver disease:


liver disease: lessons from mouse models lessons from rodent models
TGR5 is an important negative regulator of inflammation both Portal hypertension represents a common complication of
systemically but also locally in the liver as described earlier. advanced liver fibrosis and liver cirrhosis and is associated with
TGR5‐deficient mice were more susceptible towards intraperi- anatomical and biochemical changes in both LSECs and HSCs
toneal injection of lipopolysaccharide resulting in elevated [39, 108, 109]. Loss of fenestration, increased deposition of
serum levels of alanine (ALT) and aspartate (AST) aminotrans- extracellular matrix in the space of Disse, reduced generation
ferases and inflammatory cytokines, such as IL‐1β, compared to and secretion of vasodilatory mediators as well as increased
their wild‐type littermates [39, 43, 87]. Increased inflammatory expression and release of profibrotic molecules, such as trans-
infiltrates as well as enhanced hepatocyte apoptosis were pre- forming growth factor β (TGF‐β) and platelet‐derived growth
sent in the livers of TGR5‐knockout mice as compared to wild‐ factor‐C (PDGF‐C), are characteristics of LSEC dysfunction in
type animals [39, 43, 87]. Simultaneous intraperitoneal injection portal hypertension [39, 109]. HSC activation and transforma-
of both LPS and TLCA or LPS and the synthetic TGR5 ligand tion into myofibroblast‐like cells is typically associated with
INT‐777 attenuated LPS‐mediated elevation of IL‐1β and IL‐18, loss of vitamin A granules, production of extracellular matrix
an effect which was abrogated in mice deficient for TGR5 or the and increased expression of alpha‐smooth muscle actin (α‐
NLRP3 inflammasome [39, 88]. SMA) and also of TGR5 [39, 53, 109]. TGR5 modulates portal
Activation of TGR5 has been shown to improve various pressure and hepatic microcirculation under physiological con-
aspects of the metabolic syndrome in different animal models ditions [39, 50]. TGR5 activation by a synthetic agonist BAR501
[11, 43, 65–67, 74, 89, 102]. Therefore, TGR5 has emerged as in mice with hepatic fibrosis induced by carbon tetrachloride
an attractive drug target for metabolic diseases including obe- administration over nine weeks improved portal hypertension
sity, insulin resistance, atherosclerosis, and non‐alcoholic fatty significantly [43, 85]. Reduced expression of ET‐1 and increased
liver disease (NAFLD) leading to the development of different expression and activity of CSE, which is essential for the gen-
synthetic TGR5 agonists [11, 43, 65–68, 90, 103, 104]. eration of the vasodilatory molecule hydrogen sulfide, underlie
Steatohepatitis is characterized by inflammation and infiltra- this beneficial effect [39, 85]. Furthermore, cAMP‐mediated
tion of Ly‐6Chigh monocytes, and the differentiation of mac- internalization of the ETA receptor leads to reduced responsive-
rophages towards an inflammatory phenotype is a typical ness of activated HSCs towards ET‐1 and may also contribute to
finding in progressive disease [39, 105]. Application of a dual the reduction in portal pressure [86]. Despite the improvement
FXR and TGR5 agonist (6α‐ethyl‐24‐nor‐5β‐cholane‐3α,7α,23‐ in portal hypertension, BAR501 did not reverse the increased
trio‐23‐sulfate sodium salt, INT‐767) to obese db/db mice for expression levels of profibrogenic genes, such as TGF-β1,
six weeks reduced levels of Ly‐6Chigh and increased levels of collagen1α1, and α‐SMA [85]. Whether TGR5 agonists can
Ly‐6Clow monocytes, resulting in a significant improvement of improve liver fibrosis or even cirrhosis in other models of liver
hepatic inflammation, steatosis, and hepatocyte ballooning as damage is unknown.
well as profibrotic gene expression [39, 106]. Stimulation of
TGR5 with the synthetic agonist INT‐777 in mice on a high‐fat
diet reduced fat mass and obesity and also triggered intestinal
secretion of glucagon‐like peptide 1 (GLP‐1) thereby enhancing HUMAN LIVER DISEASE
glucose tolerance [43, 74, 75]. The reduction of body weight
and the GLP‐1‐mediated improvement of glucose tolerance
were also observed with other systemically available non‐bile
Role of TGR5 in biliary diseases
acid TGR5 agonists [66, 67, 103]. Furthermore, INT‐777 Genome‐wide association studies suggested an association of
administration attenuated hepatic steatosis through reduction of the TGR5 gene locus with ulcerative colitis and primary scleros-
hepatic fatty acid and triglyceride concentrations [43, 74]. ing cholangitis (PSC) [110, 111]. Resequencing of TGR5 in
Inhibition of hepatic and adipose tissue inflammation improved patients with PSC and healthy blood donors failed to support a
insulin sensitivity, enhanced energy expenditure in brown adi- role for TGR5 coding sequence (CDS) variants in the disease
pose tissue and skeletal muscle, as well as increased lipolysis, pathogenesis [41, 47, 49, 112]. Nonsynonymous variants in the
mitochondrial biogenesis, and mitochondrial fission. Induction TGR5 CDS were rare and variants on both alleles were not
of beiging and thermogenesis in white adipose tissue may thus ­identified in any of the patients or controls [41, 47, 49, 112].
underlie the beneficial effects of TGR5 agonist treatment on However, a common polymorphism rs11554825 was identified
obesity and steatohepatitis [74–76, 90, 102]. within the noncoding first exon of TGR5 and in almost com-
Interestingly, FXR and TGR5 ligands exert overlapping and plete linkage disequilibrium with a second polymorphism
complementary beneficial effects on various aspects of the meta- rs3731859 within the 5′ untranslated region (UTR), which may
bolic syndrome including NAFLD. While FXR agonists regulate affect TGR5 promotor activity and, thus, TGR5 mRNA expres-
lipid and glucose homeostasis in the liver directly as well as indi- sion levels [41, 47, 49]. TGR5 mRNA levels were lowest in
rectly through intestinal induction and secretion of FGF15/19, ­individuals homozygous for the G‐allele at rs3731859, which
TGR5 agonists promote anti‐inflammatory effects in the liver and was linked to the presence of the C‐allele at rs11554825 [41].
adipose tissue, induce energy expenditure as well as lipolysis and Furthermore, presence of the risk allele at rs11554825 (C‐allele)
thermogenesis in skeletal muscle and white adipose tissue, respec- was significantly associated to PSC across four cohorts (odds
tively, and trigger GLP‐1 secretion from the intestine [75, 107]. ratio 1.14) [39, 41]. Reduced TGR5 mRNA levels impair
24:  TGR5 (GPBAR1) in the Liver 295

TGR5‐mediated cytoprotective effects and render cholangio- SUMMARY AND PERSPECTIVE


cytes more susceptible towards bile acid‐induced cell damage,
thereby contributing to the development or progression of bil- Bile acid effects in liver are mediated through different bile
iary diseases such as PSC and PBC [39, 47, 52]. Interestingly, acid receptors, including both NRs and GPCRs. TGR5
cholangiocytes in livers of patients with PSC showed a dimin- (GPBAR1) is localized in different nonparenchymal liver cells,
ished immunofluorescence staining intensity as compared to where the receptor modulates liver microcirculation, inflam-
control livers independent of the common TGR5 rs11554825 matory response, and biliary function. In humans, a role for
polymorphism (Figure 24.1  l) [39, 47, 49]. Furthermore, a simi- TGR5 has been suggested in biliary diseases such as PBC and
lar reduction of TGR5 fluorescence staining was observed in PSC, where reduced levels of TGR5 may accelerate disease
cholangiocytes in livers of Mdr2 (Abcb4)‐knockout mice, which progression. Furthermore, an overexpression of the receptor
serve as a model for PSC [39, 43, 47, 49, 113]. The mechanisms has been reported in malignant and cystic cholangiocytes in
and timing underlying TGR5 downregulation in PSC are cur- human cholangiocarcinoma and polycystic liver disease,
rently under investigation. respectively. In both diseases, activation of the receptor pro-
In contrast to sclerosing cholangitis, an overexpression of motes disease progression and, hence, disruption of TGR5
TGR5 mRNA has been observed in intrahepatic cholangiocarci- signaling may prove beneficial in these conditions. Mice with
noma (iCCA) as well as in extrahepatic, perihilar CCA com- targeted deletion of TGR5 develop more severe liver damage
pared to tissue from the non‐malignant resection margins [47, with increased parenchymal and biliary injury, reduced adap-
52, 114]. On the cellular level, an approximately threefold tive hepatocyte and cholangiocyte proliferation as well as
increase of TGR5 immunofluorescence staining intensity was increased immune activation in response to bile acid feeding,
measured in iCCA cells compared to cholangiocytes from the partial hepatectomy, common bile duct ligation, or treatment
non‐malignant surrounding tissue (Figure  24.1m), while the with lipopolysaccharide. Furthermore, TGR5 activation
staining intensity of cytokeratin 19 was unchanged [39, 47, 52]. improves many aspects of the metabolic syndrome, including
Incubation of CCA‐derived cell lines (EGI‐1 and TFK‐1) with steatohepatitis, glucose homeostasis, adipose tissue inflamma-
different TGR5 agonists (TLCA or synthetic ligands) triggered tion, atherosclerosis, and kidney injury. Administration of
transactivation of the EGFR as well as ERK1/2 phosphoryla- TGR5 agonists also ameliorates obesity through increased
tion, leading to increased cell proliferation, which was attenu- energy expenditure in brown adipose tissue and skeletal muscle
ated by siRNA knockdown of TGR5 [39, 47, 52]. Stimulation of as well as through enhanced lipolysis and mitochondrial bio-
TGR5 also increased migration and invasiveness of CCA cell genesis in white adipose tissue. Despite these beneficial effects,
lines, which again was abolished by targeted deletion of TGR5 TGR5 ligands promote gallbladder filling via gallbladder
[47, 52, 114]. smooth muscle relaxation, which may lead to abdominal dis-
An increased expression of TGR5 has also been detected in comfort, gallstone disease, and its potential complications, all
cystic cholangiocytes in polycystic liver disease (PLD) [39, 43, of which may limit the use of systemically available TGR5
47, 70, 113]. Activation of TGR5 in cholangiocytes triggered agonists. Intestinal‐specific TGR5 ligands promote GLP‐1
cell proliferation and cyst growth through elevation of cAMP, a secretion, improve glucose homeostasis, and are devoid of the
process which could be abolished in vitro by application of a unwanted side‐effects on gallbladder function; however, many
novel TGR5 inhibitor (m‐tolyl 5‐chloro‐2‐(ethylsulfonyl)‐ of the beneficial effects of targeting systemic TGR5 will not be
pyrimidine‐4‐carboxylate; SBI‐115, Figure 24.2) or attenuated induced by these molecules. Further studies are needed to
in vivo by targeted deletion of TGR5 [39, 70]. Thus, in PLD and investigate the potential of targeting intestinal TGR5 and its
CCA inhibition of TGR5 signaling may offer novel therapeutic impact on gut liver axis as well as on the microbiome host
options in these difficult to treat disorders [39, 43, 47, 70]. interactions.

Role of TGR5 in extrahepatic complications


of chronic liver disease: hepatic ACKNOWLEDGMENTS
encephalopathy Our studies are supported by Deutsche Forschungsgemeinschaft
TGR5 is expressed in virtually all cell types in the central nerv- (DFG) through SFB 974 and KFO 217.
ous system and can even be recovered in synapses [45, 115].
Here, TGR5 acts as a neurosteroid receptor and its stimulation
triggers activation of adenylate cyclase, leads to elevation of
intracellular calcium levels, and induces the formation of ROS
REFERENCES
[45]. In hepatic encephalopathy, which is seen as the clinical 1. Li, T. and Chiang, J.Y. Bile acid signaling in metabolic disease and drug ther-
manifestation of a low‐grade cerebral edema with oxidative apy. Pharmacol Rev, 2014;66(4):948–83.
stress [116], TGR5 is strongly downregulated, which may 2. Russell, D.W. The enzymes, regulation, and genetics of bile acid synthesis.
reflect a compensatory response with regard to oxidative stress Annu Rev Biochem, 2003;72:137–74.
[45]. In microglial cells, stimulation of TGR5 with pregna- 3. Hofmann, A.F. Bile acids: the good, the bad, and the ugly. News Physiol Sci,
1999;14:24–9.
nolone inhibits expression of proinflammatory cytokines and 4. Joyce, S.A. and Gahan, C.G. Bile acid modifications at the microbe‐host
thus exerts anti‐inflammatory and neuroprotective effects in interface: potential for nutraceutical and pharmaceutical interventions in host
hepatic encephalopathy [82, 117]. health. Annu Rev Food Sci Technol, 2016;7:313–33.
296 THE LIVER:  REFERENCES

5. Wahlstrom, A., Kovatcheva‐Datchary, P., Stahlman, M., Backhed, F., and 30. Chen, X., Yang, D., Shen, W. et  al. Characterization of chenodeoxycholic
Marschall, H.U. Crosstalk between bile acids and gut microbiota and its acid as an endogenous antagonist of the G‐coupled formyl peptide receptors.
impact on farnesoid X receptor signalling. Dig Dis, 2017;35(3):246–50. Inflamm Res, 2000;49(12):744–55.
6. Kullak‐Ublick, G., Stieger, B., and Meier, P.J. Enterohepatic bile salt trans- 31. Ferrari, C., Macchiarulo, A., Costantino, G., and Pellicciari, R.
porters in normal physiology and liver disease. Gastroenterology, Pharmacophore model for bile acids recognition by the FPR receptor. J
2004;126(1):322–42. Comput Aided Mol Des, 2006;20(5):295–303.
7. Häussinger, D., Reinehr, R., and Keitel, V. Bile acid signaling in the liver and 32. Chiang, J.Y. Sphingosine‐1‐phosphate receptor 2: a novel bile acid receptor
the biliary tree, in Hepatobiliary Transport in Health and Disease (eds. D. and regulator of hepatic lipid metabolism? Hepatology,
Häussinger, V. Keitel, and R Kubitz), DeGruyter Publishing, Berlin, 2012, 2015;61(4):1118–20.
pp. 85–102. 33. Kwong, E., Li, Y., Hylemon, P.B., and Zhou, H. Bile acids and sphingo-
8. Copple, B.L. and Li, T. Pharmacology of bile acid receptors: evolution of sine‐1‐phosphate receptor 2 in hepatic lipid metabolism. Acta Pharm Sin B,
bile acids from simple detergents to complex signaling molecules. Pharmacol 2015;5(2):151–7.
Res, 2016;104:9–21. 34. Liu, R., Li, X., Qiang, X. et  al. Taurocholate induces cyclooxygenase‐2
9. Godoy, P., Hewitt, N.J., Albrecht, U. et al. Recent advances in 2D and 3D in expression via the sphingosine 1‐phosphate receptor 2 in a human cholangio-
vitro systems using primary hepatocytes, alternative hepatocyte sources and carcinoma cell line. J Biol Chem, 2015;290(52):30988–1002.
non‐parenchymal liver cells and their use in investigating mechanisms of hepa- 35. Liu, R., Zhao, R., Zhou, X. et al. Conjugated bile acids promote cholangio-
totoxicity, cell signaling and ADME. Arch Toxicol, 2013;87(8): 1315–530. carcinoma cell invasive growth through activation of sphingosine 1‐phos-
10. Keitel, V. and Häussinger, D. Perspective: TGR5 (Gpbar‐1) in liver physiol- phate receptor 2. Hepatology, 2014;60(3):908–18.
ogy and disease. Clin Res Hepatol Gastroenterol, 2012;36(5):412–19. 36. Nagahashi, M., Takabe, K., Liu, R. et al. Conjugated bile acid‐activated S1P
11. Pols, T.W., Noriega, L.G., Nomura, M., Auwerx, J., and Schoonjans, K. The receptor 2 is a key regulator of sphingosine kinase 2 and hepatic gene expres-
bile acid membrane receptor TGR5 as an emerging target in metabolism and sion. Hepatology, 2015;61(4):1216–26.
inflammation. J Hepatol, 2011;54(6):1263–72. 37. Studer, E., Zhou, X., Zhao, R. et al. Conjugated bile acids activate the sphin-
12. Makishima, M., Okamoto, A.Y., Repa, J.J. et al. Identification of a nuclear gosine‐1‐phosphate receptor 2 in primary rodent hepatocytes. Hepatology,
receptor for bile acids. Science, 1999;284(5418):1362–5. 2012;55(1):267–76.
13. Parks, D.J., Blanchard, S.G., Bledsoe, R.K. et al. Bile acids: natural ligands 38. Kawamata, Y., Fujii, R., Hosoya, M. et  al. A G protein‐coupled receptor
for an orphan nuclear receptor. Science, 1999;284(5418):1365–8. responsive to bile acids. J Biol Chem, 2003;278(11):9435–40.
14. Wang, H., Chen, J., Hollister, K., Sowers, L.C., and Forman, B.M. 39. Keitel, V. and Häussinger, D. Role of TGR5 (GPBAR1) in liver disease.
Endogenous bile acids are ligands for the nuclear receptor FXR/BAR. Mol Semin Liver Dis, 2018;38(4):333–9.
Cell, 1999;3(5):543–53. 40. Maruyama, T., Miyamoto, Y., Nakamura, T. et  al. Identification of mem-
15. Sinal, C.J., Tohkin, M., Miyata, M. et al. Targeted disruption of the nuclear brane‐type receptor for bile acids (M‐BAR). Biochem Biophys Res Commun,
receptor FXR/BAR impairs bile acid and lipid homeostasis. Cell, 2002;298(5):714–19.
2000;102(6):731–44. 41. Hov, J.R., Keitel, V., Laerdahl, J.K. et al. Mutational characterization of the
16. Watanabe, M., Houten, S.M., Wang, L. et al. Bile acids lower triglyceride bile acid receptor TGR5 in primary sclerosing cholangitis. PloS One
levels via a pathway involving FXR, SHP, and SREBP‐1c. J Clin Invest, 2010;5(8):e12403.
2004;113(10):1408–18. 42. Vassileva, G., Golovko, A., Markowitz, L. et al. Targeted deletion of Gpbar1
17. Kalaany, N.Y. and Mangelsdorf, D.J., LXRS and FXR: the yin and yang of protects mice from cholesterol gallstone formation. Biochem J,
cholesterol and fat metabolism. Annu Rev Physiol, 2006;68:159–91. 2006;398(3):423–30.
18. Modica, S., Gadaleta, R.M., and Moschetta, A. Deciphering the nuclear bile 43. Reich, M., Klindt, C., Deutschmann, K. et al. Role of the G protein‐coupled
acid receptor FXR paradigm. Nucl Recept Signal, 2010;8:e005. bile acid receptor TGR5 in liver damage. Dig Dis, 2017;35(3):235–40.
19. Inagaki, T., Choi, M., Moschetta, A. et al. Fibroblast growth factor 15 func- 44. Maruyama, T., Tanaka, K., Suzuki, J. et al. Targeted disruption of G protein‐
tions as an enterohepatic signal to regulate bile acid homeostasis. Cell coupled bile acid receptor 1 (Gpbar1/M‐Bar) in mice. J Endocrinol,
Metab, 2005;2(4):217–25. 2006;191(1):197–205.
20. Kir, S., Beddow, S.A., Samuel, V.T. et al. FGF19 as a postprandial, insulin‐ 45. Keitel, V., Görg, B., Bidmon, H.J. et  al. The bile acid receptor TGR5
independent activator of hepatic protein and glycogen synthesis. Science, (Gpbar‐1) acts as a neurosteroid receptor in brain. Glia, 2010;58(15):
2011;331(6024):1621–4. 1794–805.
21. Neuschwander‐Tetri, B.A., Loomba, R., Sanyal, A.J. et  al. Farnesoid X 46. Keitel, V., Spomer, L., Marin, J.J. et al. Effect of maternal cholestasis on TGR5
nuclear receptor ligand obeticholic acid for non‐cirrhotic, non‐alcoholic expression in human and rat placenta at term. Placenta, 2013;34(9):810–16.
steatohepatitis (FLINT): a multicentre, randomised, placebo‐controlled trial. 47. Deutschmann, K., Reich, M., Klindt, C. et al. Bile acid receptors in the biliary
Lancet, 2015;385(9972):956–65. tree: TGR5 in physiology and disease. Biochim Biophys Acta, 2018;1864(4 Pt
22. Nevens, F., Andreone, P., Mazzella, G. et al. A placebo‐controlled trial of obet- B):1319–25.
icholic acid in primary biliary cholangitis. N Engl J Med, 2016;375(7):631–43. 48. Keitel, V., Donner, M., Winandy, S., Kubitz, R., and Häussinger, D.,
23. Makishima, M., Lu, T.T., Xie, W. et al. Vitamin D receptor as an intestinal Expression and function of the bile acid receptor TGR5 in Kupffer cells.
bile acid sensor. Science, 2002;296(5571):1313–16. Biochem Biophys Res Commun, 2008;372(1):78–84.
24. Staudinger, J.L., Goodwin, B., Jones, S.A. et al. The nuclear receptor PXR is 49. Keitel, V., Reich, M., and Häussinger, D. TGR5: pathogenetic role and/or
a lithocholic acid sensor that protects against liver toxicity. Proc Natl Acad therapeutic target in fibrosing cholangitis? Clin Rev Allergy
Sci U S A, 2001;98(6):3369–74. Immunol,2015;48(2–3):218–25.
25. Xie, W., Radominska‐Pandya, A., Shi, Y. et al. An essential role for nuclear 50. Keitel, V., Reinehr, R., Gatsios, P. et  al. The G‐protein coupled bile salt
receptors SXR/PXR in detoxification of cholestatic bile acids. Proc Natl receptor TGR5 is expressed in liver sinusoidal endothelial cells. Hepatology,
Acad Sci U S A, 2001;98(6):3375–80. 2007;45(3):695–704.
26. Cheng, K., Chen, Y., Zimniak, P. et al. Functional interaction of lithocholic 51. Keitel, V., Ullmer, C., and Häussinger, D., The membrane‐bound bile acid
acid conjugates with M3 muscarinic receptors on a human colon cancer cell receptor TGR5 (Gpbar‐1) is localized in the primary cilium of cholangio-
line. Biochim Biophys Acta, 2002;1588(1):48–55. cytes. Biol Chem, 2010;391(7):785–9.
27. Raufman, J.P., Chen, Y., Cheng, K. et al. Selective interaction of bile acids 52. Reich, M., Deutschmann, K., Sommerfeld, A. et al. TGR5 is essential for
with muscarinic receptors: a case of molecular mimicry. Eur J Pharmacol, bile acid‐dependent cholangiocyte proliferation in vivo and in vitro. Gut
2002;457(2–3):77–84. 2016;65(3):487–501.
28. Raufman, J.P., Chen, Y., Zimniak, P., and Cheng, K. Deoxycholic acid con- 53. Sawitza, I., Kordes, C., Götze, S., Herebian, D., and Häussinger, D., Bile
jugates are muscarinic cholinergic receptor antagonists. Pharmacology, acids induce hepatic differentiation of mesenchymal stem cells. Sci Rep,
2002;65(4):215–21. 2015;5:13320.
29. Sheikh Abdul Kadir, S.H., Miragoli, M., Abu‐Hayyeh, S. et  al. Bile acid‐ 54. Keitel, V., Cupisti, K., Ullmer, C. et  al. The membrane‐bound bile acid
induced arrhythmia is mediated by muscarinic M2 receptors in neonatal rat receptor TGR5 is localized in the epithelium of human gallbladders.
cardiomyocytes. PloS One, 2010;5(3):e9689. Hepatology, 2009;50(3):861–70.
24:  TGR5 (GPBAR1) in the Liver 297

55. Lavoie, B., Balemba, O.B., Godfrey, C. et al. Hydrophobic bile salts inhibit 78. Rajagopal, S., Kumar, D.P., Mahavadi, S. et al. Activation of G protein‐
gallbladder smooth muscle function via stimulation of GPBAR1 receptors coupled bile acid receptor, TGR5, induces smooth muscle relaxation via
and activation of KATP channels. J Physiol, 2010;588(Pt 17):3295–305. both Epac‐ and PKA‐mediated inhibition of RhoA/Rho kinase pathway.
56. Masyuk, A.I., Huang, B.Q., Radtke, B.N. et al. Ciliary subcellular localiza- Am J Physiol Gastrointest Liver Physiol, 2013;304(5):G527–35.
tion of TGR5 determines the cholangiocyte functional response to bile acid 79. Hiller, C., Kuhhorn, J., and Gmeiner, P., Class A G‐protein‐coupled receptor
signaling. Am J Physiol Gastrointest Liver Physiol, 2013;304(11): (GPCR) dimers and bivalent ligands. J Med Chem, 2013;56(17):6542–59.
G1013–24. 80. Greife, A., Felekyan, S., Ma, Q. et al. Structural assemblies of the di‐ and
57. Keitel, V. and Häussinger, D. TGR5 in the biliary tree. Dig Dis, oligomeric G‐protein coupled receptor TGR5 in live cells: an MFIS‐FRET
2011;29(1):45–7. and integrative modelling study. Sci Rep, 2016;6:36792.
58. Keitel, V. and Häussinger, D. TGR5 in cholangiocytes. Curr Opin 81. Patowary, S., Alvarez‐Curto, E., Xu, T.R. et al. The muscarinic M3 acetyl-
Gastroenterol, 2013;29(3):299–304. choline receptor exists as two differently sized complexes at the plasma
59. Sato, H., Macchiarulo, A., Thomas, C. et al. Novel potent and selective bile membrane. Biochem J, 2013;452(2):303–12.
acid derivatives as TGR5 agonists: biological screening, structure‐activity 82. McMillin, M., Frampton, G., Tobin, R. et al. TGR5 signaling reduces
relationships, and molecular modeling studies. J Med Chem, neuroinflammation during hepatic encephalopathy. J Neurochem,
2008;51(6):1831–41. 2015;135(3):565–76.
60. Martin, R.E., Bissantz, C., Gavelle, O. et al. 2‐Phenoxy‐nicotinamides are 83. Fiorucci, S., Zampella, A., Cirino, G., Bucci, M., and Distrutti, E., Decoding
potent agonists at the bile acid receptor GPBAR1 (TGR5). ChemMedChem, the vasoregulatory activities of bile acid‐activated receptors in systemic and
2013;8(4):569–76. portal circulation: role of gaseous mediators. Am J Physiol Heart Circ
61. Gertzen, C.G., Spomer, L., Smits, S.H. et  al. Mutational mapping of the Physiol, 2017;312(1):H21–H32.
transmembrane binding site of the G‐protein coupled receptor TGR5 and 84. Renga, B., Bucci, M., Cipriani, S. et  al. Cystathionine gamma‐lyase, a
binding mode prediction of TGR5 agonists. Eur J Med Chem, H2S‐generating enzyme, is a GPBAR1‐regulated gene and contributes to
2015;104:57–72. vasodilation caused by secondary bile acids. Am J Physiol Heart Circ
62. Katritch, V. and Abagyan, R. GPCR agonist binding revealed by modeling Physiol, 2015;309(1):H114–26.
and crystallography. Trends Pharmacol Sci, 2011;32(11):637–43. 85. Renga, B., Cipriani, S., Carino, A. et al. Reversal of endothelial dysfunction by
63. Pean, N., Doignon, I., Garcin, I. et al. The receptor TGR5 protects the liver GPBAR1 agonism in portal hypertension involves a AKT/FOXOA1 dependent
from bile acid overload during liver regeneration in mice. Hepatology, regulation of H2S generation and endothelin‐1. PloS One, 2015;10(11):e0141082.
2013;58(4):1451–60. 86. Reinehr, R., Fischer, R., and Häussinger, D. Regulation of endothelin‐A
64. Cao, H., Chen, Z.X., Wang, K. et  al. Intestinally‐targeted TGR5 agonists receptor sensitivity by cyclic adenosine monophosphate in rat hepatic stel-
equipped with quaternary ammonium have an improved hypoglycemic effect late cells. Hepatology, 2002;36(4 Pt 1):861–73.
and reduced gallbladder filling effect. Sci Rep, 2016;6:28676. 87. Wang, Y.D., Chen, W.D., Yu, D., Forman, B.M., and Huang, W. The G‐pro-
65. Duan, H., Ning, M., Zou, Q. et  al. Discovery of intestinal targeted TGR5 tein‐coupled bile acid receptor, Gpbar1 (TGR5), negatively regulates
agonists for the treatment of type 2 diabetes. J Med Chem, hepatic inflammatory response through antagonizing nuclear factor kappa
2015;58(8):3315–28. light‐chain enhancer of activated B cells (NF‐kappaB) in mice. Hepatology,
66. Briere, D.A., Ruan, X., Cheng, C.C. et al. Novel small molecule agonist of 2011;54(4):1421–32.
TGR5 possesses anti‐diabetic effects but causes gallbladder filling in mice. 88. Guo, C., Xie, S., Chi, Z. et al. Bile acids control inflammation and meta-
PloS One, 2015;10(8):e0136873. bolic disorder through inhibition of NLRP3 inflammasome. Immunity,
67. Duan, H., Ning, M., Chen, X. et al. Design, synthesis, and antidiabetic activ- 2016;45(4):802–16.
ity of 4‐phenoxynicotinamide and 4‐phenoxypyrimidine‐5‐carboxamide 89. Pols, T.W., Nomura, M., Harach, T. et al. TGR5 activation inhibits athero-
derivatives as potent and orally efficacious TGR5 agonists. J Med Chem, sclerosis by reducing macrophage inflammation and lipid loading. Cell
2012;55(23):10475–89. Metab, 2011;14(6):747–57.
68. Pellicciari, R., Gioiello, A., Macchiarulo, A. et  al. Discovery of 6alpha‐ 90. Perino, A. and Schoonjans, K., TGR5 and immunometabolism: insights from
ethyl‐23(S)‐methylcholic acid (S‐EMCA, INT‐777) as a potent and selective physiology and pharmacology. Trends Pharmacol Sci, 2015;36(12):847–57.
agonist for the TGR5 receptor, a novel target for diabesity. J Med Chem, 91. Haselow, K., Bode, J.G., Wammers, M. et al. Bile acids PKA‐dependently
2009;52(24):7958–61. induce a switch of the IL‐10/IL‐12 ratio and reduce proinflammatory capa-
69. Spomer, L., Gertzen, C.G., Schmitz, B. et al. A membrane‐proximal, C‐ter- bility of human macrophages. J Leukoc Biol, 2013;94(6):1253–64.
minal alpha‐helix is required for plasma membrane localization and function 92. Hogenauer, K., Arista, L., Schmiedeberg, N. et al. G‐Protein‐coupled bile
of the G Protein‐coupled receptor (GPCR) TGR5. J Biol Chem, acid receptor 1 (GPBAR1, TGR5) agonists reduce the production of proin-
2014;289(6):3689–702. flammatory cytokines and stabilize the alternative macrophage phenotype.
70. Masyuk, T.V., Masyuk, A.I., Lorenzo Pisarello, M. et al. TGR5 contributes J Med Chem, 2014;57(24):10343–54.
to hepatic cystogenesis in rodents with polycystic liver diseases through 93. Ichikawa, R., Takayama, T., Yoneno, K. et al. Bile acids induce monocyte
cyclic adenosine monophosphate/Galphas signaling. Hepatology, differentiation toward interleukin‐12 hypo‐producing dendritic cells via a
2017;66(4):1197–218. TGR5‐dependent pathway. Immunology, 2012;136(2):153–62.
71. Hong, J., Behar, J., Wands, J. et al. Role of a novel bile acid receptor TGR5 94. Wammers, M., Schupp, A.K., Bode, J.G. et  al. Reprogramming of pro‐
in the development of oesophageal adenocarcinoma. Gut, 2010; inflammatory human macrophages to an anti‐inflammatory phenotype by
59(2):170–80. bile acids. Sci Rep, 2018;8(1):255.
72. Wacker, D., Wang, C., Katritch, V. et al. Structural features for functional 95. Li, T., Holmstrom, S.R., Kir, S. et al. The G protein‐coupled bile acid receptor,
selectivity at serotonin receptors. Science, 2013;340(6132):615–19. TGR5, stimulates gallbladder filling. Mol Endocrinol, 2011;25(6):1066–71.
73. Jensen, D.D., Godfrey, C.B., Niklas, C. et al. The bile acid receptor TGR5 does 96. Beuers, U., Hohenester, S., de Buy Wenniger, L.J. et al. The biliary HCO(3)
not interact with beta‐arrestins or traffic to endosomes but transmits sustained (‐) umbrella: a unifying hypothesis on pathogenetic and therapeutic aspects
signals from plasma membrane rafts. J Biol Chem, 2013;288(32):22942–60. of fibrosing cholangiopathies. Hepatology, 2010;52(4):1489–96.
74. Thomas, C., Gioiello, A., Noriega, L. et al. TGR5‐mediated bile acid sensing 97. Hohenester, S., Wenniger, L.M., Paulusma, C.C. et  al. A biliary HC03‐
controls glucose homeostasis. Cell Metabol, 2009;10(3):167–77. umbrella constitutes a protective mechanism against bile acid‐induced
75. Velazquez‐Villegas, L.A., Perino, A., Lemos, V. et al. TGR5 signalling pro- injury in human cholangiocytes. Hepatology, 2012;55(1):173–83.
motes mitochondrial fission and beige remodelling of white adipose tissue. 98. Merlen, G., Ursic‐Bedoya, J., Kahale, N. et al. The bile acid receptor TGR5
Nat Commun, 2018;9(1):245. regulates paracellular permeability and protects the liver through an impact
76. Perino, A., Pols, T.W., Nomura, M. et al. TGR5 reduces macrophage migra- on the tight junction protein JAM‐A. J Hepatol, 2017;66(1):S44–S45.
tion through mTOR‐induced C/EBPbeta differential translation. J Clin 99. Choi, M., Moschetta, A., Bookout, A.L. et al. Identification of a hormonal
Invest, 2014;124(12):5424–36. basis for gallbladder filling. Nat Med, 2006;12(11):1253–5.
77. Lieu, T., Jayaweera, G., Zhao, P. et al. The bile acid receptor TGR5 activates 100. Donepudi, A.C., Boehme, S., Li, F., and Chiang, J. Y. G‐protein‐coupled
the TRPA1 channel to induce itch in mice. Gastroenterology, 2014; bile acid receptor plays a key role in bile acid metabolism and fasting‐
147(6):1417–28. induced hepatic steatosis in mice. Hepatology, 2017;65(3):813–27.
298 THE LIVER:  REFERENCES

101. Häussinger, D. and Keitel, V. Dual role of the bile acid receptor Takeda G‐ 109. Iwakiri, Y., Shah, V., and Rockey, D.C. Vascular pathobiology in chronic
protein‐coupled receptor 5 for hepatic lipid metabolism in feast and famine. liver disease and cirrhosis – current status and future directions. J Hepatol,
Hepatology, 2017;65(3):767–70. 2014;61(4):912–24.
102. Watanabe, M., Houten, S.M., Mataki, C. et  al. Bile acids induce energy 110. Franke, A., Balschun, T., Karlsen, T.H. et  al. Sequence variants in IL10,
expenditure by promoting intracellular thyroid hormone activation. Nature, ARPC2 and multiple other loci contribute to ulcerative colitis susceptibil-
2006;439(7075):484–9. ity. Nat Genet, 2008;40(11):1319–23.
103. Ullmer, C., Alvarez Sanchez, R., Sprecher, U. et al. Systemic bile acid sens- 111. Karlsen, T.H., Franke, A., Melum, E. et al. Genome‐wide association anal-
ing by G protein‐coupled bile acid receptor 1 (GPBAR1) promotes PYY ysis in primary sclerosing cholangitis. Gastroenterology, 2010;
and GLP‐1 release. Br J Pharmacol, 2013;169(3):671–84. 138(3):1102–11.
104. Phillips, D.P., Gao, W., Yang, Y. et  al. Discovery of 112. Hov, J.R., Keitel, V., Schrumpf, E., Häussinger, D., and Karlsen, T.H.
trifluoromethyl(pyrimidin‐2‐yl)azetidine‐2‐carboxamides as potent, TGR5 sequence variation in primary sclerosing cholangitis. Dig Dis,
orally bioavailable TGR5 (GPBAR1) agonists: structure‐activity relation- 2011;29(1):78–84.
ships, lead optimization, and chronic in vivo efficacy. J Med Chem, 2014; 113. Masyuk, T.V., Masyuk, A.I., and LaRusso, N.F. TGR5 in the cholangiocili-
57(8):3263–82. opathies. Dig Dis, 2015;33(3):420–5.
105. Tacke, F. Targeting hepatic macrophages to treat liver diseases. J Hepatol, 114. Erice, O., Labiano, I., Arbelaiz, A. et  al. Differential effects of FXR or
2017;66(6):1300–12. TGR5 activation in cholangiocarcinoma progression. Biochim Biophys
106. McMahan, R.H., Wang, X.X., Cheng, L.L. et al. Bile acid receptor activa- Acta, 2018;1864(4 Pt B):1335–44.
tion modulates hepatic monocyte activity and improves nonalcoholic fatty 115. Doignon, I., Julien, B., Serriere‐Lanneau, V. et al. Immediate neuroendocrine
liver disease. J Biol Chem, 2013;288(17):11761–70. signaling after partial hepatectomy through acute portal hyperpressure and
107. Chavez‐Talavera, O., Tailleux, A., Lefebvre, P., and Staels, B., Bile acid cholestasis. J Hepatol, 2011;54(3):481–8.
control of metabolism and inflammation in obesity, type 2 diabetes, dyslipi- 116. Häussinger, D. and Schliess, F. Pathogenetic mechanisms of hepatic
demia, and nonalcoholic fatty liver disease. Gastroenterology, 2017; encephalopathy. Gut, 2008;57(8):1156–65.
152(7):1679–94 e3. 117. Karababa, A., Groos‐Sahr, K., Albrecht, U. et al. Ammonia attenuates LPS‐
108. Iwakiri, Y. and Groszmann, R.J. Vascular endothelial dysfunction in cirrhosis. induced upregulation of pro‐inflammatory cytokine mRNA in co‐cultured
J.Hepatol, 2007;46(5):927–34. astrocytes and microglia. Neurochem Res, 2017;42(3):737–49.
Bile Acids as Signaling
25 Molecules
Thierry Claudel and Michael Trauner
Hans Popper Laboratory of Molecular Hepatology, Division of Gastroenterology and Hepatology,
Department of Internal Medicine III, Medical University of Vienna, Vienna, Austria

INTRODUCTION acids by controlling a network of transporters and enzymes.


These sensors are nuclear, intracellular, and membrane proteins
Primary bile acids are synthesized by enzymes belonging to the that coordinate intra‐ and inter‐organ signaling and communica-
cytochrome P450 family (CYP). In humans the “classical path- tions by regulating bile acid homeostasis and a wide range of
way,” with CYP7A1 as limiting step, produces the primary bile other metabolic and immunological pathways (Figure 25.1) [1].
acids cholic acid (CA) and chenodeoxycholic acid (CDCA),
while the “alternative pathway,” involving CYP27A1, produces
CDCA [1, 2] (Table  25.1). In rodents, CDCA is metabolized
into α‐muricholic acid (αMCA) by Cyp2c70 and epimerized in BILE ACID‐ACTIVATED NUCLEAR
the intestine into β‐muricholic acid (βMCA) [3] (Table  25.1). RECEPTORS: FXR, PXR, CAR, VDR,
Primary bile acids are conjugated with taurine in rodents (TCA AND GR
and TβMCA) and glycine in humans (GCA and GCDCA).
Intestinal bacteria deconjugate these primary into secondary Nuclear receptors (NRs) are transcription factors that, upon
bile acids; βMCA can be then epimerized at C‐6 into ωMCA ligand binding or post‐translational modifications and cofactor
(Table  25.1); while CA, CDCA, and ωMCA can be 7α‐dehy- recruitment, modulate RNA polymerase II activity and gene
droxylated into deoxycholic acid (DCA), lithocholic acid expression. Several of these NRs are activated by bile acids.
(LCA), and hyodeoxycholic acid (HDCA), respectively
(Table  25.1) [4, 5]. CDCA is also converted into ursodeoxy-
cholic acid (UDCA) by the hydroxysteroid dehydrogenase and
Farnesoid X receptor
subsequently UDCA can be 7α‐dehydroxylated into LCA or The farnesoid X receptor (FXR, NR1H4) [8, 9] was originally
oxidized in the liver by Cyp2c70 into βMCA (Table 25.1). DCA, described as farnesol‐activated receptor [8], but several years
LCA, UDCA, HDCA, and ωMCA are secondary bile acids and later its key function as the main bile acid‐activated NR was
display either higher hydrophobicity and toxicity (LCA, DCA) identified [10–12]. Human FXR is maximally activated by
or hydrophilicity (UDCA, HDCA) than their primary bile acids CDCA (Table 25.1, Figure 25.1), followed by DCA, CA, LCA
precursors. Therefore bile acid composition depends on micro- and weakly UDCA, whereas mouse Fxr is marginally respon-
biota composition and, conversely, bile acids also contribute to sive to CDCA [13] which is only an intermediate in the synthe-
shape the microbiome [6]. sis of MCA in rodents (see earlier and Table  25.1). FXR can
Since bile acids are amphipathic molecules which can dam- accommodate both glycine/taurine conjugated‐ and free bile
age cell membranes, induce apoptosis, and have potential pro- acids in its ligand‐binding domain (LBD) [14]. In addition to
inflammatory actions at higher concentrations [7], various sensor bile acids, fatty acids such as arachidonic acid, docosahexaenoic
mechanisms are necessary to coordinate metabolic and trans- acid, and linoleic acid were found to activate FXR [15, 16].
port processes limiting the bile acid load and exposure, thus pro- Moreover, various semi‐synthetic bile acids such as the obeticholic
tecting the cellular integrity in the different organs handling bile acid (OCA)/6‐ethyl‐CDCA [17] and isoxazole derivatives such

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
300 THE LIVER:  BILE ACID‐ACTIVATED NUCLEAR RECEPTORS: FXR, PXR, CAR, VDR, AND GR

Table 25.1  Bile acid nomenclature, properties, and biological activities


Bile acid Full name Origin Biological activity
CA Cholic acid Liver (Cyp7a1) FXR agonist
CDCA Chenodeoxycholic acid Liver (Cyp27a1 and Cyp7a1) Most potent FXR agonist in human but not in mouse;
FPR receptor antagonist
αMCA α‐Muricholic acid Liver: oxidation of CDCA by Cyp2c70in rodents only FXR antagonist
βMCA β‐Muricholic acid Liver: oxidation of UDCA by Cyp2c70 in rodents only FXR antagonist
Intestine: epimerization of αMCA
LCA Lithocholic acid Intestine: 7α‐dehydroxylation of CDCA by Clostridium and FXR, PXR, VDR, TGR5 agonist
Eubacterium and/or 7α‐dehydroxylation of UDCA
UDCA Ursodeoxycholic acid Intestine: hydroxysteroid dehydrogenization of CDCA FXR partial agonist
GRα agonist
HDCA Hyodeoxycholic acid Intestine: 7α‐dehydroxylation of ωMCA by Clostridium and Lowers FXR antagonist ωMCA concentration
Eubacterium
DCA Deoxycholic acid Intestine: 7α‐dehydroxylation of CA by Clostridium and FXR, TGR5, S1PR2, CHMR3 agonist; FPR receptor
Eubacterium antagonist
ωMCA ω‐Muricholic acid Intestine: epimerization in C‐6 of βMCA by Bacteroides, FXR antagonist?
Escherichia, Clostridium, Eubacterium

as GW4064 [18] act as FXR ligands. These were used as discov- Initial studies showed that PXR was activated by 21‐carbon
ery tools in preclinical studies (GW4064) [19, 20] or developed steroids called pregnanes, resulting in the name pregnane X
further for treatment of primary biliary cholangitis [21] and receptor [40]. PXR can form heterodimers with RXR and stimu-
non‐alcoholic fatty liver (OCA) [22]. In addition to bile acid‐ lates expression of CYP3A4, a key enzyme for drug detoxifica-
derived FXR ligands, nonsteroidal FXR agonists were also tion and drug/drug interactions [41–43]. PXR is mainly
developed, which may differ in their pharmacokinetic and phar- expressed in liver and intestine [40–43] and is a promiscuous
macodynamic properties (see also Chapter 30) [23]. FXR can receptor activated by nearly 60% of the drugs existing due to its
form heterodimers with the retinoid X receptor alpha (RXRα; abnormally large LBD, accommodating a wide range of ligands
NR2B1), binding to FXR response elements (FXRE) and stimu- [44] including bile acids such as LCA.
lating gene transcription [8, 9, 24], but FXR also binds as a CAR is highly expressed in intestine and liver and is consti-
monomer repressing [25] or increasing gene expression [26]. tutively active even in absence of ligands, forming heterodimers
Based on genome binding analysis, FXREs are located in the with RXR [45].
vicinity of genes and clusters involved in lipid, fatty acid, ster- Despite species differences in their LBDs, PXR and CAR
oid metabolism, glycolysis, and transporters shared with another retained the capacity to respond to bile acids and to coordinate
NR, the liver receptor homolog 1 (LRH‐1; NR5A2) [27]. bile acid detoxification. Hence PXR activation lowers Cyp7a1
Whole‐body [28] or tissue‐specific [29] Fxr‐deficient mice (see later and Figure 25.2) without affecting Cyp27, Cyp8b1, or
revealed the central role of FXR signaling in lipid and bile acid Cyp7b1 in mice [46]. Moreover, LCA (Table  25.1) and the
homeostasis. At least four isoforms of FXR are expressed in metabolite 3‐keto‐LCA binds to the LBD and activates PXR at
human and mouse due to alternative first exon use and a stretch concentrations found in cholestasis [46]. PXR in turn represses
of amino acids inserted in the hinge domain between the DNA‐ Cyp7a1 and induces the bile acid transporter Oatp2, facilitating
binding domain (DBD) and the LBD [30]. Hereditary defects of LCA uptake into the hepatocytes where it is subsequently
FXR result in a severe form of progressive familial intrahepatic detoxified via coordinately induced Cyp3a11 [46]. CAR is pre-
cholestasis (PFIC) [31], while polymorphisms predispose to sent in cytosol and unable to affect gene transcription until its
intrahepatic cholestasis of pregnancy (ICP) [32] or influence translocation to the nucleus after phosphorylation, where it
fasting glycemia and free fatty acid levels [33]. A second FXR upregulates Cyp2b genes [47]. CAR activation in mouse liver
locus, called Fxrβ (NR1H5), identified in rodents and shown to results in induction of Fxr, Cyp8b1, Cy27a1, Cyp39a1, and
encode a functional protein responsive to lanosterol, but not bile Ugt1a1, which altogether not only reduces the total amount of
acid, is only a pseudogene in primates [34]. bile acid but also makes the bile acid pool composition less
toxic [48]. Taken together, CAR and PXR act as a second line of
defense against toxic bile acids as additional backup to FXR
Pregnane X receptor/constitutive signaling.
androstane receptor
In addition to FXR, other NRs such as the pregnane X receptor
Vitamin D receptor
(PXR; NRI2) [35, 36] and the constitutive androstane receptor
(CAR; NR1I3) [37] were shown to respond to bile acids Vitamin D receptor (VDR; NR1I1) forms heterodimers with
(Figure 25.1 and Table 25.1). Fxr‐deficient mice have a massive RXR and is primarily activated by vitamin D3 and as such regu-
induction of enzymes involved in bile acids transport and detox- lates calcium homeostasis as well as inflammation [49].
ification, which are target genes of PXR and CAR [38], allow- Moreover, VDR was identified as an additional bile acid‐respon-
ing these receptors and their downstream targets to partly sive NR in intestine, where it is activated by LCA (Table 25.1
compensate for loss of FXR. Moreover, combined deficiency of and Figure 25.1) [50] and induces CYP3A4, metabolizing LCA
PXR and CAR increased the sensitivity to bile acid toxicity in and counteracting its pro‐carcinogenic properties  [50]. VDR
mice [39]. activation after vitamin D3 gavage reduced intrahepatic and
25:  Bile Acids as Signaling Molecules 301

N‐Formyl CDCA
peptides DCA LCA DCA DCA
Pro-EGF EGF

TGR5 S1pR2 MMP EGFR


FPR receptor CHMR3
PLCβ Cytosol
Gα cAMP RAS

ERK AKT PLC ERK


PKC PKC

PLA2 Ca2+
COX influx
5-LO
ROS

LTB4
FXR PXR CREB SPHK2
VDR p-FXR Nucleus

PPARγ p-SHP

Chemotaxis BA/lipid/glucose NO production BA/lipid


homeostasis Anti-inflammatory metabolism

Figure 25.1  Signaling pathways activated by bile acids. CDCA, LCA, and DCA are the main ligands of the nuclear receptor FXR as key regulator
of bile acid homeostasis, lipid, and glucose metabolism (center). LCA activates the nuclear receptors PXR and VDR, regulating bile acid detoxifica-
tion and inflammation (center). LCA and DCA are also ligands for the membrane‐bound G‐coupled receptor TGR5, which stimulates the produc-
tion of cyclic AMP (cAMP) activating the transcription factor cAMP response element‐binding protein (CREB), which in turn regulates nitric oxide
(NO) levels and suppresses inflammation (center). Bile acids such as DCA can activate either the sphingosine 1 phosphate receptor 2 (S1PR2) or
the muscarinic receptor 3 (CHMR3) or the epidermal growth factor receptor (EGFR), which all reside in the cell membrane (right part of the
image). S1PR2 and CHMR3 activate the phospholipase C beta (PLCβ), which activates membrane metalloproteinase (MMP). MMP releases EGF
from the membrane‐bound precursor pro‐EGF, allowing EGF then to bind to EGFR, in turn activating the small GTPase RAS resulting in activation
of protein kinase C zeta (PKCζ), phosphorylating FXR and SHP. S1PR2 can also directly activate ERK signaling, leading to increased activity of
the sphingosine kinase 2 (SPHK2), catalyzing the phosphorylation of sphingosine and controlling angiogenesis, tumorigenesis, and lipid metabo-
lism. N‐Formyl peptides are bacterial components activating the G‐coupled receptor formyl peptide receptor (FPR), CDCA blocks ligand binding
to this receptor (left part of the image). FPR stimulates phospholipase C (PLC) and protein kinase C (PKC) signaling, which together with AKT
leads to calcium influx and reactive oxygen species (ROS) formation. Calcium influx also initiates a negative feedback desensitizing FPR/FPRL1.
FPR/FPRL1 also activates extracellular signal‐regulated kinase (ERK), which increases the activity of the phospholipase A2 (PLA2), arachidonate
5‐lipoxygenase (5‐LO) and cyclooxygenase (COX) which produce leukotriene B4 (LTB4), a potent ligand activating the nuclear receptor PPARγ.

circulating bile acid levels by promoting their urinary excretion developed which may act as new GR modulators, retaining
in mice [51]. Since LCA is a secondary bile acid obtained after anti‐inflammatory properties without disturbing glucose
bacterial modification (Table 25.1), the impact of VDR on intes- metabolism [59].
tinal microbiota and bile acid metabolism will be of future
interest.
Fibroblast growth factor 19
FXR induces the expression of Fgf15 (mice) [60]/FGF19
Glucocorticoid receptor (humans) [61] in the terminal ileum. In contrast to mouse Fgf15,
Glucocorticoids activate the glucocorticoid receptor (GR; FGF19 is also expressed in hepatocytes [61]. Intestinal FGF19/
NR3C1) [52]. Two isoforms, GRα and GRβ, exist by exon 9 Fgf15 reaches the liver via the portal vein and binds to the tyros-
splicing: GRα is a classical NR activated by glucocorticoids, ine kinase receptor FGFR4, or alternatively to FGFR1c in mac-
while GRβ acts as a dominant negative receptor [53]. rophages, adipocytes, and brain (Figure  25.3) [62]. The
Importantly. GRα is also activated by UDCA but not other bile interactions between FGF19/Fgf15 and FGFR4 or FGFR1c are
acids (Table  25.1) [54] but the binding of UDCA to GRα is only possible via the membrane‐bound protein βKlotho, which
weaker than other glucocorticoids, such as dexamethasone [55]. allows a tissue‐specific fine tuning of the ligand–receptor inter-
Activation of GRα may contribute to the immunosuppressive action since its expression is more restricted than that of FGFR4
properties of UDCA, such as reducing the hepatic expression of or FGFR1c [62]. PXR and VDR can also activate the intestinal
class I and II major histocompatibility complex in PBC [56, 57]. expression of FGF19/Fgf15 [62]. FGF19/Fgf15 reduces bile
GRα may also be required to mediate other immunomodulatory acid synthesis (by repressing CYP7A1), gluconeogenesis, and
effects of UDCA, such as suppressing interleukin 18 and inter- lipogenesis and increases hepatic glycogen and protein synthe-
feron gamma production by Kupffer cells and NKT cells, sis independently of insulin signaling (see Figures  25.2 and
respectively [58]. Novel UDCA derivatives have been 25.3) [63]. In addition, FGF19/Fgf15 controls gallbladder
302 THE LIVER:  BILE ACID‐ACTIVATED NUCLEAR RECEPTORS: FXR, PXR, CAR, VDR, AND GR

3 Alternative export BA

MRP4 MRP3 OSTα/β Phase III

CAR CAR PXR VDR FXR


OATP1A1
2
CAR UGTs
BA OATP1B2 G/S-BA Phase II
SULTs
OA CAR MRP2 OA, Bili
OATP1B3 FXR CYPs ox-BA Phase I

Canalicular export
PXR
OATP1A4 PXR BA
Uptake

CYP7A1 PPARα MDR2/3 PL

BA HNF4α FGF19 FXR BSEP BA

BA NTCP HNF4α SHP FXR ABCG5


Chol
ABCG8
BA
1

4 Bile synthesis and detoxification Hepatocyte

Figure 25.2  Nuclear receptors and transcriptional regulation of hepatocellular bile acid transport and metabolism. (1) Bile acid (BA) uptake:
HNF4α regulates basolateral Na+‐dependent (NTCP) and is repressed directly by bile acids or indirectly via FXR/SHP. FXR upregulates human
OATP1B3, while PXR increases OATP1A4 expression. (2) Canalicular export of bile acids into the bile: all canalicular transporters involved in bile
formation (BSEP for BAs, MDR3/Mdr2 for phospholipids (PL), MRP2 for BAs, conjugated bilirubin, and GSH, ABCG5/8 for cholesterol (Chol))
are upregulated by FXR. CAR and PXR positively regulate MRP2 and PPARα upregulates MDR3/Mdr2. FXR also upregulates PPARα expression
in humans. (3) Alternative export into blood to allow renal excretion: CAR positively regulates MRP3 and MRP4, while PXR and VDR upregulate
MRP3. FXR induces OSTα/β. (4) Bile acid synthesis and detoxification: CYP7A1 is the rate‐limiting step in BA synthesis. FXR represses CYP7A1
expression via the induction of the hepatic and intestinal FGF19 in humans, Fgf15 in mouse which is produced only in the terminal ileum (not
shown). FXR also represses CYP7A1 expression via the induction of SHP which in turn decreases HNF4α activity. HNF4α activity can also be
directly impaired by BAs. Bile acids can be detoxified during phase I by oxidation catalyzed by enzymes from the cytochrome P450 family (CYPs).
After oxidation, ox‐bile acids can be conjugated by sulfation catalyzed by the sulfotransferases (SULTs) or glucuronidated by the UDP‐glucurono-
syltransferases (UGTs). FXR/CAR/PXR upregulate CYPs, SULTs, and UGTs expression. After conjugation, glucuronidated bile acids (G‐BA) and
sulfated‐bile acids (S‐BA) can be excreted either into the bile via MRP2 or into the bloodstream via the alternative transporters MRP3, MRP4, and
OSTα/β. Arrows: upregulation or metabolite flux; blocked arrows: transcriptional repression.

filling (via FGFR4), lowers food intake and increases insulin promoter sequences  [67] and by promoting histone modifica-
sensitivity in the brain (via FGFR1c), metabolically activates tions reversed by Shp [68]. However, the rather weak bile acid
white adipose tissue (WAT) adipocytes (via FGFR1c), and phenotype of Shp‐deficient mice [69, 70] combined with the
decreases macrophage uptake of oxidized low‐density lipopro- marked decrease of Cyp8b1 but preserved expression of Cyp7a1
teins (LDL) (and thereby foam cell formation) by reducing in Lrh‐1‐deficient mice [71, 72], suggested that intestinal factors
CD36 expression (via FGFR1c) (Figure  25.3) [62]. Thus might also contribute to Cyp7a1 regulation [29].
FGF19/Fgf15 acts as a bile acid‐induced key regulator of inter- Indeed, experiments with liver and ileum‐specific Fxr‐defi-
mediary metabolism. cient mice suggest that the ileal route of Cyp7a1 repression via
the FGF19/Fgf15 pathway dominates over hepatic‐negative
feedback pathways, indicating that a functional gut–liver sign-
Bile acid signaling actions mediated aling is a prerequisite for bile acid homeostasis [29]
by nuclear receptors (Figure 25.3). FGF19/Fgf15 after binding to FGFR4/βKlotho
Bile acid synthesis activates the tyrosine kinase function of Fgfr4, which represses
Cyp7a1 through c‐Jun terminal kinase (Jnk) and Shp‐depend-
CYP7A1 is the rate‐limiting enzyme in bile acid synthesis and is ent pathways [60]. In addition to Fxr/Shp and FXR/FGF19,
transcriptionally regulated via a negative feedback by bile acids bile acids can directly decrease Hnf4α gene expression [73]
[64]. Bile acids binding to FXR induces expression of the NR and impair Hnf4α activation of the Cyp7a1 promoter by block-
small heterodimer partner (SHP; NR0B2), which interacts with ing the recruitment of its coactivators the peroxisome prolif-
LRH‐1 to suppress CYP7A1 transcription (Figure 25.2) [65, 66]. erator activated receptor gamma coactivator (Pgc‐1α) or the
Lrh‐1 and the NR hepatocyte nuclear factor 4 alpha (Hnf4α; cAMP response element‐binding protein (CBP) [74].
NR2A1) drive Cyp7a1 baseline expression by binding to specific Moreover, PXR (e.g. activated by hydrophobic bile acids)
25:  Bile Acids as Signaling Molecules 303

FGF15

FXR βKlotho GLP1-R


TGR5 FGFR1c
FXR FGFR4
Glycogen Neuro- Glucose Insulin
synthesis transmitters metabolism sensitivity
TG FA metabolic &
clearance oxidation rate satiety
&
TG HDL Protein feeding

synthesis synthesis synthesis


Brain

FGF15 FGF15
βKlotho βKlotho/FGFR1c TGR5/FGFR1c
FGF15
FGFR4

Liver WAT Macrophage


Less TG uptake CD36
Leptin Low intake oxLDL
Small adipocytes Foam cells
FXR VDR FGF15
VDR
Endothelium
FXR
TGR5
eNOS TGR5
AKT
GLP-1 TGR5
BAT Muscle
PXR FXR βKlotho Intestine
VDR FGFR1c/FGFR4 DIO2 energy
expenditure
SMC Insulin sensitivity
DIO2 energy
Cyps Cell migration Role for FGF19? expenditure
induction &
detoxification

Figure 25.3  FXR and TGR5 control lipid and glucose homeostasis. FXR activation in the liver promotes triglycerides (TG) clearance (green),
reduces TG synthesis (red). Concomitantly, FXR activation increases glycogen and protein synthesis (green) thus helping to control glycemia. FXR
and VDR activation in the intestine release FGF19 in humans (Fgf15 in mice) which via the portal vein circulates to the liver where it binds to its
receptor FGFR4 and βKlotho to repress bile acid synthesis. TGR5 activation in L‐cells produces the glucagon‐like peptide 1 (GLP‐1), which regu-
lates glucose sensitivity and insulin resistance. FXR activation in endothelium increases smooth muscle cell (SMC) motility in vitro and the
endothelial nitric oxide synthase expression (eNOS). FGFR4, βKlotho, and FGFR1c are expressed in smooth muscle cells but the FGF15/19 role
is not yet clear. FGF15/19 signals in white adipose tissue (WAT) where it decreases TG uptake and adipocyte size and increases leptin expression.
TGR5 activation in brown adipose tissue (BAT) increases energy expenditure via the induction of the expression of the type II iodothyronine deio-
dinase (DIO2). In muscles TGR5 activation increases AKT signaling and increases DIO2 expression, energy expenditure, and insulin sensitivity. In
macrophages TGR5 activation and FGF15/19 signaling via FGFR1c/βKlotho lowers CD36 expression, reducing oxidized LDL (ox‐LDL) intake
and thus foam cells formation. In the brain FXR and TGR5 regulates neurotransmitters. GLP‐1 resulting from TGR5 activation in the gut increases
insulin sensitivity and FGF15/19 signaling increases the metabolic rate and improves glucose metabolism.

impairs HNF4α binding to the CYP7A1 promoter by reducing acid pool composition, therefore identifying Lrh‐1 as the central
the interaction of PGC‐1α with HNF4α [75] (Figure  25.2). regulator of Cyp8b1 expression [71, 72], also indirectly control-
FXR induces the expression of peroxisome proliferator acti- ling Cyp7a1 expression by generating FXR antagonists.
vated receptor alpha (PPARα; NR1C1) in humans [76] and
Regulation of phase I and II bile acid metabolism
PPARα in turn reduces CYP7A1 transcription via reduced
by bile acid‐activated nuclear receptors
HNF4α binding, effects which might contribute to the risk of
gallstone formation after treatment with PPARα activators Bile acid hydroxylation (phase I) and conjugation (phase II)
such as fibrates but could also contribute to their anti‐choles- makes them more hydrophilic and facilitates their excretion into
tatic effects [1, 77] (Figure 25.2). urine as an additional/alternative route for bile acid elimination
Another enzyme, CYP8B1, controls the relative hydropho- under cholestatic conditions [81]. Human CYP3A4 and the
bicity of the bile acid pool through CDCA synthesis and rodent Cyp3a11 are key cytochrome P450 enzymes for bile acid
Cyp8b1‐deficient mice lack the negative feedback on Cyp7a1 oxidation (see earlier). CYP3A4 expression is induced by FXR,
[78] due to increased levels of FXR antagonists such as αMCA, PXR, VDR, and CAR, providing  –  together with sulfation or
βMCA, and UDCA (Table 25.1) [79, 80]. Lrh‐1‐deficient mice glucuronidation – a mechanism to limit hepatocellular damage
have reduced Cyp8b1 expression and low CA levels in their bile by accumulating bile acids (Figure 25.2).
304 THE LIVER:  BILE ACID‐ACTIVATED NUCLEAR RECEPTORS: FXR, PXR, CAR, VDR, AND GR

SULT2A1, the dehydroepiandrosterone sulfotransferase, cat- induces PPARα in human, FXR could also activate MDR3 indi-
alyzes sulfo‐conjugation of bile acids (an important bile acid rectly via PPARα (Figure 25.2) [1]. FXR also induces ABCG5/
detoxification pathway during cholestasis in humans) and is G8 expression, resulting in increased cholesterol excretion in
induced by FXR, PXR, VDR, and CAR [1]. Bile acid glucuro- bile [1].
nidation is catalyzed by the UDP‐glucuronosyltransferases Canalicular bile acid efflux is thus mainly mediated via bile
UGT2B4 and UGT2B7 [26, 82]. Bile acids induce human acids requiring FXR/CAR/PXR/VDR as a key transcription fac-
UGT2B4 via activation of FXR [26], but UGT2B4 is also acti- tor coordinating a balanced bile composition.
vated by PPARα agonists [82]. Since FXR upregulates PPARα
and its target genes expression [76], bile acids could induce Alternative basolateral bile acid export
UGT2B4 expression directly via FXR and indirectly via FXR‐
Basolateral bile acid excretion back into portal blood represents
mediated induction of PPARα.
an alternative route for bile acid elimination during cholestasis
(when canalicular excretion is blocked). Alternative basolateral
Regulation of bile acid transport by nuclear receptors bile acid export is mediated by MRP3 and MRP4 and the
Basolateral hepatocellular bile acid uptake organic solute transporters OSTα and OSTβ [89] (Figure 25.2).
These export systems are expressed at low levels under normal
Hepatic bile acid uptake is mediated by a high‐affinity Na+‐ conditions but can be significantly upregulated in cholestasis
dependent transporter NTCP (SLC10A1) and a family of multi‐ [89]. Since MRP3, MRP4, and OSTα/OSTβ are able to trans-
specific organic anion transporters (OATPs; SLC21A), port sulfated and glucuronidated bile acids, the induction of
mediating the quantitatively less important Na+‐independent these transporters may facilitate renal excretion of these bile
uptake of conjugated or free bile acids mediated by facilitated acids in patients with chronic cholestasis [89].
exchange with intracellular anions [83] (Figure 25.2). Induction of both Mrp3 and Mrp4 is independent of Fxr in
Regulation of NTCP by bile acids differs considerably among rodent models of cholestasis [94, 95] (Figure  25.2). PXR and
humans, mice, and rats [84]. Negative feedback inhibition of VDR are able to induce human and rodent MRP3 expression
Ntcp is mediated via Fxr‐Shp‐dependent or ‐independent mech- [96, 97], while CAR ligands induced both human and rodent
anisms to limit bile acid uptake [85, 86]. Induction of Shp by MRP3 and MRP4 [81]. The bile acid transporter OSTα/β is
bile acid‐activated FXR inhibits the activity of Hnf4α [87] induced by FXR and baseline Ostα/Ostβ expression is reduced
(Figure  25.2). In humans, the NTCP promoter lacks HNF4α in Fxr‐knockout mice [1]. Taken together, similar to canalicular
response elements [84] and SHP acts mainly by suppressing transporters, multiple NRs (FXR, PXR, VDR, and CAR) coor-
GRα‐mediated activation of NTCP [88]. Similar to NTCP, dinate basolateral bile acid efflux.
repression of the sodium‐independent bile acid uptake system in
humans, OATP1B1, is also mediated by FXR, involving SHP,
HNF4α, and HNF1α [89]. Bile acid‐activated nuclear receptor signaling
in lipid/glucose metabolism, endothelial
Canalicular bile acid excretion
function, and inflammation
Canalicular excretion of bile acids is the rate‐limiting step in
High‐density lipoprotein metabolism
bile formation. Bile acids such as CA, CDCA, and UDCA and
their tauro/glycoconjugate are excreted into the bile canaliculus High‐density lipoproteins (HDLs) are heterogeneous lipid parti-
via the bile salt export pump BSEP (ABCB11) (Figure  25.2). cles containing phospholipids, cholesterol, sphingosine 1 phos-
Human, rat, and mouse BSEP promoters are transcriptionally phate (S1P), and apolipoprotein A‐I as their major structural
activated by FXR [1, 90] and Bsep baseline expression is proteins. HDLs transport cholesterol from the periphery back to
reduced in Fxr‐knockout mice [28] and humans with FXR loss the liver where it can be excreted as free cholesterol or, after
of function [31]. A role for VDR in BSEP repression via direct conversion into bile acids, into bile to be eliminated via feces in
VDR–FXR interaction in vitro has also been postulated [91]. a process called “reverse cholesterol transport” (Figure  25.1).
Bile acids with two negative charges such as sulfated tauro‐ Resins (such as cholestyramine), bind to bile acids in the intes-
or glyco‐LCA are transported by multidrug resistance‐associ- tine, resulting in bile acid malabsorption and increasing HDL
ated protein MRP2 (ABCC2) [92]. MRP2 is regulated by several cholesterol levels, while dietary bile acid supplementation low-
NRs, reflecting its diverse substrate spectrum. FXR binds to ers HDL cholesterol [98]. This can be explained by repression
FXREs in the promoter, shared with CAR and PXR [93] of apoA‐I gene expression by FXR, resulting in reduced hepatic
(Figure 25.2). HDL synthesis (Figure 25.3) [25]. FXR activation also increases
Additional hepatobiliary transport systems in the canalicular cholesteryl ester transfer protein (CETP) mass and activity [99]
membrane include a phospholipid floppase (MDR3/Mdr2 in as well as phospholipid transfer protein (PLTP) in vitro [98].
rodents) for phosphatidylcholine translocation, the cholesterol PLTP transfers phospholipids from triglycerides to HDL,
two half‐transporters ABCG5/8 for sitosterol and cholesterol whereas CETP catalyzes exchange of triglycerides and choles-
export, a P‐type ATPase (Fic1/FIC1; ATP8B1) whose substrate/ terol esters between HDL and apoB lipoproteins and CETP
function is still unclear and an Cl−/HCO3− anion exchanger 2 activity correlates with cardiovascular disease risk. This,
(SLC4A2/AE2), all of them involved in bile formation [1, 81]. together with lowering HDL cholesterol and increasing LDL
FXR enhances human MDR3 transcription, while PPARα stim- cholesterol (see later) (Figure  25.3) may impact on long‐term
ulates the expression of rodent Mdr2 and MDR3; since FXR cardiovascular safety of FXR agonists [98].
25:  Bile Acids as Signaling Molecules 305

Triglyceride and low‐density lipoprotein metabolism and plaque accumulation favoring the development of
atherosclerosis [119]. Like FXR, VDR also induces eNOS in
Bile acid supplementation to patients lowered serum triglycer-
endothelial cells [120] (Figure 25.3).
ides [98], while Fxr‐deficient mice are hypertriglyceridemic due
to an increase of the lipoprotein lipase (LPL) activity inhibitor
Inflammation
apoC‐III and lowering of the LPL activator apoC‐II [28, 100].
FXR (via induction of SHP) also reduces de novo lipogenesis by FXR, CAR, PXR, and VDR are all negative acute phase genes
repressing Srebp1c [101] and reduces very low‐density lipopro- (i.e. their expression and signaling are reduced during the acute
tein (VLDL) synthesis via repression of HNF4α transcriptional phase response) [121–123]. FXR stimulation by bile acids (e.g.
activity on the promoter of the microsomal triglyceride transfer accumulating during sepsis‐induced cholestasis) is anti‐inflam-
protein (MTTP) [102]. Conversely FXR increases the expres- matory by antagonizing nuclear factor‐κB (NF‐κB) signaling
sion of the VLDL receptor (VLDLR) [103] and PPARα as well [124]. In addition, FXR antagonizes JNK signaling during
as β‐oxidation [76] (Figure  25.3). Lipoprotein(a) (Lp(a)) is a inflammation by inducing superoxide dismutase 3 (SOD3)
lipoprotein formed via the covalent binding of apoA with [125]. Conversely, binding of NF‐κB to well‐conserved binding
apoB‐100 in LDL and high plasma Lp(a) levels increased ath- sites in the promoters of FXR target genes resulted in repression
erogenicity [104]. FXR directly lowers Lp(a) gene expression of BSEP, ABCG5/G8, and MRP2, while promoters of OSTα
and indirectly via FGF19‐mediated mechanisms [105, 106]. and OSTβ were unexpectedly activated rather than repressed.
FXR activation in humanized mouse with chimeric livers also This may suggest that activation of NF‐κB could help to main-
increased LDL cholesterol levels due to SREBP2 repression, tain bile acid excretion via OSTα/β into the blood during inflam-
which in turn reduces LDL receptor expression in hepatocytes mation/cholestasis [126]. Activation of FXR in the small
and LDL uptake by the liver [107]. In conclusion, bile acids are intestine and colon attenuates interleukin responses (IL‐6,
hypotriglyceridemic through lowering the production and IL‐1β), controls microbiota, maintains intestinal integrity, and
increasing clearance of triglycerides, but conversely increase overall ameliorates inflammation [127, 128]. VDR and FXR
LDL cholesterol (Figure 25.3). upregulate cathelicidin expression in liver, biliary ducts, and
intestine to control bacterial growth [129]. Gut microbiota also
Glucose metabolism regulates FXR activation in intestine by controlling the release
of the FXR antagonist βMCA (see earlier) [79, 130].
FXR deletion in mouse resulted in insulin resistance coupled
with glucose intolerance; whereas treatment with FXR agonist
lowered blood glucose by repressing hepatic gluconeogenesis
and enhancing glycogen synthesis and storage (Figure  25.3)
[108], partially via FGF19 [63,109]. Moreover, absence of FXR G PROTEIN‐COUPLED‐RECEPTOR
increased plasma free fatty acids levels and glucose production FOR BILE ACIDS TAKEDA G PROTEIN‐
in liver, thus impairing peripheral glucose disposal in mice COUPLED RECEPTOR 5
[110]. Finally, FXR enhances glucose‐stimulated insulin secre-
tion in pancreatic beta‐cells [111]. Altogether, FXR activation Discovery and properties
counteracts insulin resistance and glucose intolerance, which
may be of interest for the treatment of type 2 diabetes mellitus In addition to NRs, bile acids also activate G protein‐coupled
and related/associated disorders (Figure 25.3). receptors (GPCRs) such as TGR5 (also see Chapter  25). Two
groups cloned a membrane receptor which was shown to be
activated by bile acid and named BG37 or TGR5 [131, 132].
Signaling in endothelium and smooth muscle cells
TGR5 is activated mainly by LCA or DCA, with a divergent
Vascular smooth muscle cells express FXR [112], where it con- potency from FXR [131, 132] (Figure 25.1 and Table 25.1) and
trols PLTP expression, apoptosis, and attenuates inflammation by bile acid concentrations achieved in post‐prandial state, dur-
by inducing SHP expression which in turn interferes with NFκB ing cholestasis and portosystemic shunting or therapeutic bile
signaling [113] (Figure  25.3). In addition, FXR activation acid therapy. Notably, TGR5 is expressed in tissues where FXR
reduces platelet activation/aggregation in response to thrombin is absent, such as lung, spleen, leukocytes, white (WAT) and
and collagen [114]. FGFR4 and FGFR1c are also expressed in brown adipose tissues (BAT) [131, 132], with its highest expres-
smooth muscle cells in human arteries [115], together with sion in colon and gallbladder [133] (for more details see
βKlotho, where it mediates FGF21 signaling [116], while the Chapter 24).
relevance of FGF19 signaling in vascular wall is still unclear
(Figure 25.3). FXR also regulates vascular tension by control-
ling endothelial nitric oxide synthase eNOS (Figure 25.3) [117].
Metabolism and energy homeostasis
Since chronic liver disease leads to portal hypertension and cir- Mice fed with high‐fat diet containing bile acids were protected
rhosis, FXR activation in sinusoidal endothelial cells reduces from obesity – despite preserved food intake – which was attrib-
intrahepatic vascular resistance and thereby portal pressure by uted to increased energy expenditure [134]. TGR5 activation by
increasing eNOS and reducing endothelin 1 [118]. VDR is also bile acid induced cAMP production, resulting in upregulation of
expressed in endothelial cells (outside the liver) and its disrup- deiodinase 2 (D2) expression, which converts inactive thyroid
tion promotes a pro-inflammatory phenotype with induction of hormone T4 into the active T3 form. T3, by activating the NR
interleukin 6 (IL‐6) and VCAM‐1, resulting in enhanced rolling thyroid hormone receptor alpha (TRα1, NR1A1), induced
306 THE LIVER:  SPHINGOSINE 1 PHOSPHATE RECEPTOR 2, CHOLINERGIC RECEPTOR MUSCARINIC 3

uncoupling protein 1 (UCP1) expression, thus increasing energy FORMYL PEPTIDE RECEPTORS
expenditure in BAT and skeletal muscle [134] (Figure 25.3), an
effect also principally found in humans [135]. TGR5 also stimu- Formyl peptide receptors (FPRs) are G‐coupled receptors
lates production of glucagon‐like peptide 1 (GLP‐1) in intesti- activated by bacterial chemotactic formylated peptides and play
nal enteroendocrine cells and therefore contributes to control of a key role in sensing of bacteria and mediating leukocyte recruit-
glucose homeostasis [136]. The use of bile acid‐binding resins ment to infection sites. N‐Formyl‐methionyl‐leucylphenylala-
(cholestyramine, colestipol, colesevelam), resulting in activa- nine (fMLP) is one of the most powerful ligands activating the
tion of intestinal TGR5 by shifting bile acid signaling to the high‐affinity FPR and formyl peptide receptor like 1 (FPRL1)
colon by inhibition of ileal bile acid absorption, improved the (Figure 25.1) [150]. Cholestatic patients showed an impairment
control of glycemia in diabetic patients and hypercholester- of FPR and FPRL1 activity [151], since bile acids such as
olemia [98] and reduced hepatic glycogenolysis and glucose CDCA and CA act as FPR and FPRL1 antagonists, while
production in mice [137] (Figure 25.3). Bile acid‐binding resins UDCA has only a rather weak inhibitory effect [151]. Bile acids
have beneficial metabolic effects by depleting hepatic choles- prevent binding of FPR and FPRL1 with fMLP by steric hin-
terol via stimulation of bile acid synthesis/CYP7A1 activity and drance and glycoconjugated CDCA was identified as the most
improving glucose control by GLP‐1 induction via TGR5. efficient competitor [151, 152]. Bile acid antagonism of FPR/
Therefore activating TGR5 might be an attractive approach to FPRL1 could thus contribute to immunosuppressive effects of
treat metabolic conditions such as diabetes, fatty liver, or bile acids [153, 154]. During the acute phase response up to
dyslipidemia. 25% of the proteins produced in the liver are accounted for by
serum amyloid A (SAA), which also binds and activates FPRL1.
Anti‐inflammatory effects of TGR5 FPRL1 induces cyclooxygenase 2 (COX2) and synthesis of
prostaglandins, activating the peroxisome proliferator receptor
Tgr5‐deficient mice are more susceptible to liver inflamma- γ (PPARγ; NR1C3) (Figure 25.1) [155]. How bile acid signaling
tion induced by LPS [138]. TGR5 inhibits inflammatory is involved in crosstalk with SAA/FPRL1 and synthesis of
cytokine production by c‐FOS/p65 in macrophages [139], PPARγ agonists during cholestatic conditions is still unknown
while pharmacological stimulation of TGR5 reduced myeloid (Figure 25.1) [150].
cell activation and autoimmune encephalitis [140] and pre-
vented atherosclerosis by lowering macrophage plaque con-
tent and inflammatory cytokine production [141, 142]. In
addition, TGR5 repressed CD36 and SR‐A expression in mac- SPHINGOSINE 1 PHOSPHATE
rophages, thus reducing uptake of oxidized LDL [141]
(Figure  25.3). FXR/TGR5 coactivation also enhanced the
RECEPTOR 2, CHOLINERGIC RECEPTOR
intrahepatic content of alternatively activated macrophages MUSCARINIC 3, AND EPIDERMAL
and anti‐inflammatory monocytes in mice with NAFLD GROWTH FACTOR RECEPTOR
[143]. Such anti‐inflammatory effects could contribute to the
beneficial effects of pharmacological TGR5 stimulation in Several bile acid‐activated GPCRs share a common signaling
cholestatic and metabolic disorders (see Chapter  24 for fur- pathway which can result in activation of EGFR signaling and
ther details). are therefore summarized in this section. Bile acids such as
DCA can activate the sphingosine 1 phosphate receptor 2
(S1PR2), the muscarinic receptor 3 (CHMR3), or the epidermal
Role of TGR5 in the biliary tree growth factor receptor (EGFR) in the cell membrane
In human cholangiocytes, TGR5 colocalizes with the chloride (Figure 25.1). Moreover, S1PR2 and CHMR3 activate the phos-
channel cystic fibrosis transmembrane conductance regulator pholipase Cβ (PLCβ) which in turn activates membrane metal-
(CFTR) and the apical sodium‐dependent bile salt uptake loproteinase (MMP). releasing EGF from the membrane‐bound
transporter (ASBT) in the apical membrane of gallbladder epi- precursor pro‐EGF, allowing EGF to bind to EGFR (Figure 25.1)
thelial cells, where it mediates chloride secretion via CFTR
[144]. TGR5 is also localized to primary cilium of cholangio-
Sphingosine 1 phosphate receptor 2
cytes, where it regulates ductular bile formation by increasing
bile acid reabsorption and water secretion [145]. Targeting Sphingosine can be phosphorylated by the sphingosine kinase 1
TGR5 in inflammatory biliary diseases might thus be tempting and 2 (SPHK1 and SPHK2) to generate bioactive S1P. S1P in
but TGR5 has recently been implicated in the pathogenesis of turn activates specific G protein‐coupled receptors known as
pruritus [146]. sphingosine 1 phosphate receptors (S1PR) 1 to 5. While S1PR1
Moreover, TGR5 also promotes cell proliferation in is ubiquitously expressed, S1PR2 is mainly found in liver, kid-
response to bile acid in adenocarcinoma [147] and induces ney, heart, brain, lung, and vascular smooth muscle cells [156].
cardiac hypertrophy in cardiomyocytes [148], TGR5 is also a S1PRs activate ERK, AKT, JNK, and SPHK [156]. S1PR2 acti-
neurosteroid receptor in brain involved in the generation of vation increases pro-inflammatory cytokine (IL‐1β, TNF‐α)
reactive oxygen species (ROS) [149]. Therefore the quest for secretion and S1pr2‐deficient mice have reduced TNF‐α and
safe TGR5 modulators with therapeutic applications for IL‐1β levels [157]. After a meal, bile acid activation of hepatic
cholestatic and metabolic disorders may be challenging (also S1pr2 leads to nuclear S1P accumulation – via SphK2 activity –
see Chapter 24). inhibiting histone deacetylase‐regulating enzymes and NRs
25:  Bile Acids as Signaling Molecules 307

involved in nutrient sensing and metabolism. In line with this and BSEP into the plasma membrane of hepatocytes [172].
finding, S1pr2‐knockout mice develop a fatty liver under high‐ TUDCA also raises cAMP levels and thus activates the protein
fat diet [158]. S1pr2 signaling also stimulates phospholipase C kinase A (PKA) protecting hepatocytes against apoptosis by
β (PLCβ), which in turn activates MMP, releasing EGF from the internalizing CD95, thus preventing its interaction and activa-
membrane, activating its receptor EGFR (Figure  25.1) [156, tion with pro‐apoptotic bile acids such as glyco‐CDCA [173].
159]. HDL carries the pro-inflammatory sphingolipid S1P
[160]. Interestingly, bile acid signaling (via FXR) reduces
apoA‐I production and thereby circulating HDL carrying S1P,
which could contribute to anti‐inflammatory bile acid effects by KINASES: PHOSPHATIDYLINOSITOL
inhibiting the S1PR2 pathway. Several S1PR modulators are 3‐KINASE AND PROTEIN KINASE C
currently being developed to treat autoimmune disorders, such
as ozanimod in ulcerative colitis; although these do not target Bile acids such as TCA activate phosphatidylinositol 3‐kinase
S1PR2 [161], a bile acid‐based scaffold might be used to (PI3K), which promotes the insertion of transporters such as rat
­antagonize S1PR2. Mrp2 and Bsep into the canalicular membrane, thereby increas-
ing biliary bile acid excretion [174, 175]. Since TCA activates
FXR as transcriptional regulator of MRP2 (see earlier), it would
Cholinergic receptor muscarinic 3
be of interest to explore potential crosstalks between FXR and
Bile acids can activate the cholinergic receptor muscarinic 3 PI3K in regulation of MRP2 and Bsep.
(CHRM3), a G‐coupled receptor that is naturally activated by TUDCA stimulates calcium‐dependent exocytosis into bile
acetylcholine [162]. Similar to S1PR2, CHMR3 and CHMR1 partially by activating PKCα [176] and enhances biliary excre-
activate PLC [163] and thereby EGFR (Figure 25.1 and see ear- tion by inserting Mrp2 into the canalicular membrane [177],
lier). TCA was found to activate the CHMR2 in rat cardiomyo- which has also been mechanistically linked to PKCα and PKA
cytes, possibly contributing to bile acid‐induced arrhythmia activation [178].
[164]. Moreover, CHMR3 signaling influences bile formation
and cholestasis: Mice lacking Chmr3 have a modestly reduced
bile flow and biliary bicarbonate secretion and absence of Chm3
enhances the susceptibility to cholestatic injury in DDC‐fed CYTOTOXIC EFFECTS OF BILE ACIDS
mice. Moreover, treatment of Mdr2−/− with a CHMR3 agonist
decreased liver injury, indicating that CHMR3 signaling may Bile acid cytotoxicity depends not only on their concentrations
represent a therapeutic target in cholangiopathies [165]. but also on their hydrophobicity [7]. Bile acids can induce cell
injury by stimulating apoptosis via the TNF‐related apoptosis‐
inducing ligand (TRAIL). GCDCA stimulates TRAIL
Epidermal growth factor downstream signaling such as caspases 3, 8, and 10, and phos-
EGF binds and activates the EGFR, stimulating cell growth and phorylates the cellular FLICE inhibitory protein (cFLIP) via
differentiation; as such, EGFR signaling is also involved in hep- PKC, which then is bound to the Fas‐associated death domain
atocarcinogenesis [166]. Hydrophobic bile acids such as DCA protein (FADD) and initiates cell apoptosis [179].
stimulate hepatocyte apoptosis by inducing phosphorylation of Secondary bile acid, such as DCA, activate the transmem-
the death receptor CD95 via EGFR [167]. DCA can directly brane receptor FASR, which results in mitochondrial damage
activate EGFR signaling, probably involving plasma membrane and ROS generation. [180, 181]. DCA activation of PLA2 leads
alterations leading to EGFR activation [168] (Figure 25.1), an to production of arachidonic acid (AA), a pro-inflammatory
effect which is counteracted by UDCA [169]. Bile acid‐induced fatty acid generating ROS, while DCA stimulation of NADPH
EGFR activation in quiescent rat hepatic stellate cells (HSCs) oxidase also results in ROS [180, 181]. Activation of membrane
can trigger both proliferation and apoptosis. EGF stimulation of by DCA leads to inositol 1,4,5‐trisphosphate (IP3) and diacylg-
quiescent primary rat HSCs results in increased migration with- lycerol (DAG) production, which can promote endoplasmic
out proliferation [170], whereas stimulation of EGFR by hydro- reticulum (ER) stress, release of the BCL2‐antagonist/killer 1
phobic bile acids leads to an NADPH oxidase‐driven ROS (BAK) and mitochondrial damage (Figure  25.4) [180, 181].
generation followed by a Yes‐mediated EGFR activation and a Bile acid accumulation in cholestasis also stimulates ERK
shift of the proliferative signal to an apoptotic signal when JNK ­activity, activating the early growth factor (EGR1) leading to
is costimulated [171]. These observations may be important in production of pro-inflammatory cytokines, adhesion mole-
understanding bile acid effects in the development of biliary cule expression, and amino acid synthesis all promoting liver
fibrosis. injury (Figure 25.4) [182].
Finally, bile acid can also induce necrosis. Human hepato-
cytes exposed to G‐CDCA at concentrations found in serum of
patients with cholestasis developed necrosis without increase in
INTEGRINS apoptotic parameters such as cleaved cytokeratin 18. Patients
with cholestasis displayed elevations in serum full‐length
Tauro‐UDCA (TUDCA) also signals via α5β1‐integrins, chang- cytokeratin 18, high mobility group box 1 protein (HMGB1),
ing its β1 subunit conformation to activate the focal adhesion and acetylated HMGB1, all hallmarks of necrotic inflammation
kinases, subsequently increasing choleresis by inserting NTCP [183]. Therefore, in contrast to mouse hepatocytes, human
308 THE LIVER:  REFERENCES

BA

BA

NADPHox PLA2 FASR PLC

ARA CASP8 IP3 ERK

Mitochondrial
ROS BAK ER stress
Cytosol damage

NFκB
Nucleus
EGR1

Inflammatory Inflammatory
genes genes
DNA DNA
damage repair systems

Figure 25.4  Bile acids (BA) induce ER stress, mitochondrial damage, and inflammation. Bile acid (e.g. DCA)‐mediated stimulation of NADPH
oxidase or PLA2 leads to reactive oxygen species (ROS) generation. BA activation of the receptor FASR releases caspase 8 (CASP8), leading to
mitochondrial damage with ROS production. ROS promotes DNA damage and activates NFκB, inhibiting DNA repair and promoting inflamma-
tion. BA stimulation of PLC releases inositol 1,4,5‐trisphosphate (IP3), driving endoplasmic reticulum (ER) stress, release of the BCL2‐antagonist/
killer 1 (BAK), and mitochondrial damage. BA also activate ERK, which stimulates the early growth factor (EGR1) transcriptional activity, which
induces pro-inflammatory cytokine secretion, adhesion molecules, and arachidonic acid synthesis.

hepatocytes are more resistant to human bile acid species and disease. Therefore, understanding bile acid signaling will most
are likely to preferentially die of necrosis. likely be highly relevant for understanding the pathobiology and
management of diseases of the liver and beyond.

CONCLUSIONS
The new insights into the signaling functions of bile acids via a
REFERENCES
broad range of membrane and nuclear receptors has profoundly 1. Claudel, T., Zollner, G., Wagner, M., and Trauner, M. Role of nuclear recep-
changed our understanding of liver physiology and the patho- tors for bile acid metabolism, bile secretion, cholestasis, and gallstone dis-
physiology of liver diseases. Bile acid signaling is not only criti- ease. Biochim Biophys Acta, 2011;1812(8):867–78.
cal for maintaining bile acid homeostasis by providing feedback 2. Hofmann, A.F. and Hagey, L.R. Key discoveries in bile acid chemistry and
biology and their clinical applications: history of the last eight decades. J
and feedforward loops regulating bile acid transport, synthesis,
Lipid Res, 2014;55(8):1553–95.
and metabolism, but as “enterohepatic hormones” bile acids 3. Takahashi, S., Fukami, T., Masuo, Y. et al. Cyp2c70 is responsible for the
also coordinate several key aspects of lipid and glucose metabo- species difference in bile acid metabolism between mice and humans. J Lipid
lism, intestinal microbiota, and even some key aspects of Res, 2016;57(12):2130–7.
inflammation and immunity. Such broad implications of bile 4. Jia, W., Xie, G., and Jia, W. Bile acid‐microbiota crosstalk in gastrointestinal
inflammation and carcinogenesis. Nat Rev Gastroenterol Hepatol,
acid signaling also have important potential implications for
2018;15(2):111–28.
treatment of cholestatic and metabolic liver diseases. However, 5. Wahlstrom, A., Sayin, S.I., Marschall, H.U., and Backhed, F. Intestinal
we have to keep in mind that many of the data rely either on in crosstalk between bile acids and microbiota and its impact on host metabo-
vitro or mouse studies and in vivo data in humans are still scarce. lism. Cell Metab, 2016;24(1):41–50.
Therefore, the next challenge will be to elucidate and integrate 6. Ikegami, T. and Honda, A. Reciprocal interactions between bile acids and gut
microbiota in human liver diseases. Hepatol Res, 2018;48(1):15–27.
the contributions of individual bile acid signaling pathways in 7. Jansen, P.L., Ghallab, A., Vartak, N. et al. The ascending pathophysiology of
human physiology and pathophysiology. Bile acid‐activated G cholestatic liver disease. Hepatology, 2017;65(2):722–38.
protein‐coupled receptors and NRs offer ample opportunities to 8. Forman, B.M., Goode, E., Chen, J. et al. Identification of a nuclear recep-
modulate hepatic metabolism, inflammation, and liver fibrosis tor that is activated by farnesol metabolites. Cell, 1995;81(5): 687–93.
and several ligands have already shown promising effects in 9. Seol, W., Choi, H.S., and Moore, D.D. Isolation of proteins that interact spe-
cifically with the retinoid X receptor: two novel orphan receptors. Mol
preclinical or clinical studies. Moreover, changes in bile acid Endocrinol, 1995;9(1):72–85.
composition and signaling are emerging as important biomark- 10. Makishima, M., Okamoto, A.Y., Repa, J.J. et al. Identification of a nuclear
ers at the interface of host and microbial metabolism in liver receptor for bile acids. Science, 1999;284(5418):1362–5.
25:  Bile Acids as Signaling Molecules 309

11. Parks, D.J., Blanchard, S.G., Bledsoe, R.K. et al. Bile acids: natural ligands 36. Xie, W., Radominska‐Pandya, A., Shi, Y. et al. An essential role for nuclear
for an orphan nuclear receptor. Science, 1999;284(5418):1365–8. receptors SXR/PXR in detoxification of cholestatic bile acids. Proc Natl
12. Wang, H., Chen, J., Hollister, K., Sowers, L.C., and Forman, B.M. Acad Sci U S A, 2001;98(6):3375–80.
Endogenous bile acids are ligands for the nuclear receptor FXR/BAR. Mol 37. Zhang, J., Huang, W., Qatanani, M., Evans, R.M., and Moore, D.D. The con-
Cell, 1999;3(5):543–53. stitutive androstane receptor and pregnane X receptor function coordinately
13. Cui, J., Heard, T.S., Yu, J. et al. The amino acid residues asparagine 354 and to prevent bile acid‐induced hepatotoxicity. J Biol Chem, 2004;279(47):
isoleucine 372 of human farnesoid X receptor confer the receptor with high 49517–22.
sensitivity to chenodeoxycholate. J Biol Chem, 2002;277(29):25963–9. 38. Schuetz, E.G., Strom, S., Yasuda, K. et al. Disrupted bile acid homeostasis
14. Downes, M., Verdecia, M.A., Roecker, A.J. et al. A chemical, genetic, and reveals an unexpected interaction among nuclear hormone receptors, trans-
structural analysis of the nuclear bile acid receptor FXR. Mol Cell, porters, and cytochrome P450. J Biol Chem, 2001;276(42):39411–18.
2003;11(4):1079–92. 39. Uppal, H., Toma, D., Saini, S.P. et  al. Combined loss of orphan receptors
15. Zhao, A., Yu, J., Lew, J.L. et al. Polyunsaturated fatty acids are FXR ligands PXR and CAR heightens sensitivity to toxic bile acids in mice. Hepatology,
and differentially regulate expression of FXR targets. DNA Cell Biol, 2005;41(1):168–76.
2004;23(8):519–26. 40. Kliewer, S.A., Moore, J.T., Wade, L. et al. An orphan nuclear receptor acti-
16. Cieslak, A., Trottier, J., Verreault, M. et al. N‐3 Polyunsaturated fatty acids vated by pregnanes defines a novel steroid signaling pathway. Cell,
stimulate bile acid detoxification in human cell models. Can J Gastroenterol 1998;92(1):73–82.
Hepatol, 2018;2018:6031074. 41. Lehmann, J.M., McKee, D.D., Watson, M.A. et al. The human orphan nuclear
17. Pellicciari, R., Fiorucci, S., Camaioni, E. et al. 6Alpha‐ethyl‐chenodeoxy- receptor PXR is activated by compounds that regulate CYP3A4 gene expres-
cholic acid (6‐ECDCA), a potent and selective FXR agonist endowed with sion and cause drug interactions. J Clin Invest, 1998;102(5):1016–23.
anticholestatic activity. J Med Chem, 2002;45(17):3569–72. 42. Bertilsson, G., Heidrich, J., Svensson, K. et  al. Identification of a human
18. Maloney, P.R., Parks, D.J., Haffner, C.D. et al. Identification of a chemical tool nuclear receptor defines a new signaling pathway for CYP3A induction.
for the orphan nuclear receptor FXR. J Med Chem, 2000;43(16): 2971–4. Proc Natl Acad Sci U S, A, 1998;95(21):12208–13.
19. Baghdasaryan, A., Claudel, T., Gumhold, J. et al. Dual farnesoid X receptor/ 43. Blumberg, B., Sabbagh, W., Jr., Juguilon, H. et al. SXR, a novel steroid and
TGR5 agonist INT‐767 reduces liver injury in the Mdr2‐/‐ (Abcb4‐/‐) mouse xenobiotic‐sensing nuclear receptor. Genes Dev, 1998;12(20):3195–205.
cholangiopathy model by promoting biliary HCO output. Hepatology, 44. Watkins, R.E., Wisely, G.B., Moore, L.B. et al. The human nuclear xenobi-
2011;54(4):1303–12. otic receptor PXR: structural determinants of directed promiscuity. Science,
20. Liu, Y., Binz, J., Numerick, M.J. et al. Hepatoprotection by the farnesoid X 2001;292(5525):2329–33.
receptor agonist GW4064 in rat models of intra‐ and extrahepatic cholesta- 45. Baes, M., Gulick, T., Choi, H.S. et al. A new orphan member of the nuclear
sis. J Clin Invest, 2003;112(11):1678–87. hormone receptor superfamily that interacts with a subset of retinoic acid
21. Nevens, F., Andreone, P., Mazzella, G. et al. A placebo‐controlled trial of obet- response elements. Mol Cell Biol, 1994;14(3):1544–52.
icholic acid in primary biliary cholangitis. N Engl J Med, 2016;375(7):631–43. 46. Goodwin, B., Redinbo, M.R., and Kliewer, S.A. Regulation of cyp3a gene
22. Mudaliar, S., Henry, R.R., Sanyal, A.J. et al. Efficacy and safety of the farnesoid transcription by the pregnane x receptor. Annu Rev Pharmacol Toxicol,
X receptor agonist obeticholic acid in patients with type 2 diabetes and nonalco- 2002;42:1–23.
holic fatty liver disease. Gastroenterology, 2013;145(3):574–82 e571. 47. Willson, T.M. and Kliewer, S.A. PXR, CAR and drug metabolism. Nat Rev
23. Xu, Y. Recent progress on bile acid receptor modulators for treatment of Drug Discov, 2002;1(4):259–66.
metabolic diseases. J Med Chem, 2016;59(14):6553–79. 48. Beilke, L.D., Aleksunes, L., Holland, R. et al. Car‐mediated changes in bile
24. Zavacki, A.M., Lehmann, J.M., Seol, W. et al. Activation of the orphan recep- acid composition contributes to hepatoprotection from Lca‐induced liver
tor RIP14 by retinoids. Proc Natl Acad Sci U S A, 1997;94(15):7909–14. injury in mice. Drug Metab Dispos, 2009;37(5):1035–45.
25. Claudel, T., Sturm, E., Duez, H. et al. Bile acid‐activated nuclear receptor 49. Jones, G., Strugnell, S.A., and DeLuca, H.F. Current understanding of the
FXR suppresses apolipoprotein A‐I transcription via a negative FXR molecular actions of vitamin, D. Physiol Rev, 1998;78(4):1193–231.
response element. J Clin Invest, 2002;109(7):961–71. 50. Makishima, M., Lu, T.T., Xie, W. et al. Vitamin D receptor as an intestinal
26. Barbier, O., Torra, I.P., Sirvent, A. et al. FXR induces the UGT2B4 enzyme bile acid sensor. Science, 2002;296(5571):1313–16.
in hepatocytes: a potential mechanism of negative feedback control of FXR 51. Nishida, S., Ozeki, J., and Makishima, M. Modulation of bile acid metabo-
activity. Gastroenterology, 2003;124(7):1926–40. lism by 1{alpha}‐hydroxyvitamin D3 administration in mice. Drug Metab
27. Chong, H.K., Infante, A.M., Seo, Y.K. et al. Genome‐wide interrogation of Dispos, 2009;37(10):2037–44.
hepatic FXR reveals an asymmetric IR‐1 motif and synergy with LRH‐1. 52. Hollenberg, S.M., Weinberger, C., Ong, E.S. et  al. Primary structure and
Nucleic Acids Res, 2010;38(18):6007–17. expression of a functional human glucocorticoid receptor cDNA. Nature,
28. Sinal, C.J., Tohkin, M., Miyata, M. et al. Targeted disruption of the nuclear 1985;318(6047):635–41.
receptor FXR/BAR impairs bile acid and lipid homeostasis. Cell, 53. Oakley, R.H., Jewell, C.M., Yudt, M.R., Bofetiado, D.M., and Cidlowski,
2000;102(6):731–44. J.A. The dominant negative activity of the human glucocorticoid receptor
29. Kim, I., Ahn, S.H., Inagaki, T. et  al. Differential regulation of bile acid beta isoform. Specificity and mechanisms of action. J Biol Chem,
homeostasis by the farnesoid X receptor in liver and intestine. J Lipid Res, 1999;274(39):27857–66.
2007;48(12):2664–72. 54. Tanaka, H. and Makino, I. Ursodeoxycholic acid‐dependent activation of the
30. Zhang, Y., Kast‐Woelbern, H.R., and Edwards, P.A. Natural structural vari- glucocorticoid receptor. Biochem Biophys Res Commun, 1992;
ants of the nuclear receptor farnesoid X receptor affect transcriptional activa- 188(2):942–8.
tion. J Biol Chem, 2003;278(1):104–10. 55. Weitzel, C., Stark, D., Kullmann, F. et  al. Ursodeoxycholic acid induced
31. Gomez‐Ospina, N., Potter, C.J., Xiao, R. et al. Mutations in the nuclear bile activation of the glucocorticoid receptor in primary rat hepatocytes. Eur J
acid receptor FXR cause progressive familial intrahepatic cholestasis. Nat Gastroenterol Hepatol, 2005;17(2):169–77.
Commun, 2016;7:10713. 56. Calmus, Y., Gane, P., Rouger, P., and Poupon, R. Hepatic expression of class
32. Van Mil, S.W., Milona, A., Dixon, P.H. et al. Functional variants of the cen- I and class II major histocompatibility complex molecules in primary biliary
tral bile acid sensor FXR identified in intrahepatic cholestasis of pregnancy. cirrhosis: effect of ursodeoxycholic acid. Hepatology, 1990;11(1):12–15.
Gastroenterology, 2007;133(2):507–16. 57. Trauner, M., Halilbasic, E., Kazemi‐Shirazi, L. et al. Therapeutic role of bile
33. Heni, M., Wagner, R., Ketterer, C. et al. Genetic variation in NR1H4 encod- acids and nuclear receptor agonists in fibrosing cholangiopathies. Dig Dis,
ing the bile acid receptor FXR determines fasting glucose and free fatty acid 2014;32(5):631–6.
levels in humans. J Clin Endocrinol Metab, 2013;98(7):E1224–9. 58. Takigawa, T., Miyazaki, H., Kinoshita, M. et  al. Glucocorticoid receptor‐
34. Otte, K., Kranz, H., Kober, I. et al. Identification of farnesoid X receptor beta dependent immunomodulatory effect of ursodeoxycholic acid on liver lym-
as a novel mammalian nuclear receptor sensing lanosterol. Mol Cell Biol, phocytes in mice. Am J Physiol Gastrointest Liver Physiol,
2003;23(3):864–72. 2013;305(6):G427–38.
35. Staudinger, J., Liu, Y., Madan, A., Habeebu, S., and Klaassen, C.D. 59. Sharma, R., Prichard, D., Majer, F. et al. Ursodeoxycholic acid amides as
Coordinate regulation of xenobiotic and bile acid homeostasis by pregnane novel glucocorticoid receptor modulators. J Med Chem,
X receptor. Drug Metab Dispos, 2001;29(11):1467–72. 2011;54(1):122–30.
310 THE LIVER:  REFERENCES

60. Inagaki, T., Choi, M., Moschetta, A. et al. Fibroblast growth factor 15 func- 84. Jung, D., Hagenbuch, B., Fried, M., Meier, P.J., and Kullak‐Ublick, G.A.
tions as an enterohepatic signal to regulate bile acid homeostasis. Cell Role of liver‐enriched transcription factors and nuclear receptors in regulat-
Metab, 2005;2(4):217–25. ing the human, mouse, and rat NTCP gene. Am J Physiol Gastrointest Liver
61. Holt, J.A., Luo, G., Billin, A.N. et al. Definition of a novel growth factor‐ Physiol, 2004;286(5):G752–61.
dependent signal cascade for the suppression of bile acid biosynthesis. 85. Zollner, G., Fickert, P., Silbert, D. et  al. Induction of short heterodimer
Genes Dev, 2003;17(13):1581–91. partner 1 precedes downregulation of Ntcp in bile duct‐ligated mice. Am J
62. Babaknejad, N., Nayeri, H., Hemmati, R., Bahrami, S., and Esmaillzadeh, A. Physiol Gastrointest Liver Physiol, 2002;282(1):G184–91.
An overview of FGF19 and FGF21: the therapeutic role in the treatment of 86. Geier, A., Zollner, G., Dietrich, C.G. et al. Cytokine‐independent repres-
the metabolic disorders and obesity. Horm Metab Res, 2018;50(6):441–52. sion of rodent Ntcp in obstructive cholestasis. Hepatology, 2005;
63. Kir, S., Beddow, S.A., Samuel, V.T. et al. FGF19 as a postprandial, insulin‐ 41(3):470–7.
independent activator of hepatic protein and glycogen synthesis. Science, 87. Lee, Y.K., Dell, H., Dowhan, D.H., Hadzopoulou‐Cladaras, M., and Moore,
2011;331(6024):1621–4. D.D. The orphan nuclear receptor SHP inhibits hepatocyte nuclear factor 4
64. Chiang, J.Y. Regulation of bile acid synthesis: pathways, nuclear receptors, and retinoid X receptor transactivation: two mechanisms for repression.
and mechanisms. J Hepatol, 2004;40(3):539–51. Mol Cell Biol, 2000;20(1):187–95.
65. Goodwin, B., Jones, S.A., Price, R.R. et  al. A regulatory cascade of the 88. Eloranta, J.J., Jung, D., and Kullak‐Ublick, G.A. The human Na+‐taurocho-
nuclear receptors FXR, SHP‐1, and LRH‐1 represses bile acid biosynthesis. late cotransporting polypeptide gene is activated by glucocorticoid receptor
Mol Cell, 2000;6(3):517–26. and peroxisome proliferator‐activated receptor‐gamma coactivator‐1alpha,
66. Lu, T.T., Makishima, M., Repa, J.J. et al. Molecular basis for feedback regu- and suppressed by bile acids via a small heterodimer partner‐dependent
lation of bile acid synthesis by nuclear receptors. Mol Cell, mechanism. Mol Endocrinol, 2006;20(1):65–79.
2000;6(3):507–15. 89. Zollner, G., Marschall, H.U., Wagner, M., and Trauner, M. Role of nuclear
67. Hayhurst, G.P., Lee, Y.H., Lambert, G., Ward, J.M., and Gonzalez, F.J. receptors in the adaptive response to bile acids and cholestasis: pathoge-
Hepatocyte nuclear factor 4alpha (nuclear receptor 2A1) is essential for netic and therapeutic considerations. Mol Pharm, 2006;3(3):231–51.
maintenance of hepatic gene expression and lipid homeostasis. Mol Cell 90. Ananthanarayanan, M., Balasubramanian, N., Makishima, M.,
Biol, 2001;21(4):1393–403. Mangelsdorf, D.J., and Suchy, F.J. Human bile salt export pump promoter
68. Kir, S., Zhang, Y., Gerard, R.D., Kliewer, S.A., and Mangelsdorf, D.J. is transactivated by the farnesoid X receptor/bile acid receptor. J Biol
Nuclear receptors HNF4alpha and LRH‐1 cooperate in regulating Cyp7a1 in Chem, 2001;276(31):28857–65.
vivo. J Biol Chem, 2012;287(49):41334–41. 91. Chow, E.C., Durk, M.R., Cummins, C.L., and Pang, K.S. 1α,25‐
69. Kerr, T.A., Saeki, S., Schneider, M. et  al. Loss of nuclear receptor SHP Dihydroxyvitamin D3 up‐regulates P‐glycoprotein via the vitamin D recep-
impairs but does not eliminate negative feedback regulation of bile acid syn- tor and not farnesoid X receptor in both fxr(‐/‐) and fxr(+/+) mice and
thesis. Dev Cell, 2002;2(6):713–20. increased renal and brain efflux of digoxin in mice in vivo. J Pharmacol
70. Wang, L., Lee, Y.K., Bundman, D. et al. Redundant pathways for negative Exp Ther, 2011;337(3):846–59.
feedback regulation of bile acid production. Dev Cell, 2002;2(6):721–31. 92. Keppler, D. and Konig, J. Hepatic secretion of conjugated drugs and endog-
71. Mataki, C., Magnier, B.C., Houten, S.M. et al. Compromised intestinal lipid enous substances. Semin Liver Dis, 2000;20(3):265–72.
absorption in mice with a liver‐specific deficiency of liver receptor homolog 93. Kast, H.R., Goodwin, B., Tarr, P.T. et al. Regulation of multidrug resist-
1. Mol Cell Biol, 2007;27(23):8330–9. ance‐associated protein 2 (MRP2;ABCC2) by the nuclear receptors PXR,
72. Lee, Y.K., Schmidt, D.R., Cummins, C.L. et al. Liver receptor homolog‐1 FXR, and CAR. J Biol Chem, 2001;12:12.
regulates bile acid homeostasis but is not essential for feedback regulation of 94. Wagner, M., Fickert, P., Zollner, G. et al. Role of farnesoid X receptor in
bile acid synthesis. Mol Endocrinol, 2008;22(6):1345–56. determining hepatic ABC transporter expression and liver injury in bile
73. Jung, D. and Kullak‐Ublick, G.A. Hepatocyte nuclear factor 1 alpha: a key duct‐ligated mice. Gastroenterology, 2003;125(3):825–38.
mediator of the effect of bile acids on gene expression. Hepatology, 95. Zollner, G., Fickert, P., Fuchsbichler, A. et  al. Role of nuclear bile acid
2003;37(3):622–31. receptor, FXR, in adaptive ABC transporter regulation by cholic and urso-
74. De Fabiani, E., Mitro, N., Gilardi, F. et al. Coordinated control of cholesterol deoxycholic acid in mouse liver, kidney and intestine. J Hepatol,
catabolism to bile acids and of gluconeogenesis via a novel mechanism of 2003;39(4):480–8.
transcription regulation linked to the fasted‐to‐fed cycle. J Biol Chem, 96. Teng, S., Jekerle, V., and Piquette‐Miller, M. Induction of ABCC3 (MRP3)
2003;278(40):39124–32. by pregnane X receptor activators. Drug Metab Dispos, 2003;
75. Li, T. and Chiang, J.Y. Mechanism of rifampicin and pregnane X receptor 31(11):1296–9.
inhibition of human cholesterol 7 alpha‐hydroxylase gene transcription. Am 97. McCarthy, T.C., Li, X., and Sinal, C.J. Vitamin D receptor‐dependent regu-
J Physiol Gastrointest Liver Physiol, 2005;288(1):G74–84. lation of colon multidrug resistance‐associated protein 3 gene expression
76. Pineda Torra, I., Claudel, T., Duval, C. et al. Bile acids induce the expression by bile acids. J Biol Chem, 2005;280(24):23232–42.
of the human peroxisome proliferator‐ activated receptor alpha gene via acti- 98. Claudel, T., Staels, B., and Kuipers, F. The farnesoid X receptor: a molecu-
vation of the farnesoid X receptor. Mol Endocrinol, 2003;17(2):259–72. lar link between bile acid and lipid and glucose metabolism. Arterioscler
77. Cuperus, F.J., Claudel, T., Gautherot, J., Halilbasic, E., and Trauner, M. The Thromb Vasc Biol, 2005;25(10):2020–30.
role of canalicular ABC transporters in cholestasis. Drug Metab Dispos, 99. Gautier, T., de Haan, W., Grober, J. et al. Farnesoid X receptor activation
2014;42(4):546–60. increases cholesteryl ester transfer protein expression in humans and trans-
78. Li‐Hawkins, J., Lund, E.G., Turley, S.D., and Russell, D.W. Disruption of genic mice. J Lipid Res, 2013;54(8):2195–205.
the oxysterol 7alpha‐hydroxylase gene in mice. J Biol Chem, 100. Claudel, T., Inoue, Y., Barbier, O. et al. Farnesoid X receptor agonists sup-
2000;275(22):16536–42. press hepatic apolipoprotein CIII expression. Gastroenterology,
79. Sayin, S.I., Wahlstrom, A., Felin, J. et al. Gut microbiota regulates bile acid 2003;125(2):544–55.
metabolism by reducing the levels of tauro‐beta‐muricholic acid, a naturally 101. Watanabe, M., Houten, S.M., Wang, L. et al. Bile acids lower triglyceride
occurring FXR antagonist. Cell Metab, 2013;17(2):225–35. levels via a pathway involving FXR, SHP, and SREBP‐1c. J Clin Invest,
80. Hu, X., Bonde, Y., Eggertsen, G., and Rudling, M. Muricholic bile acids are 2004;113(10):1408–18.
potent regulators of bile acid synthesis via a positive feedback mechanism. J 102. Hirokane, H., Nakahara, M., Tachibana, S., Shimizu, M., and Sato, R. Bile
Intern Med, 2014;275(1):27–38. acid reduces the secretion of very low density lipoprotein by repressing
81. Halilbasic, E., Claudel, T., and Trauner, M. Bile acid transporters and regulatory microsomal triglyceride transfer protein gene expression mediated by
nuclear receptors in the liver and beyond. J Hepatol, 2013;58(1):155–68. hepatocyte nuclear factor‐4. J Biol Chem, 2004;279(44):45685–92.
82. Barbier, O., Duran‐Sandoval, D., Pineda‐Torra, I. et al. Peroxisome prolif- 103. Sirvent, A., Claudel, T., Martin, G. et al. The farnesoid X receptor induces
erator‐activated receptor alpha induces hepatic expression of the human bile very low density lipoprotein receptor gene expression. FEBS Lett,
acid glucuronidating UDP‐glucuronosyltransferase 2B4 enzyme. J Biol 2004;566(1–3):173–7.
Chem, 2003;278(35):32852–60. 104. Tregouet, D.A., Konig, I.R., Erdmann, J. et  al. Genome‐wide haplotype
83. Trauner, M. and Boyer, J.L. Bile salt transporters: molecular characteriza- association study identifies the SLC22A3‐LPAL2‐LPA gene cluster as a
tion, function, and regulation. Physiol Rev, 2003;83(2):633–71. risk locus for coronary artery disease. Nat Genet, 2009;41(3):283–5.
25:  Bile Acids as Signaling Molecules 311

105. Chennamsetty, I., Claudel, T., Kostner, K.M. et  al. Farnesoid X receptor 128. Inagaki, T., Moschetta, A., Lee, Y.K. et  al. Regulation of antibacterial
represses hepatic human APOA gene expression. J Clin Invest, defense in the small intestine by the nuclear bile acid receptor. Proc Natl
2011;121(9):3724–34. Acad Sci U S A, 2006;103(10):3920–5.
106. Chennamsetty, I., Claudel, T., Kostner, K.M., Trauner, M., and Kostner, 129. D’Aldebert, E., Biyeyeme Bi Mve, M.J., Mergey, M. et al. Bile salts control
G.M. FGF19 signaling cascade suppresses APOA gene expression. the antimicrobial peptide cathelicidin through nuclear receptors in the
Arterioscler Thromb Vasc Biol, 2012;32(5):1220–7. human biliary epithelium. Gastroenterology, 2009;136(4):1435–43.
107. Papazyan, R., Liu, X., Liu, J. et al. FXR activation by obeticholic acid or 130. Zhang, L., Xie, C., Nichols, R.G. et  al. Farnesoid X receptor signaling
nonsteroidal agonists induces a human‐like lipoprotein cholesterol change shapes the gut microbiota and controls hepatic lipid metabolism. mSystems,
in mice with humanized chimeric liver. J Lipid Res, 2018;59(6):982–93. 2016;1(5).
108. Ma, K., Saha, P.K., Chan, L., and Moore, D.D. Farnesoid X receptor is essen- 131. Kawamata, Y., Fujii, R., Hosoya, M. et  al. A G protein‐coupled receptor
tial for normal glucose homeostasis. J Clin Invest, 2006;116(4):1102–9. responsive to bile acids. J Biol Chem, 2003;278(11):9435–40.
109. Potthoff, M.J., Boney‐Montoya, J., Choi, M. et  al. FGF15/19 regulates 132. Maruyama, T., Miyamoto, Y., Nakamura, T. et al. Identification of mem-
hepatic glucose metabolism by inhibiting the CREB‐PGC‐1alpha pathway. brane‐type receptor for bile acids (M‐BAR). Biochem Biophys Res
Cell Metab, 2011;13(6):729–38. Commun, 2002;298(5):714–19.
110. Zhang, Y., Lee, F.Y., Barrera, G. et al. Activation of the nuclear receptor 133. Maruyama, T., Tanaka, K., Suzuki, J. et al. Targeted disruption of G pro-
FXR improves hyperglycemia and hyperlipidemia in diabetic mice. Proc tein‐coupled bile acid receptor 1 (Gpbar1/M‐Bar) in mice. J Endocrinol,
Natl Acad Sci U S A, 2006;103(4):1006–11. 2006;191(1):197–205.
111. Seyer, P., Vallois, D., Poitry‐Yamate, C. et al. Hepatic glucose sensing is 134. Watanabe, M., Houten, S.M., Mataki, C. et  al. Bile acids induce energy
required to preserve beta cell glucose competence. J Clin Invest, expenditure by promoting intracellular thyroid hormone activation. Nature,
2013;123(4):1662–76. 2006;439(7075):484–9.
112. Bishop‐Bailey, D., Walsh, D.T., and Warner, T.D. Expression and activation 135. Broeders, E.P., Nascimento, E.B., Havekes, B. et al. The bile acid chenode-
of the farnesoid X receptor in the vasculature. Proc Natl Acad Sci U S A, oxycholic acid increases human brown adipose tissue activity. Cell Metab,
2004;101(10):3668–73. 2015;22(3):418–26.
113. Li, Y.T., Swales, K.E., Thomas, G.J., Warner, T.D., and Bishop‐Bailey, D. 136. Thomas, C., Gioiello, A., Noriega, L. et al. TGR5‐mediated bile acid sens-
Farnesoid x receptor ligands inhibit vascular smooth muscle cell inflamma- ing controls glucose homeostasis. Cell Metab, 2009;10(3):167–77.
tion and migration. Arterioscler Thromb Vasc Biol, 2007;27(12):2606–11. 137. Potthoff, M.J., Potts, A., H.e., T. et al. Colesevelam suppresses hepatic gly-
114. Moraes, L.A., Unsworth, A.J., Vaiyapuri, S. et al. Farnesoid X receptor and cogenolysis by TGR5‐mediated induction of GLP‐1 action in DIO mice.
its ligands inhibit the function of platelets. Arterioscler Thromb Vasc Biol, Am J Physiol Gastrointest Liver Physiol, 2013;304(4):G371–80.
2016;36(12):2324–33. 138. Wang, Y.D., Chen, W.D., Yu, D., Forman, B.M., and Huang, W. The G‐pro-
115. Hughes, S.E. Localisation and differential expression of the fibroblast tein coupled bile acid receptor Gpbar1 (TGR5) negatively regulates hepatic
growth factor receptor (FGFR) multigene family in normal and atheroscle- inflammatory response through antagonizing nuclear factor kappaB.
rotic human arteries. Cardiovasc Res, 1996;32(3):557–69. Hepatology, 2011;54(4):1421–32.
116. Cao, F., Wang, S., Cao, X. et  al. Fibroblast growth factor 21 attenuates 139. Yoneno, K., Hisamatsu, T., Shimamura, K. et al. TGR5 signalling inhibits
calcification of vascular smooth muscle cells in vitro. J Pharm Pharmacol, the production of pro‐inflammatory cytokines by in vitro differentiated
2017;69(12):1802–16. inflammatory and intestinal macrophages in Crohn’s disease. Immunology,
117. Zhang, R., Ran, H., Peng, L. et al. Farnesoid X receptor regulates vasoreac- 2013;139(1):19–29.
tivity via angiotensin II type 2 receptor and the kallikrein‐kinin system in 140. Lewis, N.D., Patnaude, L.A., Pelletier, J. et al. A GPBAR1 (TGR5) small
vascular endothelial cells. Clin Exp Pharmacol Physiol, 2016;43(3):327–34. molecule agonist shows specific inhibitory effects on myeloid cell activa-
118. Schwabl, P., Hambruch, E., Seeland, B.A. et al. The FXR agonist PX20606 tion in vitro and reduces experimental autoimmune encephalitis (EAE) in
ameliorates portal hypertension by targeting vascular remodelling and sinu- vivo. PLoS One, 2014;9(6):e100883.
soidal dysfunction. J Hepatol, 2017;66(4):724–33. 141. Pols, T.W., Nomura, M., Harach, T. et al. TGR5 activation inhibits athero-
119. Bozic, M., Alvarez, A., de Pablo, C. et al. Impaired vitamin D signaling in sclerosis by reducing macrophage inflammation and lipid loading. Cell
endothelial cell leads to an enhanced leukocyte‐endothelium interplay: impli- Metab, 2011;14(6):747–57.
cations for atherosclerosis development. PLoS One, 2015;10(8):e0136863. 142. Kida, T., Tsubosaka, Y., Hori, M., Ozaki, H., and Murata, T. Bile acid
120. Molinari, C., Uberti, F., Grossini, E. et al. 1alpha,25‐dihydroxycholecalcif- receptor TGR5 agonism induces NO production and reduces monocyte
erol induces nitric oxide production in cultured endothelial cells. Cell adhesion in vascular endothelial cells. Arterioscler Thromb Vasc Biol,
Physiol Biochem, 2011;27(6):661–8. 2013;33(7):1663–9.
121. Kim, M.S., Shigenaga, J., Moser, A., Feingold, K., and Grunfeld, C. 143. McMahan, R.H., Wang, X.X., Cheng, L.L. et al. Bile acid receptor activa-
Repression of farnesoid X receptor during the acute phase response. J Biol tion modulates hepatic monocyte activity and improves nonalcoholic fatty
Chem, 2003;278(11):8988–95. liver disease. J Biol Chem, 2013;288(17):11761–70.
122. Beigneux, A.P., Moser, A.H., Shigenaga, J.K., Grunfeld, C., and Feingold, 144. Keitel, V., Cupisti, K., Ullmer, C. et  al. The membrane‐bound bile acid
K.R. Reduction in cytochrome P‐450 enzyme expression is associated with receptor TGR5 is localized in the epithelium of human gallbladders.
repression of CAR (constitutive androstane receptor) and PXR (pregnane X Hepatology, 2009;50(3):861–70.
receptor) in mouse liver during the acute phase response. Biochem Biophys 145. Keitel, V., Ullmer, C., and Haussinger, D. The membrane‐bound bile acid
Res Commun, 2002;293(1):145–9. receptor TGR5 (Gpbar‐1) is localized in the primary cilium of cholangio-
123. Pramanik, R., Asplin, J.R., Lindeman, C. et al. Lipopolysaccharide nega- cytes. Biol Chem, 2010;391(7):785–9.
tively modulates vitamin D action by down‐regulating expression of vita- 146. Alemi, F., Kwon, E., Poole, D.P. et al. The TGR5 receptor mediates bile
min D‐induced VDR in human monocytic THP‐1 cells. Cell Immunol, acid‐induced itch and analgesia. J Clin Invest, 2013;123(4):1513–30.
2004;232(1–2):137–43. 147. Cao, W., Tian, W., Hong, J. et al. Expression of bile acid receptor TGR5 in
124. Wang, Y.D., Chen, W.D., Wang, M. et al. Farnesoid X receptor antagonizes gastric adenocarcinoma. Am J Physiol Gastrointest Liver Physiol,
nuclear factor kappaB in hepatic inflammatory response. Hepatology, 2013;304(4):G322–7.
2008;48(5):1632–43. 148. Desai, M.S., Shabier, Z., Taylor, M. et  al. Hypertrophic cardiomyopathy
125. Wang, Y.D., Chen, W.D., Li, C. et  al. Farnesoid X receptor antagonizes and dysregulation of cardiac energetics in a mouse model of biliary fibrosis.
JNK signaling pathway in liver carcinogenesis by activating SOD3. Mol Hepatology, 2010;51(6):2097–107.
Endocrinol, 2015;29(2):322–31. 149. Keitel, V., Gorg, B., Bidmon, H.J. et al. The bile acid receptor TGR5 (Gpbar‐1)
126. Balasubramaniyan, N., Ananthanarayanan, M., and Suchy, F.J. Nuclear factor‐ acts as a neurosteroid receptor in brain. Glia, 2010;58(15):1794–805.
kappaB regulates the expression of multiple genes encoding liver transport 150. Le, Y., Oppenheim, J.J., and Wang, J.M. Pleiotropic roles of formyl peptide
proteins. Am J Physiol Gastrointest Liver Physiol, 2016;310(8):G618–28. receptors. Cytokine Growth Factor Rev, 2001;12(1):91–105.
127. Vavassori, P., Mencarelli, A., Renga, B., Distrutti, E., and Fiorucci, S. The 151. Chen, X., Yang, D., Shen, W. et al. Characterization of chenodeoxycholic
bile acid receptor FXR is a modulator of intestinal innate immunity. J acid as an endogenous antagonist of the G‐coupled formyl peptide recep-
Immunol, 2009;183(10):6251–61. tors. Inflamm Res, 2000;49(12):744–55.
312 THE LIVER:  REFERENCES

152. Ferrari, C., Macchiarulo, A., Costantino, G., and Pellicciari, R. and FAS receptor in primary hepatocytes: inhibition of EGFR/mitogen‐
Pharmacophore model for bile acids recognition by the FPR receptor. J activated protein kinase‐signaling module enhances DCA‐induced apopto-
Comput Aided Mol Des, 2006;20(5):295–303. sis. Mol Biol Cell, 2001;12(9):2629–45.
153. Calmus, Y., Guechot, J., Podevin, P. et al. Differential effects of chenode- 169. Im, E. and Martinez, J.D. Ursodeoxycholic acid (UDCA) can inhibit deox-
oxycholic and ursodeoxycholic acids on interleukin 1, interleukin 6 and ycholic acid (DCA)‐induced apoptosis via modulation of EGFR/Raf‐1/
tumor necrosis factor‐alpha production by monocytes. Hepatology, ERK signaling in human colon cancer cells. J Nutr, 2004;134(2):483–6.
1992;16(3):719–23. 170. Yang, C., Zeisberg, M., Mosterman, B. et al. Liver fibrosis: insights into
154. Calmus, Y., Weill, B., Ozier, Y. et al. Immunosuppressive properties of che- migration of hepatic stellate cells in response to extracellular matrix and
nodeoxycholic and ursodeoxycholic acids in the mouse. Gastroenterology, growth factors. Gastroenterology, 2003;124(1):147–59.
1992;103(2):617–21. 171. Sommerfeld, A., Reinehr, R., and Haussinger, D. Bile acid‐induced epidermal
155. Li, H., Ooi, S.Q., and Heng, C.K. The role of NF‐kappaB in SAA‐induced growth factor receptor activation in quiescent rat hepatic stellate cells can trig-
peroxisome proliferator‐activated receptor gamma activation. Atherosclerosis, ger both proliferation and apoptosis. J Biol Chem, 2009;284(33):22173–83.
2013;227(1):72–8. 172. Gohlke, H., Schmitz, B., Sommerfeld, A., Reinehr, R., and Haussinger, D.
156. Kwong, E., Li, Y., Hylemon, P.B., and Zhou, H. Bile acids and sphingo- alpha5 beta1‐integrins are sensors for tauroursodeoxycholic acid in hepato-
sine‐1‐phosphate receptor 2 in hepatic lipid metabolism. Acta Pharm Sin B, cytes. Hepatology, 2013;57(3):1117–29.
2015;5(2):151–7. 173. Sommerfeld, A., Reinehr, R., and Haussinger, D. Tauroursodeoxycholate
157. Skoura, A., Michaud, J., Im, D.S. et  al. Sphingosine‐1‐phosphate recep- protects rat hepatocytes from bile acid‐induced apoptosis via beta1‐integ-
tor‐2 function in myeloid cells regulates vascular inflammation and athero- rin‐ and protein kinase a‐dependent mechanisms. Cell Physiol Biochem,
sclerosis. Arterioscler Thromb Vasc Biol, 2011;31(1):81–5. 2015;36(3):866–83.
158. Nagahashi, M., Takabe, K., Liu, R. et  al. Conjugated bile acid‐activated 174. Misra, S., Ujhazy, P., Gatmaitan, Z., Varticovski, L., and Arias, I.M. The
S1P receptor 2 is a key regulator of sphingosine kinase 2 and hepatic gene role of phosphoinositide 3‐kinase in taurocholate‐induced trafficking of
expression. Hepatology, 2015;61(4):1216–26. ATP‐dependent canalicular transporters in rat liver. J Biol Chem,
159. Dent, P., Fang, Y., Gupta, S. et al. Conjugated bile acids promote ERK1/2 1998;273(41):26638–44.
and AKT activation via a pertussis toxin‐sensitive mechanism in murine 175. Misra, S., Varticovski, L., and Arias, I.M. Mechanisms by which cAMP
and human hepatocytes. Hepatology, 2005;42(6):1291–9. increases bile acid secretion in rat liver and canalicular membrane vesicles.
160. Sachinidis, A., Kettenhofen, R., Seewald, S. et al. Evidence that lipopro- Am J Physiol Gastrointest Liver Physiol, 2003;285(2):G316–24.
teins are carriers of bioactive factors. Arterioscler Thromb Vasc Biol, 176. Beuers, U., Throckmorton, D.C., Anderson, M.S. et al. Tauroursodeoxycholic
1999;19(10):2412–21. acid activates protein kinase C in isolated rat hepatocytes. Gastroenterology,
161. Dyckman, A.J. Modulators of sphingosine‐1‐phosphate pathway biology: 1996;110(5):1553–63.
recent advances of sphingosine‐1‐phosphate receptor 1 (S1P1) agonists and 177. Beuers, U., Bilzer, M., Chittattu, A. et  al. Tauroursodeoxycholic acid
future perspectives. J Med Chem, 2017;60(13):5267–89. inserts the apical conjugate export pump, Mrp2, into canalicular mem-
162. Raufman, J.P., Chen, Y., Cheng, K. et al. Selective interaction of bile acids branes and stimulates organic anion secretion by protein kinase C‐
with muscarinic receptors: a case of molecular mimicry. Eur J Pharmacol, dependent mechanisms in cholestatic rat liver. Hepatology,
2002;457(2–3):77–84. 2001;33(5):1206–16.
163. Raufman, J.P., Cheng, K., and Zimniak, P. Activation of muscarinic recep- 178. Wimmer, R., Hohenester, S., Pusl, T. et  al. Tauroursodeoxycholic acid
tor signaling by bile acids: physiological and medical implications. Dig Dis exerts anticholestatic effects by a cooperative cPKC alpha‐/PKA‐dependent
Sci, 2003;48(8):1431–44. mechanism in rat liver. Gut, 2008;57(10):1448–54.
164. Sheikh Abdul Kadir, S.H., Miragoli, M., Abu‐Hayyeh, S. et al. Bile acid‐ 179. Higuchi, H., Yoon, J.H., Grambihler, A. et al. Bile acids stimulate cFLIP
induced arrhythmia is mediated by muscarinic M2 receptors in neonatal rat phosphorylation enhancing TRAIL‐mediated apoptosis. J Biol Chem,
cardiomyocytes. PLoS One, 2010;5(3):e9689. 2003;278(1):454–61.
165. Durchschein, F., Krones, E., Pollheimer, M.J. et al. Genetic loss of the mus- 180. Barrasa, J.I., Olmo, N., Lizarbe, M.A., and Turnay, J. Bile acids in the colon,
carinic M3 receptor markedly alters bile formation and cholestatic liver from healthy to cytotoxic molecules. Toxicol In Vitro, 2013;27(2):964–77.
injury in mice. Hepatol Res, 2018;48(3):E68–E77. 181. Bernstein, H., Bernstein, C., Payne, C.M., Dvorakova, K., and Garewal, H.
166. Lanaya, H., Natarajan, A., Komposch, K. et al. EGFR has a tumour‐pro- Bile acids as carcinogens in human gastrointestinal cancers. Mutat Res,
moting role in liver macrophages during hepatocellular carcinoma forma- 2005;589(1):47–65.
tion. Nat Cell Biol, 2014;16(10):972–7. 182. Allen, K., Jaeschke, H., and Copple, B.L. Bile acids induce inflammatory
167. Reinehr, R., Graf, D., and Haussinger, D. Bile salt‐induced hepatocyte genes in hepatocytes: a novel mechanism of inflammation during obstruc-
apoptosis involves epidermal growth factor receptor‐dependent CD95 tive cholestasis. Am J Pathol, 2011;178(1):175–86.
tyrosine phosphorylation. Gastroenterology, 2003;125(3):839–53. 183. Woolbright, B.L., Dorko, K., Antoine, D.J. et al. Bile acid‐induced necrosis
168. Qiao, L., Studer, E., Leach, K. et  al. Deoxycholic acid (DCA) causes in primary human hepatocytes and in patients with obstructive cholestasis.
ligand‐independent activation of epidermal growth factor receptor (EGFR) Toxicol Appl Pharmacol, 2015;283(3):168–77.
Hepatic Adenosine
26 Triphosphate‐Binding
Cassette Transport Proteins
and Their Role in Physiology
Peter L.M. Jansen
Department of Hepatology and Gastroenterology, Academic Medical Center, Amsterdam, The Netherlands
and LiSyM Research Network, University of Freiburg, Germany

HISTORY apparent Tm (transport maximum) which suggests that bile salt


secretion is rate‐limiting and may involve specific transport pro-
Bile flow in the liver is driven by active transport of organic ions teins [8]. Mathematical modeling revealed that these transport
across the canalicular membrane of hepatocytes, followed by processes behave like enzyme‐mediated activities and follow
secretion of water in the bile ductules [1]. Studies in rodents Michaelis–Menten kinetics with a specific Km and Vmax (Km is
identified bile salts as the main drivers of bile secretion [2]. Bile the substrate affinity at half maximal transport rate, Vmax the
flow generated by the active secretion of bile salts, is called maximal transport rate of a substrate).
“bile salt dependent” bile flow [3]. A “bile salt independent” Until 1989, the prevailing view was that bile acid secretion is
fraction of bile flow has been proposed based on linear regres- driven by the electrochemical gradient (−35mV) generated by
sion analysis indicating that when flow and bile salt secretion the Na+K+‐ATPase residing in the basolateral plasma domain of
are extrapolated to zero bile salt secretion, there still is consider- the hepatocyte. A change in paradigm occurred following the
able bile flow. This suggests a flow of bile [3] resulting from demonstration that ATP hydrolysis directly drives transport of
secretion of glutathione and other non‐bile salt organic anions organic anions out of canalicular membrane vesicles [9]. Studies
across the canalicular membrane as well as bicarbonate and with canalicular membrane vesicles revealed that bile acid and
water secretion in the small intrahepatic bile ductules [4]. The organic anion transport is directly dependent on ATP hydrolysis
magnitude of the bile salt independent fraction of bile flow is [10, 11]. In vivo studies in wild type and mutant sheep, monkeys
species dependent and in humans is about one third of the and rats confirmed the presence of distinct ATP‐dependent
total bile flow [5]. However, the concept of “bile salt independent transport processes for bile acids and non‐bile acid organic ani-
bile flow” has been challenged on the basis that extrapolation ons [12–14].
of bile flow and bile acid secretion may not be linear and fails to Proof for the functional activity of more than one canalicular
account for the effect of aquaporins [6]. Intrahepatic water transport protein came from animals with natural or experimen-
secretion was first thought to occur via leaky tight junctions but tally induced mutations. For instance, sheep, rats, and monkeys
discovery of aquaporins revealed that water secretion into bile with genetically impaired canalicular secretion of bilirubin con-
ductules is by aquaporins driven by osmotic pressure differ- jugates showed preserved bile salt and organic cation secretion
ences [7]. [11, 12, 14, 15]. Definitive proof had to wait for the cloning of
Studies in which increasing amounts of bile salts were intra- the affected transporters: ABCC2/MRP2 for non‐bile salt
venously administered to intact animals or into isolated per- organic anions, ABCB11/BSEP (sister of Pgp) for bile salts,
fused livers, showed that bile salt secretion and bile flow ABCB1 (Pgp, MDR1) [16–18], for organic cations and some
increased to a maximum called “biliary secretory maximum” or organic anions. Additional evidence for multiple transport

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
314 THE LIVER:  ABC TRANSPORTERS

proteins was obtained by the targeted deletion of genes encoding flow in the canaliculi is from the pericentral to the periportal
separate proteins for transport of bilirubin, bile salts, and zone, from where bile is delivered to bile ductuli and ducts [23].
phosphatidylcholine (PC) in mice [19–21]. Other studies suggest that canalicular bile does not really flow
but represents a basin in which solutes move by diffusion. In
this model bile only flows in bile ductules and ducts (Vartak and
Hengstler, 2019, personal communication). Computational
PHYSIOLOGY OF BILE FLOW modeling suggests that water is added to bile mainly in the peri-
portal area or in the bile ductules, where the biliary pressure is
The idea that organic ion transport drives bile flow needs refine- low. Low biliary pressure allows the passive flow of water to
ment in view of current insights in the 3D reconstruction of so‐ bile via water channels supported by the osmotic gradient cre-
called liver “lobules”. Liver lobules consist of a number of liver ated by the high concentration of organic solutes in canalicular
acini. A liver acinus is the smallest functional unit of the liver. bile [23].
Acini are composed of hepatocytes, sinusoids, and bile canali- To fully understand this model, the enterohepatic cycle has to
culi. Blood flows from the periportal to the pericentral area of be taken into consideration. Bile salts are reabsorbed in the
the liver acinus and bile from the pericentral to the periportal intestine and delivered to the liver via the portal blood. Studies
area. Whether bile really flows in the bile canaliculus remains using tracer dose radiolabeled taurocholate indicated that uptake
hypothetical until proved by techniques such as dynamic intra- of bile salts mainly occurs in the periportal area of the hepatic
vital microscopy that enable observations at the subcellular acinus. At higher concentrations, pericentral hepatocytes also
level in live tissues. participate in bile acid uptake [29]. In a normal physiological
The hepatic artery and portal vein deliver blood to the liver state, bile salts are delivered to the liver in the micromolar con-
sinusoids. The sinusoidal perfusion of blood from the portal to centration range. At this concentration bile salt uptake is
the central vein passes about 15 hepatocytes per acinus. These expected to engage most hepatocytes in the hepatic acinus.
hepatocytes are separated from the bloodstream by a layer of Transport via ABC transporters is dependent on sufficient sup-
fenestrated endothelium. This makes the liver sinusoids differ- ply of ATP and this is better guaranteed in the oxygen‐rich peri-
ent from capillaries in the general circulation. A single cell portal zone than in the oxygen‐poor pericentral zone of the
RNAseq study revealed that hepatocytes along the axis of the hepatic acinus. Therefore, the main ABC transporter activity is
hepatic acini each have their own gene expression profile and in the periportal oxygen‐rich region of the hepatic acinus. If the
their own metabolic program. From these studies a picture canalicular flow model is accepted it should be realized that the
emerges with endogenous synthesis of chenodeoxycholic acid pericentral‐to‐periportal pressure differences cannot entirely
(CDCA) mainly occurring in centrizonal hepatocytes, while the explain the velocity of canalicular flow, particularly not in the
12α‐hydroxylase CYP8B1, for synthesis of cholic acid (CA), is midzonal area. Experimental evidence indicates that peristaltic
mainly located in the midzonal cells and the bile salt conjugat- contractions probably support canalicular bile flow [23] (see
ing enzyme coenzyme A:amino acid N‐acyltransferase (BAAT) Figure 26.1).
mainly in the periportal cells [22]. The functional consequence
of this is that unconjugated bile salts have to re‐enter the peri-
portal hepatocytes before being secreted as conjugates.
Currently this is merely a hypothesis based on the distribution ABC TRANSPORTERS
of enzymes. Final proof has to wait for detailed molecular meta-
bolic studies in which metabolites are identified in situ at the
Discovery, nomenclature, and structure
subcellular level.
Bile canaliculi are the space between two hepatocytes where Victor Ling’s laboratory in 1976 identified a protein conferring
they form an interconnected mesh within hepatocyte plates. multidrug resistance by supporting the efflux of daunorubicin and
Each canaliculus is surrounded by the plasma membrane a number of other agents in Chinese ovary cells. This 170 kDa
domains of adjacent hepatocytes. Canaliculi can be visualized protein was called P‐glycoprotein (Pgp) [18, 30, 31]. P‐glycopro-
by immune fluorescent microscopy using antibodies directed at tein belongs to the superfamily of ATP‐binding cassette (ABC)
canalicular antigens (dipeptidyl peptidase IV, MDR1/ABCB1, transport proteins. The name ABC transporter indicates an evolu-
BSEP/ABCB11, MRP2/ABCC2) or by dynamic intravital tionary conserved ATP‐binding domain and a transport function
microscopy using fluorescent cholephilic tracers (e.g. bile salt [32, 33]. P‐glycoprotein is now known as ABCB1 or MDR1.
tracers such as cholyl‐lysylfluorescein [CLF] and organic anion ABC transport proteins are expressed in prokaryotes as well
tracers such as 6‐carboxyfluorescein diacetate [6‐CFDA]) [23– as eukaryotes which points to their evolutionary conserved
28]. An ideal fluorescent marker for BSEP/ABCB11‐mediated function as protectors of the interior milieu. For an overview see
transport is not available. The specificity of BSEP does allow Table 26.1.
transport of bile acid analogues with a big fluorescein moiety. The human ABC transporter superfamily comprises 48 genes
However, these labeled molecules are also transported by and proteins, classified in 7 subfamilies [34]. Members of the
MRP2, the non‐bile salt organic anion transporting protein. ABC transport superfamily are expressed in liver, pancreas, kid-
Raman imaging enables the study of transport and flow of natu- ney, intestine, testis, and brain. These are tissues at the boundary of
ral bile salts without the need for fluorescent labeling. the interior milieu and the external environment. ABC transporters
Visualization of bile flow in the canaliculus is challenging. are ATP‐dependent pumps that support the active transport of a
Intravital microscopic studies suggest that the direction of bile wide variety of organic and inorganic anions, metals, peptides,
26:  Hepatic Adenosine Triphosphate‐Binding Cassette Transport Proteins and Their Role in Physiology 315

Chol
CYP7A1 FXR
Bile salts Organic
H2O H2O H2O
CA anions Chol PC CDCA

aquaporins BSEP
MRP2 G5/G8 MDR3 BSEP

Bile ducts Bile

Periportal Pericentral
zone zone
Bile salts Organic Chol PC CDCA
H2O H2O H2O
CA anions

NTCP OATP

EHC: bile salts, C4 Organic anions

Periportal Midzonal Pericentral


Relative transporter 0.73 1.0 2.07
density
Intraluminal ~100 ~2500
pressure (Pa)
Bile velocity 1.86 0.39 0.11
(μm/sec)

Figure 26.1  Contribution of ABC transporters and water channels to bile secretion. A single bile canaliculus extends from the pericentral to
periportal area. According to the computational modeling study of Meyer et al. the relative density of transport proteins in the pericentral area is
about 2.8 times higher than in the periportal area. The calculated intraluminal pressure is about 25 times higher in the pericentral than in the peri-
portal area and the bile flow velocity is 17 times higher in the periportal than in the pericentral area [23]. Abbreviations: OATP, organic anion
transporting polypeptides OATP1B1 and OATP 1B3; NTCP, sodium taurocholate cotransporting polypeptide; EHC, enterohepatic cycle; Chol,
cholesterol; PC, phosphatidylcholine; CDCA, chenodeoxycholic acid, CA, cholic acid.

amino acids, and sugars. In liver, these ATP‐dependent pumps spanning helices linked to an intracellular nucleotide‐binding
transport solutes from liver to bile. In the intestine ABC transport- domain (NBD1 and NBD2) for ATP binding and hydrolysis
ers mainly pump across the mucosa cell membrane toward the (Walker A and B and ABC signature sequence). Together TMD1
intestinal lumen. By this orientation they form an active barrier and TMD2 form a channel with sites for ligand binding and trans-
against the intestinal uptake of solutes, toxins, certain food ingre- port. Upon ATP hydrolysis and ligand binding the channel
dients, and certain anti‐cancer medications (e.g. topotecan). Also, undergoes conformational changes wherein the intracellular
in the blood–brain barrier ABC transport proteins pump agents out ligand‐accepting mode shifts to an extracellular efflux mode [35].
of the brain thus presenting a barrier for uptake of drugs and toxic The molecular structure of the nine‐membered C family of
agents, the so‐called blood–brain barrier. Most ABC transport pro- 190 kDa MRP glycoproteins is similar to that of the B family
teins support the transport of solutes across membranes but there but includes an additional N‐terminal TMD (TMD0) with five
are exceptions. For instance, CFTR (cystic fibrosis transmem- membrane‐spanning helices connected via a linker region to
brane conductance regulator) structurally belongs to the ABC TMD1. ABC C family members differ in the structure of the
transporter family but functionally acts as a chloride channel. TMD0 domain.
CFTR mutations are the cause of cystic fibrosis. Proteins of the G family of “white” proteins only have one
TMD of six membrane spanning helices. These proteins act
as homo‐ (ABCG2) or hetero‐ (ABCG5/G8) dimers and as
Structure functional transporters they resemble the B family proteins.
The molecular structure of the MDR (multidrug resistance) pro- The mouse homologues of ABCB1 (MDR1) are Abcb1a and
teins (ABCB1, ABCB4, ABCB11) comprises two transmem- Abcb2 (Mdr1a and Mdr1b); for ABCB4 (MDR3) this is
brane domains (TMD1 and TMD2) each with six membrane Abcb4 (Mdr2).
Table 26.1  Human hepatic ABC transporters
Gene Protein Chromosome Typical substrates NHR Regulation Clinical phenotype Refs
ABCB1 MDR1 7q21.12 Drugs, chemotherapeutic agents PXR Neurotoxicity (mice and dogs) [77, 78]
ABCB4 MDR3 7q21.12 PC FXR, PPARα PFIC3, LPAC, ICP [21, 57]
ABCB11 BSEP 2q31.1 Bile salts FXR PFIC2, BRIC‐2, ICP [20, 25, 65, 165]
ABCC1 MRP1 16p13.11 LTC4, cyclic nucleotides, drugs PXR Increased chemosensitivity (mice) [109]
ABCC2 MRP2, cMOAT 10q24.2 Bilirubin glucuronides, GSH, glutathione conjugates, LTC4, anti‐HIV drugs FXR, PXR, CAR Dubin–Johnson syndrome [16, 90]
ABCC3 MRP3 7q21.33 Bilirubin glucuronides, bile salts, estradiol‐17 β glucuronide PXR, CAR Increased chemosensitivity (mice) [109, 169]
ABCC4 MRP4 13q32.1 Bile salts, bile salt sulfo‐conjugates, cyclic nucleotides (cAMP, cGMP), PXR, CAR Increased chemosensitivity (mice) [109, 111]
DHEA 3‐sulfate
ABCC6 MRP6 16p13.11 ATP HNF4α Pseudoxanthoma elasticum [116, 117]
ABCC7 CFTR 7q31.2 Chloride — Cystic fibrosis [118]
ABCG2 BCRP 4q22.1 Anticancer drugs, protoporphyrin IX — Diet‐induced phototoxicity (mice) [119]
ABCG5/ — 2p21 Cholesterol — Sitosterolemia, gallstones [41]
G8

c26.indd 316 03-11-2019 19:50:05


26:  Hepatic Adenosine Triphosphate‐Binding Cassette Transport Proteins and Their Role in Physiology 317

Nomenclature G8 is interdependent. Perturbation of any of these transporters


can give rise to unstable bile, prone to gallstone formation.
In this chapter we will follow the international literature in using Although ABCG5/G8 is considered as a “gallstone gene”
capitals for human ABC proteins and capital italics for the ABC (Lith9), malfunction of MDR3 is equally important for gall-
protein encoding genes. For ABC proteins and genes in rodents stone formation as in “low phospholipid‐associated cholelithia-
we will use lower case. sis (LPAC)” [43–45]. LPAC is characterized by intrahepatic
microlithiasis, liver fibrosis, and gallstones.
From multidrug resistance to liver function The function of ABC transporters in the canalicular mem-
brane depends on the cholesterol content of microdomains or
B‐, C‐, and G‐family ABC proteins have originally been discov- “lipid rafts” in which these transporters are located. The stabil-
ered as mediators of multidrug resistance of cancer cells [18, ity of these rafts is perturbed when phospholipid transport is
36]. This is based on findings in the late eighties, indicating that impaired as in MDR3 and ATP8B1 deficiency diseases.
the expression of these proteins in cancer cells increases upon Cholesterol lipid raft microdomains are essential for the proper
exposure to chemotherapeutic agents, thus conferring drug function of BSEP, MRP2, and ABCG2 [46–49].
resistance. These findings were originally met with great enthu-
siasm as this was thought to provide an instrument to deal with
multidrug resistance in cancer. Meanwhile, however, it became ATP8B1
apparent that multidrug resistance in cancer is more complex Although ATP8B1 (FIC1) does not belong to the ABC transport
and does not only depend on the overexpression of a single pro- protein family, it is discussed here since it is important for
tein [37]. Nevertheless, for liver research this burgeoning field understanding the hepatic secretory function. ATP8B1 is a P‐
has been of considerable importance for the identification of type ATPase, encoded by the FIC1 gene. ATP8B1 supports the
transport proteins supporting hepatic secretory function. reverse translocation of phosphatidylserine and phosphatidyle-
MDR1, BSEP, MDR3, MRP2, and ABCG5/G8 are located thanolamine from the outer to the inner leaflet of the canalicular
in  the apical or canalicular membrane of the hepatocyte. membrane domain [50]. FIC1 gene mutations have dramatic
Cell  biological studies have elucidated the complex pathway effects on BSEP and MRP2 function (see Chapter 29). Children
accounting for post-translational regulation and cellular locali- with ATP8B1 deficiency have severe cholestatic liver disease
zation (see Chapter 4). ABCC3 (MRP3) and ABCC4 (MRP4) called progressive familial intrahepatic cholestasis (sporadic
are located in the basolateral membrane of the hepatocyte. PFIC type 1 in non‐Amish populations, endemic Byler’s disease
These proteins are expressed during cholestasis and function to in the Amish population). Characteristically these patients have
protect the hepatocyte against high concentrations of toxic a low serum gamma‐glutamyltransferase activity. Adults, hete-
metabolites by supporting active transport out of the cell into the rozygous for the FIC1 mutation, have a less severe disease
blood. Physiologically this is “reverse” transport. Agents leav- called benign recurrent intrahepatic cholestasis (BRIC type 1)
ing the liver via this route are secreted by the kidney. [51, 52].
ATP8B1−/− mice show an increased bile salt‐induced extrac-
tion of cholesterol into bile with, as a result, a reduced choles-
terol/PC ratio in the canalicular membrane and impaired BSEP
ABC‐MEDIATED TRANSPORT and MRP2 function [46]. Children with PFIC1 temporarily ben-
IN THE LIVER efit from partial external biliary drainage but most eventually
need to be transplanted. However, liver transplantation does not
The apical or canalicular hepatocyte membrane is home to completely resolve the clinical manifestations of this disease as
ABCB1 (P‐glycoprotein, multidrug resistance protein 1 patients suffer from steatosis of the liver graft, diarrhea, pancre-
[MDR1]), ABCB4 (MDR3), ABCB11 (BSEP, bile salt export atic disease, and continue to have poor growth [53]. The effect
protein), ABCG2 (BCRP, breast cancer resistance protein), of external biliary drainage suggests that secondary bile salts
ABC G5/G8 (the cholesterol transport protein), and ABCC2 (deoxy‐ and lithocholic acid), play a role in the functional
(MRP2, multidrug resistance‐associated protein 2, formerly defect. External biliary drainage causes enrichment of the bile
called cMOAT, canalicular multispecific organic anion transporter). salt pool with primary bile salts and this may stabilize the cana-
For normal liver function, MDR3, BSEP, MRP2, and ABCG5/ licular membrane and improve the function of transport
G8 are most relevant as they support the transport of the major proteins.
ingredients of bile: phosphatidylcholine (PC) (MDR3), bile
salts (BSEP), conjugated bilirubin, glutathione and glutathione‐
conjugates (MRP2), and cholesterol (ABCG5/G8). BSEP sup-
ABCB4 (MDR3, Mdr2)
ports the transport of monovalent bile salts and, by action of the The ABCB4/MDR3 gene encodes a 170 kDa MDR3 protein in
solubilizing effect of bile salts at the canalicular membrane, humans and Mdr2 in mice. The function of this apical hepato-
indirectly facilitates the secretion of cholesterol and PC [38]. cyte protein became clear when Mdr2−/− mice became available
MDR3 mediates the translocation of PC from the inner to the [21]. These Mdr2−/− mice have normal bile salt secretion but
outer leaflet of the canalicular membrane [39]. ABCG5/G8 sup- lack PC in bile. This has major consequences for liver and bile
ports the transport of cholesterol and MRP2 the transport of ani- ducts. In normal bile, cytotoxic bile salts are packed in phospho-
ons with at least two negative charges including bilirubin lipid and cholesterol‐containing micelles or vesicles. This pro-
conjugates [40–42]. The function of BSEP, MDR3, and ABCG5/ tects hepatocytes and cholangiocytes from the cytotoxic effect
318 THE LIVER:  ABCB1 (MDR1)

of bile salts. Without phospholipids bile is toxic and this may A mutation affecting BSEP underlies PFIC type 2 in chil-
underlie periductal inflammation and fibrosis as in Mdr2−/− mice dren and benign recurrent intrahepatic cholestasis (BRIC 2) in
and children with MDR3 deficiency (PFIC type 3). The histopa- adults heterozygous for ABCB11/BSEP gene mutations [25,
thology in Mdr2−/− mice bears some resemblance to the 65, 66] (see Chapter  29). In PFIC2 the accumulation of bile
pathology in primary sclerosing cholangitis (PSC). PSC is salts causes inflammation and fibrosis of the liver and is asso-
characterized by concentric periductular fibrosis, biliary cirrho- ciated with an increased risk for hepatocellular carcinoma
sis, and an increased risk for cholangiocellular cancer. However, [67]. Typically children with PFIC2 develop cholestasis with a
apart from single nucleotide polymorphisms functional ABCB4/ low serum gamma‐glutamyltransferase activity. Partial exter-
MDR3 gene defects have not been detected in PSC patients nal biliary diversion provides temporary relief particularly in
[54, 55]. patients with some residual BSEP expression. Most patients
Mutations of the ABCB4/MDR 3 gene underlie PFIC type 3 eventually need a liver transplantation. After liver transplanta-
in children, low‐phospholipid‐associated cholelithiasis (LPAC) tion patients usually do well and show a catch‐up growth [53].
and selected cases of intrahepatic cholestasis of pregnancy Recurring PFIC2 has been described after liver transplantation
(ICP) in adults [43, 56–60]. PFIC type 3 is a pediatric disease due to development of BSEP‐directed antibodies [68, 69].
characterized by cirrhosis, bile duct abnormalities, severe pruri- Patients with residual BSEP activity may respond to ursode-
tus, and elevated serum gamma‐glutamyltransferase. PFIC type oxycholic acid therapy. A future option may be treatment with
3 in humans is associated with a more severe histopathology chaperone molecules that help reparation of misfolded BSEP
than that seen in Mdr2−/− mice. This is probably due to human protein and insertion into the apical membrane [70]. This
bile being more cytotoxic than mouse bile with more hydro- would help a subgroup of patients.
philic bile salts (e.g. α‐ and β‐muricholic acid). Ursodeoxycholic BSEP supports the transport of taurine‐ and glycine‐con-
acid administration provides temporary relief in a subset of patients jugated bile salts, unconjugated bile salts, and ursodeoxy-
but most patients eventually need a liver transplantation. cholic acid. Nor‐ursodeoxycholic acid and obeticholic acid
The MDR3 protein plays an important role in bile physiol- may also be BSEP substrates but this has not been tested yet.
ogy. It acts together with BSEP and ABCG5/G8 in producing BSEP has a preference for transporting bile salts but shows
PC‐, bile salt‐, and cholesterol‐containing vesicles at the outer low activity with pravastatin as substrate [71]. Cyclosporin
leaflet of the canalicular membrane. Mdr2−/− mice produce bile A, troglitazone, bosentan, glibenclamide, and rifampicin are
with bile salts in normal concentration but without PC and BSEP inhibitors and may cause drug‐induced liver disease
reduced cholesterol. This shows that a normal MDR3 function [72–74].
not only is required for PC secretion but cholesterol secretion
indirectly also depends on PC secretion [38]. Bile salts and PC
are needed for cholesterol solubilization. Hence a decreased PC
concentration in bile represents a risk for gallstone and biliary ABCB1 (MDR1)
sludge formation.
Mdr2 expression in mice is increased by treatment with MDR1/ABCB1 confers multidrug resistance in cancer. In the
PPARα agonists [61]. Thus PPARα agonists are expected to liver ABCB1 is located in the hepatocyte canalicular mem-
increase PC secretion in bile and therefore this may represent a brane domain where it supports the biliary secretion of
“bile detoxifying” therapy that may be of therapeutic value. organic cations, hormones, and drugs [75, 76]. ABCB1 is
Norursodeoxycholic acid may also be a useful drug as it also expressed in epithelial cells of the bile duct, small intes-
increases bicarbonate secretion thus protecting bile duct epithe- tine, kidney, blood–brain barrier, and placenta. ABCB1 and
lium against bile salt toxicity [62]. ABCB11 show structural similarity but considerable func-
tional differences. While ABCB1 mediates the transport
of a great variety of substrates (analgesics, anti‐arrhythmic
Bile salt export pump, ABCB11 (BSEP) agents, antibiotics, anti‐cancer agents, antihistamines, anti‐
A protein originally called “sister of P‐glycoprotein (SPGP)” epileptic drugs, immunosuppressive agents, antilipidemic
was identified as the principal transporter for glycine‐ and tau- agents), ABCB11 supports the transport of bile salts with a
rine‐conjugated bile salts. SPGP or BSEP is located in the cana- preference for conjugated bile salts [35]. The role of ABCB1
licular membrane domain of hepatocytes [17]. Sister of in normal liver physiology is unclear. Natural mutations of
P‐glycoprotein is generically called bile salt export pump the Abcb1 gene in dogs or targeted deletion of Abcb1a in
(BSEP) and formally ABCB11. Children with BSEP deficiency, mice cause sensitivity of the brain to drugs (e.g. ivermectin),
caused by ABCB11 mutations, have a severe cholestatic liver in otherwise healthy dogs and mice, suggesting a role of this
disease called PFIC type 2. In mice the targeted deletion of the transporter in the blood–brain barrier [77, 78].
Abcb11 gene causes cholestasis albeit not as severe as in humans MDR1 may serve as a backup transport system. For instance
[63]. This is due to the extensive tetrahydroxylation of bile acids in Abcb11/Bsep−/− mice, Mdr1a expression is increased and sup-
and their subsequent renal and biliary excretion via upregulated ports transport of tetrahydroxylated bile salts [20]. Because of
Mdr1a and Mrp2. Tetrahydroxylation of bile salts is a pathway this detoxifying pathway, Abcb11/Bsep−/− mice do not have a
available in mice but not in humans [64]. Definite proof for real cholestatic phenotype. The varied phenotype in patients
BSEP being a bile acid transporter came from studies by Gerloff with progressive familial intrahepatic cholestasis (PFIC type 2)
et al. who showed bile salt transport in Xenopus laevis oocytes may be related to the expression level of MDR1 [79]. In hepato-
and Sf9 cells upon expression of the mouse Abcb11 gene [17]. cellular carcinoma (HCC) MDR1 expression is variable and
26:  Hepatic Adenosine Triphosphate‐Binding Cassette Transport Proteins and Their Role in Physiology 319

inversely correlated with disease free survival and response to glucuronide‐conjugates. MRP1 expression in normal liver is
chemotherapy [80]. marginal but is increased in severe hepatic failure, in hepatic
progenitor cells, and in reactive hepatic ductules [99, 100].
Deletion of MRP1 has not been linked to a particular liver dis-
ABCC2 (MRP2) ease. MRP1 plays a role in the natural course of cancer and is
The multidrug resistance‐associated proteins (MRPs) are 190 responsible for multidrug resistance in some patients. Among
kDa glycoproteins expressed in various organs including liver, the MRPs, MRP1/ABCC1 has the highest homology with
MRP2 is encoded by the ABCC2 gene. Compared to proteins CFTR/ABCC7 and may act as a modifier gene in cystic fibrosis
of the ABCB family (the MDRs) the ABCC family proteins [101]. While MRP2 is exclusively present in the canalicular
(the MRPs) have an extra transmembrane domain comprised domain of hepatocytes and the apical domain of intestinal
of five helices connected by a linker region to the rest of the mucosa and tubular epithelial cells, MRP1 is located in the
protein. The function of this extra domain is unclear since basolateral membrane. This difference in localization is due to a
without this domain MRP1 is fully active and retains its mem- PDZ motif in the C‐terminus of the MRP2 protein that is lack-
brane localization [81]. ing in MRP1[81, 102]. This motif interacts with other PDZ‐con-
MRP2 is the apical transporter of amphiphilic “multivalent” taining proteins, directing MRP2 to the apical membrane.
organic anions, mainly agents with more than one negative MRP1 expression in HCC is associated with a more aggressive
charge. Thus conjugated (mono‐ and diglucuronidated) biliru- phenotype [103].
bin, oxidized and reduced glutathione (GSSG and GSH), MRP3 is an inducible ATP‐dependent transporter of organic
numerous drug conjugates, tetra‐hydroxylated, sulphated and anions in the basolateral membrane of hepatocytes, bile duct
divalent glucuronidated bile salts, leukotriene C4, conjugated and intestine epithelial cells, adrenals, pancreas, kidney, and
hormones, endothelin I, and vasopressin are MRP2/mrp2 sub- prostate [100, 104–106]. MRP3 is upregulated when canalicular
strates [15, 82–86]. Positively charged agents (e.g. metals) are MRP2 is not or poorly expressed as in Dubin–Johnson syn-
transported by MRP2 complexed to glutathione [87]. drome or when transport across the canalicular membrane is
In bile, bile salts and glutathione are the solutes with the impaired as in cholestasis [104, 105, 107]. The substrate spec-
highest concentration (millimolar range). Thus, glutathione acts trum of MRP3 is similar but not identical of that of MRP2 [108].
as an osmolyte contributing to the “bile‐salt independent bile MRP3 transports bilirubin mono‐ and diglucuronide,
flow”. In TR− rats, with a genetic deletion of MRP2, bile flow is estradiol‐17β‐glucuronide, glutathione‐conjugates, bile salts
reduced to ~50% [13, 15]. and their conjugates. In contrast to MRP2, MRP3 does mediate
The MRP2 protein is expressed in the liver, intestine, renal the transport of reduced or oxidized glutathione and has a low
proximal tubule, gallbladder, and placenta [88, 89]. The absence affinity for LTC4 [109]. MRP3 and/or MRP4 are highly
of Mrp2 caused by natural mutations in TR− rats, monkeys, and expressed in hepatic progenitor cells [99, 100].
sheep is associated with conjugated hyperbilirubinemia and a Like MRP3, MRP4 is an inducible ATP‐dependent trans-
phenotype similar to the human Dubin–Johnson syndrome porter in the basolateral membrane of hepatocytes. MRP4 is
(DJS). Like humans with DJS, TR− rats have a brown‐black expressed in liver and proximal tubules of the kidney. MRP4
lysosomal pigment in the liver upon feeding a tyrosine‐, tryp- supports transport of conjugated, sulfated, and unconjugated
tophane‐, and phenylalanine‐enriched diet [13, 16, 90, 91]. The bile salts, cyclic nucleotides (cAMP), LTC4, dehydroepiandros-
targeted deletion of Abcc2 in mice causes mild conjugated terone (DHEA)‐3‐sulphate, estradiol‐17β‐glucuronide, and
hyperbilirubinemia with defective biliary secretion of glu- cyclic nucleotides (cAMP) [110, 111]. MRP4 protein expres-
tathione and organic anions [92]. MRP2 is expressed in most sion in rodent and human liver is induced by cholestasis [112–
HCCs [80, 93, 94]. 114]. Basolateral MRP3 and MRP4 take over the function of
MRP2 and BSEP when the function of these canalicular trans-
porters is impaired as in cholestasis. They support “reverse”
Jaundice transport from hepatocyte to blood leading to elevated serum
Jaundice is the clinical hallmark of cholestatic liver disease. On levels of conjugated bilirubin and bile salts.
a molecular level jaundice may be caused by MRP2 dysfunc-
tion, be it genetic or acquired. Jaundice of sepsis is a typical
example of acquired jaundice by reduced MRP2 expression.
ABCC6, MRP6
Inflammatory cytokines suppress MRP2 expression to a level There is strong genetic linkage between the more than 300
that impairs the hepatobiliary transport with conjugated hyper- ­identified mutations of the ABCC6 gene and pseudoxanthoma
bilirubinemia as a result [95, 96]. MRP2 downregulation may elasticum, a rare disease with abnormalities in skin, retina, car-
be the underlying mechanisms of jaundice in a variety of liver diovascular system, and gastrointestinal system [115]. ABCC6/
diseases [97, 98]. MRP6 is a transmembrane transporter of the ABC C family
expressed in liver and kidney. Pseudoxanthoma elasticum is
characterized by increased mineralization of connective tissue
ABCC1, 3, 4 (MRP 1, 3, 4) and arteries. The relation between disease manifestations and
MRP1 is encoded by the ABCC1 gene. MRP1 is present in the the gene defect has long been elusive. In particular the relation
basolateral membrane of most epithelial cells where it supports between the gene product in the liver, and increased mineraliza-
the transmembrane transport of leukotriene C4 (LTC4), reduced tion in almost all tissues, was puzzling. Recent literature indi-
and oxidized glutathione, and numerous drugs, glutathione‐ and cates that ABCC6/MRP6 supports transport of ATP [116, 117].
320 THE LIVER:  REGULATION OF ABC TRANSPORTER EXPRESSION

ATP is degraded to inorganic pyrophosphate PPi, a potent inhib- preventing accumulation in the brain. This indicates that muta-
itor of mineralization. Reduced PPi in the circulation leads to tions of ABC transporters can affect the clearance of endoge-
the precipitation of calcium‐phosphate in arteries, skin, and nous and exogenous agents and, depending on the substrate,
retina. Oxidative stress in NASH may reduce HNF4α‐mediated allows the intestinal absorption or penetration of the blood–
expression of MRP6 and thereby reduce ATP release and stimu- brain barrier by agents that in normal wild‐type animals would
late mineralization of the arteries leading to coronary athero- not be absorbed or cross the blood–brain barrier.
sclerosis, a prominent cause of death in NASH [115]. ABCG5/G8 are also known as gallstone susceptibility
genes (Lith9 gene in mice). In human populations the ABCG5‐
R50C and ABCG8‐D19H gene variants are associated with
ABCC7 gallstones [123]. The ABCG8‐D19H variant, when tested in
ABCC7 deserves special mention. ABCC7 is the cystic fibrosis a  3H‐cholesterol export assay, showed increased transport
transmembrane conductance regulator (CFTR) that, when activity. The association between the D19H variant and an
mutated, underlies cystic fibrosis [118]. This protein is excep- increased gallstone risk (odds ratio 2.2; P = 1.4 × 10−14) has
tional as it is not an ATP‐dependent pump but a channel that been confirmed in German, Chilean, Chinese, and Indian
allows the cellular efflux of chloride and bicarbonate. CFTR is ­populations [124].
expressed in lungs, pancreas, bile ducts, intestine, and vas defer-
ens. The molecular structure shows two transmembrane domains
each with six membrane‐spanning helices, two nucleotide bind-
ing domains, and a regulatory “R” domain activated by cAMP. REGULATION OF ABC TRANSPORTER
An inside negative electrochemical gradient drives chloride EXPRESSION
efflux out of the cell. CFTR constitutes a gated channel that can
open and close upon ATP binding.
Transcriptional regulation
ABC transporter expression at the apical membrane can change
ABCG2 rapidly or more gradually. Transcriptional and post‐transcrip-
ABCG2 mediates transport of a number of chemotherapeutic tional mechanisms underlie these regulations. A number of
agents (hence the name breast cancer resistance protein 1), elegant studies has shown that ABC transporter expression is
estradiol, progesterone, and chlorophyll metabolites. ABCG2−/− regulated by hepatocyte nuclear factor 4α (HNF4α), nuclear
mice on an alfalfa‐containing diet accumulate phototoxic chlo- hormone receptors (NHRs such as FXR [farnesoid X‐receptor]
rophyll metabolites [119]. and PXR [pregnane X‐receptor], PPARα and CAR [constitu-
Bcrp1−/− mice show increased topotecan serum levels after tive androstane X‐receptor] [125–127 ]) (see Figure 26.2). The
oral administration of this anticancer drug showing that in nor- NHRs act as intracellular sensors for bile salts, endogenous
mal mice Bcrp1 in the intestine acts as a barrier for orally metabolites, and xenobiotics. They reside in the cytosol and
administered drugs (transporting drugs out of the organism when bound to their respective ligands they bind to the DNA of
toward the intestine). In the liver, ABCG2, MRP3, and MRP4 target genes. In general they function to maintain intracellular
are expressed in progenitor cells and may serve a cytoprotective homeostasis by increasing outward transport and reducing
function [99, 100, 120]. uptake. Bile salts provide a clear example. When intracellular
bile salt concentrations increase due to cholestasis, the expres-
sion of the FXR‐dependent bile salt transporter OSTαβ as well
ABCG5/G8 as MRP3 and MRP4 at the basolateral domain increases [104,
ABCG5/G8 expression in liver and intestine is important for 107, 112, 114, 128]. MRP3 supports the cellular efflux of con-
cholesterol transport. In the liver ABCG5/G8 supports the jugated bilirubin from hepatocyte to blood thereby contributing
biliary secretion of cholesterol, and in the small intestine to elevated levels of conjugated bilirubin in serum and clinical
the protein mediates the efflux of cholesterol and plant ­sterols jaundice [129]. MRP4 supports the cellular efflux of conju-
thereby inhibiting their absorption. ABCG5 and G8 are gated and unconjugated bile salts and bile salt sulfate conju-
half  transporters, both proteins are needed for transport gates. MRP4−/− mice are more vulnerable and sensitive to
­function. Patients with mutations in the ABCG8 gene develop cholestasis [130].
­hypercholesterolemia, premature coronary atherosclerosis, and FXR acts as a bile acid sensor that is activated upon bile
phytosterolemia but in meta‐analysis the association with salt binding [131, 132]. This causes upregulation of SHP
hypercholesterolemia was found not to be strong [41, 121]. (short heterodimer partner), a protein that blocks the LRH‐1
Targeted deletion of the Abcg5 and Abcg8 genes in mice causes and HNF4α‐mediated regulation of CYP7A1, an enzyme
impaired cholesterol secretion in bile and increased sensitivity responsible for bile salt synthesis [133]. In addition, FXR
toward changes in the cholesterol content of their diet. On the activation causes upregulation of BSEP, MRP3 and 4, and
other hand, transgenic expression of ABCG5 and G8 causes downregulation of NTCP [134]. Also the expression of OSTαβ,
high cholesterol secretion in bile and lowered absorption of a non‐ATP‐dependent basolateral transporter, is increased in
­cholesterol in the intestine [122]. cholestasis in humans [135].
The function of the blood–brain barrier largely depends on The importance of FXR for biliary transport becomes appar-
the presence of ABC transporters. ABC transporters actively ent when studying the phenotype of patients with genetic FXR
pump out neurotoxic compounds (e.g. ivermectin) thus defects. Homozygous mutations of the NR1H4 gene, encoding
26:  Hepatic Adenosine Triphosphate‐Binding Cassette Transport Proteins and Their Role in Physiology 321

Function
Chemotherapeutics Bilirubin glucuronides
Cyclosporin A Bile salts GSH/GSSG
(Bile salts) PC UDCA Organic anions Cholesterol Chemotherapeutics

MDR1 MDR3 BSEP MRP2 G5/G8 BCRP

Regulation

AhR
PXR MDR1 PPARα MDR3 FXR BSEP PXR MRP2 LXR G5/G8 BCRP

CAR ER

Fenofibrate CDCA Rifampicin Oxysterols Estradiol


Rifampicin
Benzofibrate Obeticholic acid Dexamethasone T0901317 Progesteron
Dexamethasone
PX-104 GW3965
miR 331-5 p miR 33 miR 328
Phenobarbital
miR 451 TCDD miR 181a
miR 298 miR 379 Benzopyrene miR 487a
miR 27a miR 133a miR 212
miR 122 miR 297 miR 200c
miR 137 miR 520h
miR 200c miR 519c
miR 1253
mR 145

Figure 26.2  Function and regulation of canalicular ABC transporters. Upper panel: Function of canalicular ABC transporters depicted with lead-
ing substrates. Lower panel: Regulation of ABC transporters by nuclear hormone receptors and inhibitory activity by miRNA [35, 144, 170–172].

FXR, causes severe neonatal cholestasis with undetectable hepatocellular secretory failure”, a clinical condition wherein
BSEP and preserved MDR3 expression. The disease in these intrahepatic cholestasis persists despite the removal of bile
children is characterized by low serum gamma‐glutamyltrans- duct obstruction or the removal of cholestasis‐inducing medi-
ferase activity, severe jaundice, vitamin K dependent coagulop- cation [141].
athy, and elevated α‐fetoprotein [136]. This shows that FXR not
only regulates BSEP expression at increased bile salt levels but
also plays a role in the household regulation of BSEP.
MicroRNA
Regulatory adaptations in patients with progressive famil- MicroRNAs are small non‐coding RNA molecules that
ial intrahepatic cholestasis (PFIC2) provide examples of originate as small transcripts of hundreds of nucleotides.
cytoprotection when export of toxic solutes is perturbed. In These single‐stranded RNAs interfere with gene transcrip-
PFIC2 BSEP function is reduced or absent, therefore normal tion. MicroRNAs are studied as therapeutic targets to coun-
bile salt secretion across the canalicular membrane is not teract ABC‐transporter‐mediated multidrug resistance in
possible and the intracellular bile salt concentration is ele- cancer chemotherapy [93]. However, single microRNAs
vated. Bile salts bind FXR and this activates the transcrip- target multiple genes. Applications of microRNA (e.g. gene
tion of the ABCC4 gene. MRP4 and OSTαβ at the basolateral silencing) may therefore be complicated by bystander
membrane are upregulated where they support the efflux of effects [142].
bile salts from liver to blood thereby lowering the intracel- Haenisch et  al. reported that microRNA 379 inhibits
lular bile salt level. rifampicin‐induced MRP2 expression by post‐transcriptional
FXR is regulated mainly by conjugated primary bile salts mechanisms [143]. In addition the expression of ABCB1,
while PXR is regulated by non‐bile salt ligands (e.g. rifampicin, BCRP/ABCG2, and MRP1/ABCC1 is affected by microRNAs
corticosteroids) and hydrophobic secondary bile salts. CAR [144]. Although this is early work, eventually microRNAs may
binds bilirubin and phenobarbital; VDR hydrophobic bile salts be therapeutic targets that can be used in the treatment of chole-
and 1,25‐dihydroxyvitamin D3 [125, 137–140]. This knowl- static liver disease. For instance, studies in miR 155−/− mice
edge translates into clinical medicine. Phenobarbital lowers show that miR 155 protects against acetaminophen‐induced
serum bilirubin levels by increasing the activity of the conju- liver injury in mice [145]. The use of microRNAs as biomarkers
gating enzyme UGT1A1 and the transporter MRP2; rifampicin in drug‐induced cholestasis and intrahepatic cholestasis of
is used to enhance MRP2‐mediated transport in “persistent pregnancy is debated [146–148].
322 THE LIVER:  CONCLUSION

Post‐transcriptional regulation anabolic steroid cause “bland cholestasis” and clavulanic acid,
fluoroquinolones, ketoconazole, itraconazole, diclofenac, and
Transcriptional regulatory changes typically take minutes or tricyclic antidepressants are known to cause cholestatic hepati-
longer, post‐transcriptional regulations occur faster. In hepato- tis. MRP2 inhibitors giving rise to “bland jaundice” (conjugated
cytes the BSEP protein is located in a subapical vesicular com- hyperbilirubinemia with or without mild AST/ALT elevation)
partment from where it shuttles to and from the apical membrane include the anticancer drug sorafenib and the antivirals saquina-
[149]. This process is stimulated by ursodeoxycholic acid, hor- vir, ritonavir, tenofovir, emtricitabine, and lamivudine. It is not
mones, oxidative stress, and cell swelling [150–153]. Proteins always clear if drug‐induced cholestasis is due to interactions at
that tether ABC transporters to the cytoskeleton play a role in the level of canalicular ABC transporters, drug‐metabolizing
providing an anchor for these proteins that determine their loca- enzymes, or due to immune‐mediated effects. This limits the
tion in the canalicular membrane [154]. Studies in rat liver and predictive power of in vitro drug toxicity models and makes
Madin–Darby canine kidney cells by Ortiz et al. indicate that drug‐induced cholestasis unpredictable. This poses a health haz-
HAX‐1 and the actin‐protein cortactin are involved in the retrac- ard and a considerable (financial) risk for drug development.
tion of BSEP, MDR1, and MDR2 from the plasma membrane to Interaction of drugs with ABC transporters in vivo and in ani-
the cell interior, presumably the subapical compartment. HAX‐1 mal experiments shows a complicated pattern. For instance,
binds to the linker domain of BSEP [155]. HAX‐1 and cortactin studies in BSEP‐expressing Sf9 cells showed that ATP‐depend-
are known to participate in clathrin‐mediated endocytosis sug- ent bile salt transport is inhibited by cyclosporin A, rifamycin,
gesting that this mechanism plays a role in the retraction of rifampicin, and glibenclamide. This is called cis‐inhibition
BSEP and the MDR proteins from the apical membrane. (drugs acting on the cellular inside). E2 17βG only inhibits
During bile‐duct ligation‐ or estradiol 17β glucuronide (E2 BSEP‐mediated bile salt transport when MRP2 is also expressed.
17βG) ‐induced cholestasis in rats, the ABC transporters BSEP E2 17βG first has to be transported by MRP2 before it inhibits
and MRP2 have been shown to be redirected from their apical BSEP. This is called transinhibition (inhibition by interacting
orientation towards intrahepatic compartments by clathrin‐ with the transport protein on the canalicular side) [162].
mediated endocytosis [156, 157]. Underlying mechanisms for Cyclosporin A, troglitazone, and bosentan are BSEP inhibitors
BSEP and MRP2 relocalization in cholestasis differ. For its and cause cholestatic injury in humans [163]. The liver injury
localization in the apical domain Mrp2 needs to be associated caused by these drugs most probably results from the cytotoxic-
with a protein call radixin. Radixin belongs to a group called ity of accumulated bile salts. Comparing the severe clinical phe-
ERM proteins (ezrin, radixin, moesin). These proteins cross‐ notype of patients with PFIC2 and PFIC3 and the mild phenotype
link integral membrane proteins to actin filaments in the of patients with Dubin–Johnson syndrome allows the prediction
cytoskeleton. Radixin−/− mice have conjugated hyperbilirubine- that drugs interfering with BSEP or MDR3 are more likely to
mia because of a disturbed orientation of Mrp2 [158]. induce liver injury than drugs interfering with MRP2 function.
Interestingly the localization of BSEP and MDR1 is not dis- Since systematic studies are lacking it is unclear how often
turbed in these radixin−/− mice. Studies in rats and humans, have drug‐induced cholestasis is due to genetic variants of ABCB11
shown that in cholestasis MRP2 and radixin do not colocalize and ABCB4. Both genes have been confirmed to be associated
anymore. As a result of this disconnection MRP2 loses its local- with drug‐induced cholestasis and with intrahepatic cholestasis
ization in the canalicular membrane [159, 160]. In cholestasis, of pregnancy [164–167]. Vanishing bile duct syndrome is the
MRP2 binds to ezrin and this results in the redirection of MRP2 most severe and potentially irreversible form of drug‐induced
to intracellular vesicles where the MRP2 protein is ubiquit- liver injury. Implicated drugs are amoxycillin–clavulanic acid,
inated and subjected to proteosomal breakdown. The phospho- flucloxacillin, diclofenac, and ibuprofen [168]. How (pro-
rylation of ezrin at a critical domain (T567) by protein kinase C longed) inhibition of ABC‐protein‐mediated transport leads to
α, δ, and ε is key to this mechanism [161]. disappearance of bile ducts is unclear.
The MDR1, MDR3, and BSEP proteins differ from MRP2 in
that MRP2 contains a PDZ domain at the C‐terminus that deter-
mines its localization in the apical membrane by interacting
with other PDZ domain containing proteins. The MDR/ABCB CONCLUSION
proteins do not have this PDZ domain. MDR1 and BSEP recy-
cle between similar subapical and apical compartments and in ATP‐dependent ABC transporters support the secretory func-
case of BSEP deficiency MDR1 is often also mislocalized con- tion of the liver. Genetic or acquired disturbances of ABC trans-
tributing to the severity of the phenotype. porter activity cause accumulation of conjugated bilirubin, bile
salts, and other solutes, both in blood, liver, and other organs.
Since the liver has transporters for uptake of bile salts, organic
anions, and cations hepatocytes suffer most from hepatotoxic
ABC TRANSPORTERS AND DRUG‐ drugs, either directly or indirectly via increased cytotoxic bile
INDUCED LIVER DISEASE salts. Bile salts at concentrations of 200 μm and above are toxic
as they cause apoptosis and necrosis. Mitochondrial toxicity
The list of drugs causing cholestasis and cholestatic hepatitis is occurs at lower concentrations and result in loss of hepatocel-
long. Frequently implicated cholestasis‐inducing drugs are lular polarity (see Chapters 4 and 8).
amoxicillin–clavulanic acid, flucloxacillin, rifampicin, oral Nuclear hormone receptors such as FXR, PXR, CAR, and
contraceptives, and diclofenac. Oral contraceptives and PPARα and δ act as transcriptional regulators of ABC
26:  Hepatic Adenosine Triphosphate‐Binding Cassette Transport Proteins and Their Role in Physiology 323

transporters and bile salt synthesis. High affinity ligands for 18. Juliano, R.L. and Ling, V. A surface glycoprotein modulating drug permea-
these receptors are candidate drugs for the treatment of choles- bility in Chinese hamster ovary cell mutants. Biochim Biophys Acta,
1976;455:152–62.
tatic liver disease (see Chapters 30 and 31). Ursodeoxycholic 19. Vlaming, M.L., Mohrmann, K., Wagenaar, E. et al. Carcinogen and antican-
acid and norursodeoxycholic acid act at the post‐transcriptional cer drug transport by Mrp2 in vivo: studies using Mrp2 (Abcc2) knockout
level by facilitating the insertion of transport proteins into the mice. J Pharmacol Exp Ther, 2006;318:319–27.
canalicular membrane. In particular ursodeoxycholic acid has 20. Lam P., Wang R., and Ling, V. Bile acid transport in sister of P‐glycoprotein
(ABCB11) knockout mice. Biochemistry, 2005;44:12598–605.
proved its value in the treatment of cholestatic liver disease.
21. Smit, J.J., Schinkel, A.H., Oude Elferink, R.P. et al. Homozygous disruption
of the murine mdr2 P‐glycoprotein gene leads to a complete absence of
phospholipid from bile and to liver disease. Cell, 1993;75:451–62.
22. Halpern, K.B., Shenhav, R., Matcovitch‐Natan, O. et al. Single‐cell spatial
ACKNOWLEDGEMENT reconstruction reveals global division of labour in the mammalian liver.
Nature, 2017;542:352–6.
23. Meyer, K., Ostrenko, O., Bourantas, G. et al. A predictive 3D multi‐scale model
I thank Drs Jan Hengstler and Nachiket Vartak (Department of
of biliary fluid dynamics in the liver lobule. Cell Syst, 2017;4:277–90.
Toxicology, Leibniz Research Centre for Working Environment 24. Fu, D., Wakabayashi, Y., Ido, Y., Lippincott‐Schwartz, J., and Arias, I.M.
and Human Factors, Dortmund, Germany) for critically reading Regulation of bile canalicular network formation and maintenance by AMP‐
the manuscript. activated protein kinase and LKB1. J Cell Sci, 2010;123:3294–302.
25. Jansen, P.L., Strautnieks, S.S., Jacquemin, E. et al. Hepatocanalicular bile
salt export pump deficiency in patients with progressive familial intrahepatic
cholestasis. Gastroenterology, 1999;117:1370–9.
26. Kruglov, E.A., Gautam, S., Guerra, M.T., and Nathanson, M.H. Type 2 ino-
REFERENCES sitol 1,4,5‐trisphosphate receptor modulates bile salt export pump activity in
rat hepatocytes. Hepatology, 2011;54:1790–9.
1. Sperber, I. Secretion of organic anions in the formation of urine and bile. 27. Mills, C.O., Milkiewicz, P., Muller, M. et al. Different pathways of canalicu-
Pharmacol Rev, 1959;11:109–34. lar secretion of sulfated and non‐sulfated fluorescent bile acids: a study in
2. Javitt, N.B. and Emerman, S. Effect of sodium taurolithocholate on bile flow isolated hepatocyte couplets and TR‐ rats. J Hepatol, 1999;31:678–84.
and bile acid exeretion. J Clin Invest, 1968;47:1002–14. 28. Porat‐Shliom, N., Tietgens, A.J., Van Itallie, C.M. et  al. Liver kinase B1
3. Boyer, J.L. and Klatskin, G. Canalicular bile flow and bile secretory pres- regulates hepatocellular tight junction distribution and function in vivo.
sure. Evidence for a non‐bile salt dependent fraction in the isolated perfused Hepatology, 2016;64:1317–29.
rat liver. Gastroenterology, 1970;59:853–9. 29. Groothuis, G.M., Hardonk, M.J., Keulemans, K.P., Nieuwenhuis, P., and
4. Russell, T.R., Searle, G.L., and Jones, R.S. The choleretic mechanisms of Meijer, D.K. Autoradiographic and kinetic demonstration of acinar heteroge-
sodium taurocholate, secretin, and glucagon. Surgery, 1975;77:498–504. neity of taurocholate transport. Am J Physiol, 1982;243:G455–62.
5. Boyer, J.L. and Bloomer, J.R. Canalicular bile secretion in man. Studies uti- 30. Ling, V., Kartner, N., Sudo, T., Siminovitch, L., and Riordan, J.R. Multidrug‐
lizing the biliary clearance of (14C)mannitol. J Clin Invest, resistance phenotype in Chinese hamster ovary cells. Cancer Treat Rep,
1974;54:773–81. 1983;67:869–74.
6. Javitt, N.B. History of hepatic bile formation: old problems, new approaches. 31. Kartner, N., Shales, M., Riordan, J.R., and Ling, V. Daunorubicin‐resistant
Adv Physiol Educ, 2014;38:279–85. Chinese hamster ovary cells expressing multidrug resistance and a cell‐sur-
7. Marinelli, R.A. and LaRusso N.F. Aquaporin water channels in liver: their face P‐glycoprotein. Cancer Res, 1983;43:4413–9.
significance in bile formation. Hepatology, 1997;26:1081–4. 32. Higgins, C.F. ABC transporters: from microorganisms to man. Annu Rev
8. O’Maille E.R. and Hofmann, A.F. Relatively high biliary secretory maxi- Cell Biol, 1992;8:67–113.
mum for non‐micelle‐forming bile acid: possible significance for mecha- 33. Childs, S. and Ling, V. The MDR superfamily of genes and its biological
nism of secretion. Q J Exp Physiol, 1986;71:475–82. implications. Important Adv Oncol, 1994:21–36.
9. Kamimoto Y., Gatmaitan Z., Hsu J., and Arias, I.M. The function of Gp170, 34. Vasiliou, V., Vasiliou, K., and Nebert, D.W. Human ATP‐binding cassette
the multidrug resistance gene product, in rat liver canalicular membrane (ABC) transporter family. Hum Genomics, 2009;3:281–90.
vesicles. J Biol Chem, 1989;264:11693–8. 35. Chen, Z., Shi, T., Zhang, L. et al. Mammalian drug efflux transporters of the
10. Nishida T., Gatmaitan Z., Che M., and Arias, I.M. Rat liver canalicular mem- ATP binding cassette (ABC) family in multidrug resistance: a review of the
brane vesicles contain an ATP‐dependent bile acid transport system. Proc past decade. Cancer Lett, 2016;370:153–64.
Natl Acad Sci U S A, 1991;88:6590–4. 36. Ueda, K., Cornwell, M.M., Gottesman, M.M., Pastan, I., Roninson, I.B., Ling,
11. Kitamura T., Jansen P., Hardenbrook C., Kamimoto Y., Gatmaitan Z., and V., and Riordan, J.R. The mdr1 gene, responsible for multidrug‐resistance,
Arias, I.M. Defective ATP‐dependent bile canalicular transport of organic codes for P‐glycoprotein. Biochem Biophys Res Commun, 1986;141:956–62.
anions in mutant (TR‐) rats with conjugated hyperbilirubinemia. Proc Natl 37. Gottesman, M.M. and Pastan, I.H. The role of multidrug resistance efflux
Acad Sci U S A, 1990;87:3557–61. pumps in cancer: revisiting a JNCI publication exploring expression of the
12. Schulman, F.Y. Montali, R.J., Bush, M. et al. Dubin‐Johnson‐like Syndrome MDR1 (P‐glycoprotein) gene. J Natl Cancer Inst, 2015;107.
in Golden Lion Tamarins (Leontopithecus rosalia rosalia). Vet Pathol, 38. Crawford, A.R., Smith, A.J., Hatch, V.C., Oude Elferink, R.P., Borst, P., and
1991;30:8. Crawford, J.M. Hepatic secretion of phospholipid vesicles in the mouse criti-
13. Jansen, P.L., Peters, W.H., and Lamers, W.H. Hereditary chronic conjugated cally depends on mdr2 or MDR3 P‐glycoprotein expression. Visualization
hyperbilirubinemia in mutant rats caused by defective hepatic anion trans- by electron microscopy. J Clin Invest, 1997;100:2562–7.
port. Hepatology, 1985;5:573–9. 39. Oude Elferink, R.P., Ottenhoff, R., van Wijland, M., Smit, J.J., Schinkel,
14. Alpert S., Mosher M., Shanske, A., and Arias, I.M. Multiplicity of hepatic A.H., and Groen, A.K. Regulation of biliary lipid secretion by mdr2 P‐glyco-
excretory mechanisms for organic anions. J Gen Physiol, 1969;53:238–7. protein in the mouse. J Clin Invest, 1995;95:31–8.
15. Jansen, P.L., Groothuis, G.M., Peters, W.H., and Meijer, D.F. Selective hepato- 40. Lee, S.D., Thornton, S.J., Sachs‐Barrable, K., Kim, J.H., and Wasan, K.M.
biliary transport defect for organic anions and neutral steroids in mutant rats Evaluation of the contribution of the ATP binding cassette transporter, P‐gly-
with hereditary‐conjugated hyperbilirubinemia. Hepatology, 1987;7:71–6. coprotein, to in vivo cholesterol homeostasis. Mol Pharm, 2013;10:3203–12.
16. Paulusma, C.C., Bosma, P.J., Zaman, G.J. et al. Congenital jaundice in rats 41. Berge, K.E., Tian, H., Graf, G.A. et al. Accumulation of dietary cholesterol
with a mutation in a multidrug resistance‐associated protein gene. Science, in sitosterolemia caused by mutations in adjacent ABC transporters. Science,
1996;271:1126–8. 2000;290:1771–5.
17. Gerloff, T., Stieger, B., Hagenbuch, B. et  al. The sister of P‐glycoprotein 42. Elferink, R.P., Ottenhoff, R., Liefting, W., de Haan, J., and Jansen, P.L.
represents the canalicular bile salt export pump of mammalian liver. J Biol Hepatobiliary transport of glutathione and glutathione conjugate in rats with
Chem, 1998;273:10046–50. hereditary hyperbilirubinemia. J Clin Invest, 1989;84:476–83.
324 THE LIVER:  REFERENCES

43. Rosmorduc, O., Hermelin, B., and Poupon, R. MDR3 gene defect in adults 65. Strautnieks, S.S., Bull, L.N., Knisely, A.S. et al. A gene encoding a liver‐spe-
with symptomatic intrahepatic and gallbladder cholesterol cholelithiasis. cific ABC transporter is mutated in progressive familial intrahepatic choles-
Gastroenterology, 2001;120:1459–67. tasis. Nat Genet, 1998;20:233–8.
44. Rebholz, C., Krawczyk, M., and Lammert, F. Genetics of gallstone disease. 66. van Mil, S.W., van der Woerd, W.L., van der Brugge, G. et al. Benign recur-
J Clin Invest,Eur J Clin Invest, 2018;48:e12935. rent intrahepatic cholestasis type 2 is caused by mutations in ABCB11.
45. Hirschfield, G.M., Chapman, R.W., Karlsen, T.H., Lammert, F., Lazaridis, Gastroenterology, 2004;127:379–84.
K.N., and Mason, A.L. The genetics of complex cholestatic disorders. 67. Knisely, A.S., Strautnieks, S.S., Meier, Y. et al. Hepatocellular carcinoma in
Gastroenterology, 2013;144:1357–74. ten children five years of age with bile salt export deficience. Hepatology,
46. Paulusma, C.C., de Waart, D.R., Kunne, C., Mok, K.S., and Elferink, R.P. 2006;44:478–86.
Activity of the bile salt export pump (ABCB11) is critically dependent on 68. Siebold, L., Dick, A.A., Thompson, R. et al. Recurrent low gamma‐glutamyl
canalicular membrane cholesterol content. J Biol Chem, 2009; transpeptidase cholestasis following liver transplantation for bile salt export
284:9947–54. pump (BSEP) disease (posttransplant recurrent BSEP disease). Liver
47. Guyot, C., Hofstetter, L., and Stieger, B. Differential effects of membrane Transpl, 2010;16:856–63.
cholesterol content on the transport activity of multidrug resistance‐associ- 69. Kubitz, R., Droge, C., Kluge, S. et  al. High affinity anti‐BSEP antibodies
ated protein 2 (ABCC2) and of the bile salt export pump (ABCB11). Mol after liver transplantation for PFIC‐2 – successful treatment with immunoad-
Pharmacol, 2014;85:909–20. sorption and B‐cell depletion. Pediatr Transplant, 2016;20:987–93.
48. Szilagyi, J.T., Vetrano, A.M., Laskin, J.D., and Aleksunes, L.M. Localization 70. Hayashi, H. and Sugiyama, Y. 4‐phenylbutyrate enhances the cell surface
of the placental BCRP/ABCG2 transporter to lipid rafts: role for cholesterol expression and the transport capacity of wild‐type and mutated bile salt
in mediating efflux activity. Placenta, 2017;55:29–36. export pumps. Hepatology, 2007;45:1506–16.
49. Telbisz, A., Hegedus, C., Varadi, A., Sarkadi, B., and Ozvegy‐Laczka, C. 71. Hirano, M., Maeda, K., Hayashi, H., Kusuhara, H., and Sugiyama, Y. Bile
Regulation of the function of the human ABCG2 multidrug transporter by salt export pump (BSEP/ABCB11) can transport a nonbile acid substrate,
cholesterol and bile acids: effects of mutations in potential substrate and ster- pravastatin. J Pharmacol Exp Ther, 2005;314:876–82.
oid binding sites. Drug Metab Dispos, 2014;42:575–85. 72. Fattinger, K., Funk, C., Pantze, M. et  al. The endothelin antagonist bosentan
50. Paulusma, C.C., Groen, A., Kunne, C. et  al. Atp8b1 deficiency in mice inhibits the canalicular bile salt export pump: a potential mechanism for hepatic
reduces resistance of the canalicular membrane to hydrophobic bile salts and adverse reactions. Pharmacol Ther Clin Pharmacol Ther, 2001;69:223–31.
impairs bile salt transport. Hepatology, 2006;44:195–204. 73. Funk, C., Ponelle, C., Scheuermann, G., and Pantze, M. Cholestatic potential
51. Bull, L.N., Carlton, V.E., Stricker, N.L. et  al. Genetic and morphological of troglitazone as a possible factor contributing to troglitazone‐induced
findings in progressive familial intrahepatic cholestasis (Byler disease hepatotoxicity: in vivo and in vitro interaction at the canalicular bile salt
[PFIC‐1] and Byler syndrome): evidence for heterogeneity. Hepatology, export pump (Bsep) in the rat. Mol Pharmacol, 2001;59:627–35.
1997;26:155–64. 74. Byrne, J.A., Strautnieks, S.S., Mieli‐Vergani, G., Higgins, C.F., Linton, K.J.,
52. Sinke, R.J., Carlton, V.E., Juijn, J.A. et  al. Benign recurrent intrahepatic and Thompson, R.J. The human bile salt export pump: characterization of
cholestasis (BRIC): evidence of genetic heterogeneity and delimitation of substrate specificity and identification of inhibitors. Gastroenterology,
the BRIC locus to a 7‐cM interval between D18S69 and D18S64. Hum 2002;123:1649–58.
Genet, 1997;100:382–7. 75. Nicolaou, M., Andress, E.J., Zolnerciks, J.K., Dixon, P.H., Williamson, C.,
53. Bull, L.N., Pawlikowska, L., Strautnieks, S. et al. Outcomes of surgical man- and Linton, K.J. Canalicular ABC transporters and liver disease. J Pathol,
agement of familial intrahepatic cholestasis 1 and bile salt export protein 2012;226:300–15.
deficiencies. Hepatol Commun, 2018;2:515–28. 76. Tanigawara, Y., Okamura, N., Hirai, M. et al. Transport of digoxin by human
54. Pauli‐Magnus, C., Kerb, R., Fattinger, K. et al. BSEP and MDR3 haplotype P‐glycoprotein expressed in a porcine kidney epithelial cell line (LLC‐PK1).
structure in healthy Caucasians, primary biliary cirrhosis and primary scle- J Pharmacol Exp Ther, 1992;263:840–5.
rosing cholangitis. Hepatology, 2004;39:779–91. 77. Roulet, A., Puel, O., Gesta, S. et al. MDR1‐deficient genotype in Collie dogs
55. Ji, S.G., Juran, B.D., Mucha, S. et al. Genome‐wide association study of pri- hypersensitive to the P‐glycoprotein substrate ivermectin. Eur J Pharmacol,
mary sclerosing cholangitis identifies new risk loci and quantifies the genetic 2003;460:85–91.
relationship with inflammatory bowel disease. Nat Genet, 2017;49:269–73. 78. Schinkel, A.H., Smit, J.J., van Tellingen, O. et  al. Disruption of the
56. de Vree, J.M., Jacquemin, E., Sturm, E. et al. Mutations in the MDR3 gene mouse mdr1a P‐glycoprotein gene leads to a deficiency in the blood‐
cause progressive familial intrahepatic cholestasis. Proc Natl Acad Sci U S A, brain barrier and to increased sensitivity to drugs. Cell, 1994;
1998;95:282–7. 77:491–502.
57. Smith, A.J., de Vree, J.M., Ottenhoff, R., Oude Elferink, R.P., Schinkel, 79. Ellis, J.L., Bove, K.E., Schuetz, E.G. et al. Zebrafish abcb11b mutant reveals
A.H., and Borst, P. Hepatocyte‐specific expression of the human MDR3 P‐ strategies to restore bile excretion impaired by bile salt export pump defi-
glycoprotein gene restores the biliary phosphatidylcholine excretion absent ciency. Hepatology, 2018;67:1531–45.
in Mdr2 (‐/‐) mice. Hepatology, 1998;28:530–6. 80. Wicek, K. and Stieger, B. ATP‐binding cassette transporters in liver.
58. Deleuze, J.F., Jacquemin, E., Dubuisson, C. et al. Defect of multidrug‐resist- Biofactors, 2014;40:188–98.
ance 3 gene expression in a subtype of progressive familial intrahepatic chol- 81. Bakos, E., Evers, R., Szakacs, G. et al. Functional multidrug resistance pro-
estasis. Hepatology, 1996;23:904–8. tein (MRP1) lacking the N‐terminal transmembrane domain. J Biol Chem,
59. Jacquemin, E., Cresteil, D., Manouvrier, S., Boute, O., and Hadchouel, M. 1998;273:32167–75.
Heterozygous non‐sense mutation of the MDR3 gene in familial intrahepatic 82. Megaraj, V., Iida, T., Jungsuwadee, P., Hofmann, A.F., and Vore, M.
cholestasis of pregnancy. Lancet, 1999;353:210–1. Hepatobiliary disposition of 3alpha,6alpha,7alpha,12alpha‐tetrahydroxy‐
60. Dixon, P.H., Weerasekera, N., Linton, K.J. et al. Heterozygous MDR3 mis- cholanoyl taurine: a substrate for multiple canalicular transporters. Drug
sense mutation associated with intrahepatic cholestasis of pregnancy: evi- Metab Dispos, 2010;38:1723–30.
dence for a defect in protein trafficking. Hum Mol Genet, 2000;9:1209–17. 83. Huber, M., Guhlmann, A., Jansen, P.L., and Keppler, D. Hereditary defect of
61. Kok, T., Wolters, H., Bloks, V.W. et al. Induction of hepatic ABC transporter hepatobiliary cysteinyl leukotriene elimination in mutant rats with defective
expression is part of the PPARalpha‐mediated fasting response in the mouse. hepatic anion excretion. Hepatology, 1987;7:224–8.
Gastroenterology, 2003;124:160–71. 84. Cui, Y., Konig, J., Buchholz, J.K., Spring, H., Leier, I., and Keppler, D. Drug
62. Fickert, P., Wagner, M., Marschall, H.U. et al. 24‐norUrsodeoxycholic acid resistance and ATP‐dependent conjugate transport mediated by the apical
is superior to ursodeoxycholic acid in the treatment of sclerosing cholangitis multidrug resistance protein, MRP2, permanently expressed in human and
in Mdr2 (Abcb4) knockout mice. Gastroenterology, 2006;130:465–81. canine cells. Mol Pharmacol, 1999;55:929–37.
63. Wang, R., Salem, M., Yousef, I.M. et al. Targeted inactivation of sister of P‐ 85. Kamisako, T., Leier, I., Cui, Y. et  al. Transport of monoglucuronosyl and
glycoprotein gene (spgp) in mice results in nonprogressive but persistent bisglucuronosyl bilirubin by recombinant human and rat multidrug resist-
intrahepatic cholestasis. Proc Natl Acad Sci U S A, 2001;98:2011–6. ance protein 2. Hepatology, 1999;30:485–90.
64. Hrycay, E., Forrest, D., Liu, L. et  al. Hepatic bile acid metabolism and 86. Madon, J., Hagenbuch, B., Landmann, L., Meier, P.J., and Stieger, B.
expression of cytochrome P450 and related enzymes are altered in Bsep (‐/‐) Transport function and hepatocellular localization of mrp6 in rat liver. Mol
mice. Mol Cell Biochem, 2014;389:119–32. Pharmacol, 2000;57:634–41.
26:  Hepatic Adenosine Triphosphate‐Binding Cassette Transport Proteins and Their Role in Physiology 325

87. Dijkstra, M., Havinga, R., Vonk, R.J., and Kuipers, F. Bile secretion of cad- 110. Keppler, D. Cholestasis and the role of basolateral efflux pumps. Z
mium, silver, zinc and copper in the rat. Involvement of various transport Gastroenterol, 2011;49:1553–7.
systems. Life Sci, 1996;59:1237–46. 111. Zelcer, N., Reid, G., Wielinga, P. et al. Steroid and bile acid conjugates are
88. Sandusky, G.E., Mintze, K.S., Pratt, S.E., and Dantzig, A.H. Expression of substrates of human multidrug‐resistance protein (MRP) 4 (ATP‐binding
multidrug resistance‐associated protein 2 (MRP2) in normal human tissues cassette C4). Biochem, J., 2003;371:361–7.
and carcinomas using tissue microarrays. Histopathology, 2002;41:65–74. 112. Assem, M., Schuetz, E.G., Leggas, M. et al. Interactions between hepatic
89. Rost, D., Konig, J., Weiss, G., Klar, E., Stremmel, W., and Keppler, D. Mrp4 and Sult2a as revealed by the constitutive androstane receptor and
Expression and localization of the multidrug resistance proteins MRP2 and Mrp4 knockout mice. J Biol Chem, 2004;279:22250–7.
MRP3 in human gallbladder epithelia. Gastroenterology, 2001;121:1203–8. 113. Rius, M., Hummel‐Eisenbeiss, J., Hofmann, A.F., and Keppler, D. Substrate
90. Kartenbeck, J., Leuschner, U., Mayer, R., and Keppler, D. Absence of the specificity of human ABCC4 (MRP4)‐mediated cotransport of bile acids
canalicular isoform of the MRP gene‐encoded conjugate export pump from and reduced glutathione. Am J Physiol Gastrointest Liver Physiol,
the hepatocytes in Dubin–Johnson syndrome. Hepatology, 1996;23:1061–6. 2006;290:G640–9.
91. Kitamura, T., Alroy, J., Gatmaitan, Z. et al. Defective biliary excretion of 114. Denk, G.U., Soroka, C.J., Takeyama, Y., Chen, W.S., Schuetz, J.D., and
epinephrine metabolites in mutant (TR‐) rats: relation to the pathogenesis Boyer, J.L. Multidrug resistance‐associated protein 4 is up‐regulated in
of black liver in the Dubin–Johnson syndrome and Corriedale sheep with liver but down‐regulated in kidney in obstructive cholestasis in the rat.
an analogous excretory defect. Hepatology, 1992;15:1154–9. J Hepatol, 2004;40:585–91.
92. Lagas, J.S., Vlaming, M.L., van Tellingen, O. et al. Multidrug resistance 115. Favre, G., Laurain, A., Aranyi, T. et  al. The ABCC6 transporter: a new
protein 2 is an important determinant of paclitaxel pharmacokinetics. Clin player in biomineralization. Int J Mol Sci, 2017;18. e1941.
Cancer Res Cancer Res, 2006;12:6125–32. 116. Jansen, R.S., Duijst, S., Mahakena, S. et al. ABCC6‐mediated ATP secre-
93. Borel, F., Han, R., Visser, A. et al. Adenosine triphosphate‐binding cassette tion by the liver is the main source of the mineralization inhibitor inorganic
transporter genes up‐regulation in untreated hepatocellular carcinoma is pyrophosphate in the systemic circulation‐brief report. Arterioscler Thromb
mediated by cellular microRNAs. Hepatology, 2012;55:821–32. Vasc Biol, 2014;34:1985–9.
94. Nies, A.T., Konig, J., Pfannschmidt, M., Klar, E., Hofmann, W.J., and 117. Jansen, R.S., Kucukosmanoglu, A., de Haas, M. et  al. ABCC6 prevents
Keppler, D. Expression of the multidrug resistance proteins MRP2 and ectopic mineralization seen in pseudoxanthoma elasticum by inducing cel-
MRP3 in human hepatocellular carcinoma. Int J Cancer, 2001;94:492–9. lular nucleotide release. Proc Natl Acad Sci U S A, 2013;110:20206–11.
95. Recknagel, P., Gonnert, F.A., Westermann, M. et al. Liver dysfunction and 118. Riordan, J.R., Rommens, J.M., Kerem, B. et al. Identification of the cystic
phosphatidylinositol‐3‐kinase signalling in early sepsis: experimental stud- fibrosis gene: cloning and characterization of complementary DNA.
ies in rodent models of peritonitis. PLoS Med, 2012;9:e1001338. Science, 1989;245:1066–73.
96. Trauner, M., Arrese, M., Soroka, C.J. et al. The rat canalicular conjugate 119. Jonker, J.W., Buitelaar, M., Wagenaar, E. et al. The breast cancer resistance
export pump (Mrp2) is down‐regulated in intrahepatic and obstructive chol- protein protects against a major chlorophyll‐derived dietary phototoxin and
estasis. Gastroenterology, 1997;113:255–64. protoporphyria. Proc Natl Acad Sci U S A, 2002;99:15649–54.
97. Hinoshita, E., Taguchi, K., Inokuchi, A. et al. Decreased expression of an 120. Vander Borght, S., Libbrecht, L., Katoonizadeh, A. et  al. Breast cancer
ATP‐binding cassette transporter, MRP2, in human livers with hepatitis C resistance protein (BCRP/ABCG2) is expressed by progenitor cells/reac-
virus infection. J Hepatol, 2001;35:765–73. tive ductules and hepatocytes and its expression pattern is influenced by
98. Zinchuk, V., Zinchuk, O., Akimaru, K., Moriya, F., and Okada, T. Ethanol con- disease etiology and species type: possible functional consequences. J
sumption alters expression and colocalization of bile salt export pump and mul- Histochem Cytochem, 2006;54:1051–9.
tidrug resistance protein 2 in the rat. Histochem Cell Biol, 2007;127:503–12. 121. Jakulj, L., Vissers, M.N., Tanck, M.W. et al. ABCG5/G8 polymorphisms
99. Ros, J.E., Libbrecht, L., Geuken, M., Jansen, P.L., and Roskams, T.A. and markers of cholesterol metabolism: systematic review and meta‐analy-
High expression of MDR1, MRP1, and MRP3 in the hepatic progenitor sis. J Lipid Res, 2010;51:3016–23.
cell compartment and hepatocytes in severe human liver disease. J 122. Yu, L., Li‐Hawkins, J., Hammer, R.E. et al. Overexpression of ABCG5 and
Pathol, 2003;200:553–60. ABCG8 promotes biliary cholesterol secretion and reduces fractional
100. Ros, J.E., Roskams, T.A., Geuken, M. et  al. ATP binding cassette trans- absorption of dietary cholesterol. J Clin Invest, 2002;110:671–80.
porter gene expression in rat liver progenitor cells. Gut, 2003;52:1060–7. 123. von Kampen, O., Buch, S., Nothnagel, M. et  al. Genetic and functional
101. Silverton, L., Dean, M., and Moitra, K. Variation and evolution of the ABC identification of the likely causative variant for cholesterol gallstone dis-
transporter genes ABCB1, ABCC1, ABCG2, ABCG5 and ABCG8: impli- ease at the ABCG5/8 lithogenic locus. Hepatology, 2013;57:2407–17.
cation for pharmacogenetics and disease. Drug Metabol Drug Interact, 124. Buch, S., Schafmayer, C., Volzke, H. et al. A genome‐wide association scan
2011;26:169–79. identifies the hepatic cholesterol transporter ABCG8 as a susceptibility fac-
102. Harris, M.J., Kuwano, M., Webb, M., and Board, P.G. Identification of the tor for human gallstone disease. Nat Genet, 2007;39:995–9.
apical membrane‐targeting signal of the multidrug resistance‐associated 125. Makishima, M., Okamoto, A.Y., Repa, J.J. et al. Identification of a nuclear
protein 2 (MRP2/MOAT). J Biol Chem, 2001;276:20876–81. receptor for bile acids. Science, 1999;284:1362–5.
103. Vander Borght, S., Komuta, M., Libbrecht, L. et al. Expression of multid- 126. Kliewer, S.A., Moore, J.T., Wade, L. et al. An orphan nuclear receptor acti-
rug resistance‐associated protein 1 in hepatocellular carcinoma is associ- vated by pregnanes defines a novel steroid signaling pathway. Cell,
ated with a more aggressive tumour phenotype and may reflect a progenitor 1998;92:73–82.
cell origin. Liver Int, 2008;28:1370–80. 127. Baes, M., Gulick, T., Choi, H.S., Martinoli, M.G., Simha, D., and Moore,
104. Soroka, C.J., Lee, J.M., Azzaroli, F., and Boyer, J.L. Cellular localization D.D. A new orphan member of the nuclear hormone receptor superfamily
and up‐regulation of multidrug resistance‐associated protein 3 in hepato- that interacts with a subset of retinoic acid response elements. Mol Cell
cytes and cholangiocytes during obstructive cholestasis in rat liver. Biol, 1994;14:1544–52.
Hepatology, 2001;33:783–91. 128. Boyer, J.L., Trauner, M., Mennone, A. et al. Upregulation of a basolateral
105. Konig, J., Rost, D., Cui, Y., and Keppler, D. Characterization of the human FXR‐dependent bile acid efflux transporter OSTalpha‐OSTbeta in choles-
multidrug resistance protein isoform MRP3 localized to the basolateral tasis in humans and rodents. Am J Physiol Gastrointest Liver Physiol,
hepatocyte membrane. Hepatology, 1999;29:1156–63. 2006;290:G1124–30.
106. Scheffer, G.L., Kool, M., de Haas, M. et al. Tissue distribution and induc- 129. Keppler, D. The roles of MRP2, MRP3, OATP1B1, and OATP1B3 in con-
tion of human multidrug resistant protein 3. Lab Invest, 2002;82:193–201. jugated hyperbilirubinemia. Drug Metab Dispos, 2014;42:561–5.
107. Donner, M.G. and Keppler, D. Up‐regulation of basolateral multidrug resist- 130. Mennone, A., Soroka, C.J., Cai, S.Y. et al. Mrp4‐/‐ mice have an impaired
ance protein 3 (Mrp3) in cholestatic rat liver. Hepatology, 2001;34:351–9. cytoprotective response in obstructive cholestasis. Hepatology,
108. Hirohashi, T., Suzuki, H., and Sugiyama, Y. Characterization of the trans- 2006;43:1013–21.
port properties of cloned rat multidrug resistance‐associated protein 3 131. Parks, D.J., Blanchard, S.G., Bledsoe, R.K. et al. Bile acids: natural ligands
(MRP3). J Biol Chem, 1999;274:15181–5. for an orphan nuclear receptor. Science, 1999;284:1365–8.
109. Conseil, G., Deeley, R.G., and Cole, S.P. Polymorphisms of MRP1 132. Wang, H., Chen, J., Hollister, K., Sowers, L.C., and Forman, B.M.
(ABCC1) and related ATP‐dependent drug transporters. Pharmacogenet Endogenous bile acids are ligands for the nuclear receptor FXR/BA.R. Mol
Genomics, 2005;15:523–33. Cell, 1999;3:543–53.
326 THE LIVER:  REFERENCES

133. Kir, S., Zhang, Y., Gerard, R.D., Kliewer, S.A., and Mangelsdorf, D.J. 154. Elferink, R.P. and Paulusma, C.C. MRP2 in cholestasis: putting down the
Nuclear receptors HNF4alpha and LRH‐1 cooperate in regulating Cyp7a1 anchor. J Hepatol, 2015;63:1309–10.
in vivo. J Biol Chem, 2012;287:41334–41. 155. Ortiz, D.F., Moseley, J., Calderon, G., Swift, A.L., Li, S., and Arias, I.M.
134. Cai, S.Y. and Boyer, J.L. FXR: a target for cholestatic syndromes? Expert Identification of HAX‐1 as a protein that binds bile salt export protein and
Opin Ther Targets, 2006;10:409–21. regulates its abundance in the apical membrane of Madin‐Darby canine
135. Schaap, F.G., van der Gaag, N.A., Gouma, D.J., and Jansen, P.L. High kidney cells. J Biol Chem, 2004;279:32761–70.
expression of the bile salt‐homeostatic hormone fibroblast growth factor 19 156. Paulusma, C.C., Kothe, M.J., Bakker, C.T. et al. Zonal down‐regulation and
in the liver of patients with extrahepatic cholestasis. Hepatology, redistribution of the multidrug resistance protein 2 during bile duct ligation
2009;49:1228–35. in rat liver. Hepatology, 2000;31:684–93.
136. Gomez‐Ospina, N., Potter, C.J., Xiao, R. et al. Mutations in the nuclear bile 157. Miszczuk, G.S., Barosso, I.R., Larocca, M.C. et al. Mechanisms of canali-
acid receptor FXR cause progressive familial intrahepatic cholestasis. Nat cular transporter endocytosis in the cholestatic rat liver. Biochim Biophys
Commun, 2016;7:10713. Acta, 2018;1864:1072–85.
137. Staudinger, J.L., Goodwin, B., Jones, S.A. et al. The nuclear receptor PXR 158. Kikuchi, S., Hata, M., Fukumoto, K. et al. Radixin deficiency causes con-
is a lithocholic acid sensor that protects against liver toxicity. Proc Natl jugated hyperbilirubinemia with loss of Mrp2 from bile canalicular mem-
Acad Sci U S A, 2001;98:3369–74. branes. Nat Genet, 2002;31:320–5.
138. Xie, W., Radominska‐Pandya, A., Shi, Y. et al. An essential role for nuclear 159. Kojima, H., Sakurai, S., Uemura, M. et al. Disturbed colocalization of mul-
receptors SXR/PXR in detoxification of cholestatic bile acids. Proc Natl tidrug resistance protein 2 and radixin in human cholestatic liver diseases.
Acad Sci U S A, 2001;98:3375–80. J Gastroenterol Hepatol, 2008;23:e120–8.
139. Huang, W., Zhang, J., Chua, S.S. et al. Induction of bilirubin clearance by 160. Kojima, H., Sakurai, S., Yoshiji, H., Uemura, M., Yoshikawa, M., and
the constitutive androstane receptor (CAR). Proc Natl Acad Sci U S A, Fukui, H. The role of radixin in altered localization of canalicular conjugate
2003;100:4156–61. export pump Mrp2 in cholestatic rat liver. Hepatol Res, 2008;38:202–10.
140. Makishima, M., Lu, T.T., Xie, W. et al. Vitamin D receptor as an intestinal 161. Chai, J., Cai, S.Y., Liu, X. et  al. Canalicular membrane MRP2/ABCC2
bile acid sensor. Science, 2002;296:1313–6. internalization is determined by Ezrin Thr567 phosphorylation in human
141. van Dijk, R., Kremer, A.E., Smit, W. et al. Characterization and treatment obstructive cholestasis. J Hepatol, 2015;63:1440–8.
of persistent hepatocellular secretory failure. Liver Int, 2015;35:1478–88. 162. Stieger, B., Fattinger, K., Madon, J., Kullak‐Ublick, G.A., and Meier, P.J.
142. Janssen, H.L., Reesink, H.W., Lawitz, E.J. et al. Treatment of HCV infec- Drug‐ and estrogen‐induced cholestasis through inhibition of the hepato-
tion by targeting microRNA. N Engl J Med, 2013;368:1685–94. cellular bile salt export pump (Bsep) of rat liver. Gastroenterology,
143. Haenisch, S., Laechelt, S., Bruckmueller, H. et al. Down‐regulation of ATP‐ 2000;118:422–30.
binding cassette C2 protein expression in HepG2 cells after rifampicin treat- 163. Ryan, J., Morgan, R.E., Chen, Y., Volak, L.P., Dunn, R.T., 2nd, and Dunn,
ment is mediated by microRNA‐379. Mol Pharmacol 2011; 80:314–20. K.W. Intravital multiphoton microscopy with fluorescent bile salts in rats as
144. Haenisch, S., Werk, A.N., and Cascorbi, I. MicroRNAs and their relevance an in vivo biomarker for hepatobiliary transport inhibition. Drug Metab
to ABC transporters. Br J Clin Pharmacol, 2014;77:587–96. Dispos, 2018;46:704–18.
145. Yuan, K., Zhang, X., Lv, L., Zhang, J., Liang, W., and Wang, P. Fine‐tuning 164. Pauli‐Magnus, C., Meier, P.J., and Stieger, B. Genetic determinants of drug‐
the expression of microRNA‐155 controls acetaminophen‐induced liver induced cholestasis and intrahepatic cholestasis of pregnancy. Semin Liver
inflammation. Int Immunopharmacol, 2016;40:339–46. Dis, 2010;30:147–59.
146. Church, R.J., Kullak‐Ublick, G.A., Aubrecht, J. et al. Candidate biomark- 165. Dixon, P.H., van Mil, S.W., Chambers, J. et  al. Contribution of variant
ers for the diagnosis and prognosis of drug‐induced liver injury: an interna- alleles of ABCB11 to susceptibility to intrahepatic cholestasis of preg-
tional collaborative effort. Hepatology, 2019;69:760–73. nancy. Gut, 2009;58:537–44.
147. Howell, L.S., Ireland, L., Park, B.K., and Goldring, C.E. MiR‐122 and 166. Pauli‐Magnus, C., Lang, T., Meier, Y. et al. Sequence analysis of bile salt
other microRNAs as potential circulating biomarkers of drug‐induced liver export pump (ABCB11) and multidrug resistance p‐glycoprotein 3
injury. Expert Rev Mol Diagn, 2018;18:47–54. (ABCB4, MDR3) in patients with intrahepatic cholestasis of pregnancy.
148. Rao, Z.Z., Zhang, X.W., Ding, Y.L., and Yang, M.Y. miR‐148a‐mediated Pharmacogenet, 2004;14:91–102.
estrogen‐induced cholestasis in intrahepatic cholestasis of pregnancy: role 167. Lang, C., Meier, Y., Stieger, B. et al. Mutations and polymorphisms in the
of PXR/MRP3. PLoS One, 2017;12:e0178702. bile salt export pump and the multidrug resistance protein 3 associated with
149. Kipp, H., Pichetshote, N., and Arias, I.M. Transporters on demand: intrahe- drug‐induced liver injury. Pharmacogenet Genomics, 2007;17:47–60.
patic pools of canalicular ATP binding cassette transporters in rat liver. J 168. Sundaram, V. and Bjornsson, E.S. Drug‐induced cholestasis. Hepatol
Biol Chem, 2001;276:7218–24. Commun, 2017;1:726–35.
150. Beuers, U., Boyer, J.L., and Paumgartner, G. Ursodeoxycholic acid in chol- 169. Zelcer, N., Saeki, T., Bot, I., Kuil, A., and Borst, P. Transport of bile acids
estasis: potential mechanisms of action and therapeutic applications. in multidrug‐resistance‐protein 3‐overexpressing cells co‐transfected with
Hepatology, 1998;28:1449–53. the ileal Na+‐dependent bile‐acid transporter. Biochem J, 2003;
151. Crocenzi, F.A., Mottino, A.D., Cao, J. et al. Estradiol‐17beta‐D‐glucuron- 369:23–30.
ide induces endocytic internalization of Bsep in rats. Am J Physiol 170. Ghonem, N.S., Ananthanarayanan, M., Soroka, C.J., and Boyer, J.L.
Gastrointest Liver Physiol, 2003;285:G449–59. Peroxisome proliferator‐activated receptor alpha activates human multid-
152. Perez, L.M., Milkiewicz, P., Elias, E., Coleman, R., Sanchez Pozzi, E.J., rug resistance transporter 3/ATP‐binding cassette protein subfamily B4
and Roma, M.G. Oxidative stress induces internalization of the bile salt transcription and increases rat biliary phosphatidylcholine secretion.
export pump, Bsep, and bile salt secretory failure in isolated rat hepatocyte Hepatology, 2014;59:1030–42.
couplets: a role for protein kinase C and prevention by protein kinase A. 171. Allen, R.M., Marquart, T.J., Albert, C.J. et al. miR‐33 controls the expres-
Toxicol Sci, 2006;91:150–8. sion of biliary transporters, and mediates statin‐ and diet‐induced hepato-
153. Haussinger, D., Kurz, A.K., Wettstein, M., Graf, D., Vom Dahl, S., and toxicity. EMBO Mol Med, 2012;4:882–95.
Schliess, F. Involvement of integrins and Src in tauroursodeoxycholate‐ 172. To, K.K., Robey, R., Zhan, Z., Bangiolo, L., and Bates, S.E. Upregulation
induced and swelling‐induced choleresis. Gastroenterology, 2003; of ABCG2 by romidepsin via the aryl hydrocarbon receptor pathway. Mol
124:1476–87. Cancer Res, 2011;9:516–27.
Basolateral Plasma
27 Membrane Organic Anion
Transporters
M. Sawkat Anwer1 and Allan W. Wolkoff2
1
Tufts Clinical and Translational Science Institute, Tufts University Cummings School of Veterinary
Medicine, Department of Biomedical Sciences, North, Grafton, MA, USA
2
Division of Hepatology, Marion Bessin Liver Research Center, Albert Einstein College of Medicine and
Montefiore Medical Center, Bronx, NY, USA

INTRODUCTION SODIUM‐TAUROCHOLATE
COTRANSPORTING POLYPEPTIDE (NTCP)
The liver plays a central role in vectorial transport of solutes
from the portal vein to the canaliculus and this involves trans­ Sodium‐dependent bile salt transport in rat hepatocytes was first
port of solutes via an array of transporters located at the baso­ reported in 1978 [1] and NTCP was cloned from rat liver using
lateral and the canalicular membrane of hepatocytes. Solutes a functional expression cloning approach [2, 3]. It is the found­
transported by hepatocytes include organic anions and cations, ing member of the seven member SLC10A gene family [4, 5]
inorganic ions, and neutral compounds. Many organic solutes and is named Slc10a1 [6]. NTCP is a typical secondary active
(such as bilirubin, bile acids) are albumin bound enabling transporter and strictly sodium‐dependent requiring the binding
them to circulate in blood despite limited aqueous solubility. of two sodium ions together with a bile salt molecule for the
The liver is ideally designed for extraction of protein‐bound transport step. The transport activity is electrogenic for monoan­
molecules using specific transporters. Apart from removing ionic bile salts, while it is electroneutral for the dianionic bile
solutes from the circulation for subsequent metabolic altera­ salt, cholylglycylamidofluorescein [7]. Thus, the driving force
tions and reflux back to the circulation, vectorial transport of of the uptake can be the energy available from the chemical and/
solutes across the bile canaliculus provides the osmotic driv­ or the electrochemical gradient of sodium.
ing force for bile formation. Uptake of organic anions from Uptake of bile salts into hepatocytes occurs predominantly by
the  circulation across the basolateral plasma membrane and NTCP (SLC10A1) and to a lesser extent by OATPs (SLCOs)
subsequent biliary excretion, contributes significantly to
­ [8, 9]. A recent study based on clinical findings in two infants
bile ­formation. Organic anions are primarily transported with NTCP deficiency (based on Sanger sequencing of the
via  sodium‐dependent and sodium‐independent transporters SLC10A1 gene) suggests the primary role of NTCP in hepatic
and include sodium‐taurocholate cotransporting polypeptide bile acid clearance [10]. While human NTCP (hNTCP) and rat
(NTCP) and several members of the organic anion transport­ NTCP (rNTCP) are capable of transporting conjugated bile salts
ing protein family (OATPs). This chapter will discuss func­ as well as unconjugated bile acids [9], OATPs show a preference
tions and regulations of these transporters. for unconjugated bile acids [11–13].

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
328 THE LIVER:  SODIUM‐TAUROCHOLATE COTRANSPORTING POLYPEPTIDE (NTCP)

Transport functions of NTCP A novel function for NTCP as a receptor and transporter for
hepatitis B virus (HBV) and hepatitis D virus (HDV) has been
The list of substrates transported by NTCP has been extensively described [28] and human NTCP plays an important role in viral
reviewed [4, 5, 9]. Solutes other than bile salts including steroid entry and consequent infection [29]. Both HBV and HDV con­
hormones, thyroid hormones, drugs, and drugs conjugated tain three envelope proteins known as small (S), middle (M),
with  bile acids can be transported by hNTCP/rNTCP and large (L). These proteins share the same C‐terminal domain
(SLC10A1/Slc10a1) [5, 14]. In addition to endogenous sub­ corresponding to the S protein but differ at the N‐terminal
strates, drugs including statins and an antifungal agent domains by the presence of the pre‐S2 domain in M protein and
micafungin [15], are transported by NTCP [5]. Interestingly, pre‐S1 and pre‐S2 domains in L protein. The entry of both HBV
rosuvastatin is transported by hNTCP, but not by rNTCP [16]. and HDV is determined by the pre‐S1 domain of L protein after
Thus, despite a large substrate overlap between rat, mouse, and interacting with cellular receptor(s) on hepatocytes [30–32].
human NTCP [9] transport data should be extrapolated with Myrcludex B, a drug based on myristoylated pre‐S1 domain,
caution across different species. Clinical implications of trans­ has entered clinical trials [33] and an initial report shows prom­
port functions of NTCP is discussed below. ise [34]. However, viral entry is only one step in efficient viral
NTCP can serve as a transporter to deliver drugs into hepat­ infection and replication in human hepatocytes and other fac­
ocytes. Hepatocellular carcinomas expressing NTCP mediate tors, such as host components, are likely to be involved [35–38].
high‐affinity Na+‐dependent uptake of chlorambucil‐taurocho­ This novel function of NTCP raises interesting questions
late [17]. Thus, cytotoxic drugs conjugated with TC could pro­ about the mechanism of viral transport. One possibility is that
vide a potential therapeutic strategy to deliver these drugs to these viruses are transported by NTCP the same way as it trans­
hepatocellular carcinomas. The role of transport systems in ports bile salts. In that case, it would be important to know
liver function tests has been known for a long time [18] and is whether the transport of HBV and HDV into hepatocytes is Na+
increasingly appreciated today [19, 20]. Serum bile salt con­ dependent and is inhibited by known substrates/inhibitors of
centration is a good indicator of NTCP function and serves as NTCP. A likely possibility is that NTCP is internalized by endo­
an indicator of liver function. Several exogenous compounds cytosis following binding to the virus allowing the virus to enter
used to assess dynamic liver functions in humans are taken up the cell [35, 36]. Internalization of HBV/HDV may be initiated
by hNTCP among other transporters [21]. Thus, the functional by protein–protein interactions between the envelope protein of
status of NTCP can be an important determinant of liver func­ the virus and NTCP leading to endocytosis. In that case, what
tion testing and should be taken into consideration when would be the signal for internalization? Calcium dependent con­
patients are treated with drugs known to interfere with NTCP ventional protein kinase Cs (PKCs) and most likely PKCα
transport activity. induces rNTCP retrieval [39, 40]. Whether PKCα or other sign­
Drugs may inhibit solute transport by NTCP in two major aling molecules may be involved remains to be investigated.
ways: (i) drugs may be transported by the NTCP resulting in Since NTCP traffics with the epidermal growth factor (EGF)
competitive inhibition and (ii) drugs may inhibit solute uptake receptor via endocytic vesicles [41], HBV and HDV may use
by NTCP without being themselves substrates for transport. this pathway to enter hepatocytes. Irrespective of the transport
Indeed, there are drugs that inhibit hNTCP/rNTCP‐mediated mechanism involved, the finding that NTCP is involved in HBV
bile salt uptake, but are not transported by them [5, 9, 22]. Drugs and HDV infection is expected to inspire further studies leading
that inhibit solute transport by transporters with or without to a better understanding of the role of NTCP in HBV/HDV
being transported can have significant effects on the disposition infection and development of agents to inhibit viral entry and
of endogenous substrates and drugs. These effects are important infection.
considerations in drug development, as they may result in
adverse effects or provide therapeutic strategies. Thus, inhibi­
Transcriptional regulation
tion of NTCP can impair disposition of bile salts resulting in
higher systemic concentration of bile salts and consequent NTCP undergoes transcriptional as well as post‐transcriptional
adverse effects [23, 24]. regulation. Transcriptional regulation involving various nuclear
Drug–drug interaction (DDI) is an important determinant of receptors (NRs) has been reviewed by others [4, 42]. Briefly,
drug safety and efficacy. It is becoming evident that transporters NRs regulate NTCP expression by responding to changes in the
often work together with drug metabolizing enzymes in drug intracellular bile acid concentration (BAi). Increased BAi in
disposition. As a result, a role for transporters in drug dis­ cholestasis activates the farnesoid X receptor (FXR) response
position, therapeutic efficacy, and adverse drug reactions has element in the small heterodimer partner (SHP) promoter lead­
increasingly been recognized [25]. Some transporters are con­ ing to the expression of SHP, which in turn inhibits retinoid X
sidered to be clinically important in drug absorption and dis­ receptor (RXR)/retinoic acid receptor (RAR) activation of
position and therefore could mediate DDIs [26]. Although a NTCP (Figure 27.1). In addition, bile acid‐induced suppression
number of SLC transporters are considered to have clinical of hepatocyte nuclear factor (HNF1α), via activation of HNF4,
importance in the absorption and disposition of drugs [26], is involved in cholestasis‐induced downregulation of NTCP,
NTCP is not included. This is mainly because the clinical which is transcriptionally activated by HNF1α. Bile acids can
importance of these transporters is not well established. Since also suppress HNF1α via SHP. Downregulation of NTCP by
drugs can be transported by NTCP and can also inhibit solute bile salts serves to decrease BAi and thereby limit or minimize bile
transport by NTCP, a potential role of NTCP in DDI is very acid‐induced cellular toxicity in cholestasis. Proinflammatory
likely and has been reported [15, 27]. cytokines suppress RXR and may be involved in the loss of
27:  Basolateral Plasma Membrane Organic Anion Transporters 329

NTCP BA

LRH-1 RXR FXR


↑SHP

RAR RXR
↓NTCP

HNF1α

HNF4α BA

Figure 27.1  Transcriptional regulation of NTCP by nuclear receptors. Bile acids (BA) stimulate expression of SHP by activating FXR response
elements in the SHP promoters. SHP, in turn, interferes with RXR : RAR transactivation of NTCP promoter, resulting in decreased expression of
NTCP. HNF1α transcriptionally activates NTCP. BA decreases NTCP expression also by suppressing HNF1α via its inhibitory effect on HNF4α.
SHP also inhibits HNF4α‐mediated transactivation of the HNF1α promoter and autoregulates its own promoter by repressing LRH‐1 (liver receptor
homologue).

RXR, RARα, and SHP in bile duct ligated rats [43]. The down­ cAMP, increased by glucagon, may allow hepatocytes to
regulation of NTCP in pregnancy and by endotoxin may also be ­efficiently absorb bile salts returning from the intestine under
due to suppression of HNF1α and RXR : RAR [44, 45]. NTCP physiological conditions, while the decrease in uptake may
expression is stimulated by the glucocorticoid receptor (GR), ­protect hepatocytes from further increases in BAi observed in
the vitamin D receptor (VDR), the peroxisome proliferator‐acti­ cholestasis [57]. Functionally important polymorphisms in
vated receptor‐a (PPARa), signal transducer and activator of NTCP may result in decreased plasma membrane expression.
transcription 5, as well as various transcription factors and hor­ Multiple single nucleotide polymorphisms in NTCP are present
mones [42, 46]. in populations of European, African, Chinese, and Hispanic
Americans, and the altered transport activity of Ile223 to Thr
variant seen only in African Americans was due at least in part
Post‐translational regulation to decreased plasma membrane expression [58]. The resulting
Plasma membrane transporters, after translation, need to be phenotype in people with these mutations, however, is not
translocated to the plasma membrane to carry out their intended known.
functions. The level of a transporter in the plasma membrane Studies to date suggest that the post‐translational regulation
may be increased or decreased based on the need to regulate of NTCP involves plasma membrane translocation/retrieval.
solute transport activity. This is a highly regulated post‐transla­ Several signaling pathways, including cAMP, intracellular Ca2+,
tional event requiring involvement of various cellular signaling nitric oxide, phosphoinositide‐3‐kinase (PI3K), PKCs, and pro­
pathways. The post‐translational event also involves protein tein phosphatases are implicated in this process. Some of these
quality control to select and target dysfunctional proteins for pathways have previously been reviewed [21, 59–61] and are
degradation by the ubiquitin–proteasome system [47]. NTCP is briefly described below.
degraded by the ubiquitin–proteasome system at the level of The PI3K/Akt pathway plays an important role in cell ­survival
ER‐associated degradation. It is speculated that dysfunction of and translocation of hepatocellular transporters to the plasma
this pathway may result in intracellular NTCP deposits in chole­ membrane, indicating a beneficial role of PI3K in hepatic cells
static livers [48]. [60, 62–64]. PI3Ks are a family of lipid kinases (classes I, II,
Bile salt uptake by NTCP undergoes dynamic regulation. and III) that phosphorylate the inositol ring of phosphatidylin­
Cyclic AMP (cAMP) stimulates TC uptake [49–52] and increases ositides (PIs) at 3 position known as D3 phosphorylation [65].
the plasma membrane content of NTCP in rat hepatocytes and in The resulting phosphorylated PIs (PIPs), acting in concert
hepatic cells stably transfected with hNTCP [52–54]. In contrast, with phosphoinositide‐dependent kinases (PDKs), are involved
cholestatic bile salts (taurochenodeoxycholate [TCDC] and in the activation of downstream kinases, such as PKCζ/λ, Akt/
­taurolithocholate [TLC]) inhibit TC uptake [40, 55] and decrease PKB, and p70S6K [66]. The PI3K/Akt pathway is involved in
plasma membrane rNTCP [40, 56]. The increase in uptake by cell swelling and cAMP‐induced increases in TC uptake and
330 THE LIVER:  SODIUM‐TAUROCHOLATE COTRANSPORTING POLYPEPTIDE (NTCP)

↑Ca2+

↑PI3K
↑Ca2+/CAM

↑PKCζ ↑PKCδ
↑PP2B

NTCP NTCP NTCP pNTCP

Rab4-GTP Rab4-GDP
Kinesin1

Mt

Figure 27.2  Signaling pathways regulating NTCP translocation to the basolateral plasma membrane. Insertion into the plasma membrane is medi­
ated via two major pathways; increase in intracellular Ca2+ and activation of PI3K. Increases in intracellular Ca2+ leads to the activation of Ca2+‐
dependent protein phosphatase 2B (PP2B) via Ca2+‐calmodulin kinase (Ca2+‐CAM). PP2B dephosphorylates NTCP (pNTCP to NTCP) allowing
translocation to the plasma membrane by vesicular trafficking. Activation of PI3K leads to activations of PKCδ and PKCζ, which in turn induce
translocation of NTCP from the intracellular compartment to the plasma membrane by stimulating vesicular trafficking. NTCP containing vesicles
colocalize with Rab4, which cycles between GTP bound (Rab4‐GTP) active form and GDP bound (Rab4‐GDP) inactive form. Activation of PKCδ
leads to the conversion of inactive Rab4‐GDP to active Rab4‐GTP, which moves NTCP containing vesicles toward the plus end of microtubules
(Mt) by interacting with kinesin‐1. This movement is further facilitated by PKCζ. On the other hand, activation of PKCα by a cholestatic bile acid,
taurochenodeoxycholate (TCDC), stimulates retrieval of NTCP from the plasma membrane by an unknown mechanism except that inhibition of
PI3K enhances the effect of TCDC (not shown).

rNTCP translocation to the plasma membrane [67, 68] and cAMP motility of rNTCP‐containing vesicles along the microtubules
inhibits TLC‐induced inhibition of TC uptake by activating the [78]. Other studies showed that Rab4 facilitates cAMP‐induced
PI3K pathway [56]. NTCP translocation [52], NTCP‐containing vesicles colocalize
NTCP translocation to the plasma membrane is regulated by with Rab4 [78], cAMP‐induced rNTCP translocation is
phosphorylation and dephosphorylation, although kinase(s) dependent on PI3K [79], actin cytoskeleton and microtubules
involved in phosphorylation of NTCP has yet to be identified. [53, 79], and activation of PKCδ and PKCζ is PI3K dependent
NTCP is a serine/threonine phosphorylated protein [69] and [80]. Taken together, these results may suggest that activa­
cAMP‐induced increases in TC uptake and plasma membrane tion of PI3K/PKCδ by cAMP leads to activation of Rab4 in
rNTCP is associated with dephosphorylation [70] at Ser226 [71]. NTCP‐containing Rab4 vesicles and this is followed by
It is proposed (Figure 27.2) that cAMP activates protein phos­ PI3K/PKCζ‐mediated increased motility of NTCP/Rab4‐
phatase 2B (PP2B) by increasing intracellular Ca2+ [50, 72] and containing vesicles along the microtubules to the plasma
PP2B in turn dephosphorylates rNTCP facilitating its transloca­ membrane (Figure 27.2).
tion to the plasma membrane (Figure 27.2). Interestingly, inhi­ S‐nitrosylation also affects plasma membrane localization of
bition of PP2B reverses TCDC‐induced retrieval of rNTCP [40]. NTCP. Treatment with thiol‐binding agents inhibits TC uptake
Whether TCDC‐induced retrieval of rNTCP involves PP2B‐ in isolated rat hepatocytes [81, 82] and by hNTCP [83] indicat­
mediated NTCP dephosphorylation remains to be established. ing a role for cysteine residues in transport function. Nitric
Protein kinase C isoforms differentially regulate NTCP trans­ oxide (NO) inhibits TC uptake in rat hepatocytes [84], HuH7
location to and retrieval from the plasma membrane (Figure 27.2). cells stably transfected with hNTCP [85], and human hepato­
Thus, PKCδ [51, 73] and PKCζ [74] mediate cAMP‐induced cytes [86]. One of the ways that NO exerts its cellular effects is
rNTCP translocation, while conventional PKCs (most likely through a post‐translational modification termed S‐nitrosyla­
PKCα) mediate rNTCP retrieval [39, 40]. The activation of PKCδ tion, which involves binding of NO to reactive thiol groups on
and PKCζ by cAMP, but not TCDC‐induced activation of PKCα cysteine residues [87, 88]. Indeed, inhibition of TC uptake by
[40], is PI3K‐dependent [51, 74, 75]. On the other hand, TLC NO involves S‐nitrosylation of NTCP‐Cys96 and this is associ­
induces retrieval of rNTCP [56], but inhibits PKCα [76]. Thus, ated with a decrease in plasma membrane hNTCP [85, 89].
the retrieval of NTCP initiated by cholestatic bile acids may Such a mechanism may be involved in LPS‐induced cholestasis,
involve mediators in addition to PKCα. which is associated with a burst of NO production by iNOS in
Limited studies suggest a role for the small GTPase Rab4 liver cells [90]. NO also produces its cellular effects by nitration
and microtubules in NTCP translocation by PKCs. Rab4 cycles of tyrosine residues [91] and increases tyrosine nitration of
between GTP bound (Rab4‐GTP) active form and GDP bound NTCP [86]. Two tyrosine residues on the cytoplasmic tail of rat
(Rab4‐GDP) inactive form [77]. PKCδ facilitates hNTCP NTCP appear to be involved in targeting to the basolateral mem­
translocation by activating Rab4 [51] and PKCζ increases brane [92]. Thus, tyrosine nitration of NTCP by NO may also
27:  Basolateral Plasma Membrane Organic Anion Transporters 331

result in decreased translocation to the basolateral membrane Transport functions of OATPs


resulting in decreased TC uptake.
There are species differences in the regulation of plasma The oatps have broad and overlapping substrate specificities
membrane localization of hNTCP and rNTCP. TLC inhibits TC [11, 100]. Macrolide antibiotics, antihistamines, and statins are
uptake by hNTCP and rNTCP. However, TLC decreases plasma among the many drugs that are used clinically as well as various
membrane rNTCP but not hNTCP [56]. Like Bosentan [93], xenobiotics that can be transported by members of the oatp fam­
TLC inhibits TC uptake by rNTCP and hNTCP non‐competi­ ily. As oatps have ligand specificities that may overlap between
tively [55] and competitively [56], respectively. NO inhibits TC family members as well as other transporters (e.g. NTCP), stud­
uptake by hNTCP/rNTCP. However, NO decreases plasma ies to determine transport activity of a single member of the oatp
membrane level of hNTCP [85] and does not affect the plasma family performed in vitro may be difficult to extrapolate to what
membrane level of rNTCP [89]. In contrast, cAMP‐induced actually occurs in vivo in a living organism. Studies performed
increases in TC uptake are associated with translocation of in specific oatp knockout mouse models have provided some
hNTCP as well as rNTCP. These differences should be taken into insight into their physiologic function, although the resulting
account when extrapolating mechanisms regulating rNTCP to phenotypes are often subtle. For example, knockout of oatp1b2
hNTCP for the purpose of human drug development and also for resulted in reduced clearance of rifampicin, and lovastatin, but
understanding the mechanistic relevance in human diseases. not of pravastatin, simvastatin, rifamycin SV, or cerivastatin
[101]. Interestingly, reduced hepatic clearance of phalloidin and
microcystin resulted in reduced toxicity of these compounds in
oatp1b2 knockout mice, confirming that they are ligands for this
ORGANIC ANION TRANSPORT PROTEINS transporter [102]. Of note is the finding that there was a modest
(OATPs) increase in conjugated bilirubin in the circulation of these mice
and reduced clearance of dibromosulfophthalein (DBSP) [102,
Similar to hepatic uptake of bile acids, uptake of organic ani­ 103]. Reduced liver to plasma ratio of pravastatin was also noted
ons such as bilirubin and bromosulfophthalein (BSP) is rapid following continuous infusion to reach steady state, again show­
and has carrier‐mediated kinetics [94]. Studies performed in ing that changes in drug clearance can be subtle, requiring spe­
overnight cultured rat hepatocytes revealed saturable uptake cial studies to demonstrate them [104]. Studies in mice in which
of BSP that was inhibited by bilirubin [95, 96]. Ligand was Oatp1a1 or Oatp1a4 were knocked out showed reduced uptake
taken up by cells, while albumin remained extracellular [96]. of estradiol‐17beta‐D‐glucuronide, estrone‐3‐sulfate, and tauro­
Of interest was the finding that isosmotic substitution of NaCl cholic acid in primary hepatocytes prepared from both mouse
in medium by sucrose resulted in an over 80% reduction of strains [104].
BSP uptake [96]. This was not due to a requirement for extra­ Mice, in which all members of the Oatp1a and Oatp1b were
cellular Na+, as substitution of Na+ by K+, Li+, or choline had simultaneously deleted without organ specificity had markedly
no effect. However, replacement of Cl− by HCO3‐ or gluconate delayed plasma clearance of methotrexate and fexofenadine
reduced BSP uptake by approximately 40% [96]. Similar [13]. Additional studies showed that doxorubicin clearance was
results were found for bilirubin uptake by cultured rat hepato­ also reduced by 40% in these mice [105]. Interestingly, total
cytes as well as by isolated perfused rat liver. The basis for bile acid levels were increased in these mice and this was due
this chloride‐dependency is not clear. Studies performed with to increased plasma levels of unconjugated but not conjugated
36
Cl revealed that BSP uptake requires the presence of exter­ bile acids [13]. However, another study in which oatp1a1 or
nal Cl− and is not stimulated by unidirectional Cl− gradients, oatp1a4 were knocked out individually in mice revealed
suggesting that BSP transport is not coupled to Cl− transport reduced uptake of the conjugated bile acid taurocholic acid
[95]. These studies supported the existence of a hepatocyte [104]. Still other studies looking at transport mediated by rat
surface membrane organic anion transporter. Later studies oatp1a1 showed that it was an avid transporter of several con­
used the transport assay developed in cultured rat hepatocytes jugated bile acids, including tauromuricholic and taurocholic
as the basis for an expression cloning strategy in Xenopus lae- acids [106], the major bile acids found in mice [107]. The rea­
vis oocytes [97]. Injection of oocytes with total rat liver poly son for these seeming inconsistencies with respect to oatp‐
(A)+ RNA resulted in the functional expression of chloride‐ mediated bile acid transport remains unknown. The Oatp1a/
dependent BSP extraction from albumin. Using a subselection Oatp1b double knockout mice also had elevated levels of total
cloning protocol, a single cDNA was isolated [98]. The bilirubin that was due to increased conjugated but not unconju­
encoded protein was named organic anion transporting poly­ gated bilirubin [13]. Total bilirubin levels in single knockout
peptide 1 (oatp1), now known as oatp1a1. This protein medi­ mice for Oatp1a1 or Oatp1a4 were normal [104], although as
ates uptake of BSP and also mediates transport of various bile noted above, single knockout of Oatp1b2 resulted in a modest
salts such as taurocholate in a Na+‐independent fashion. Since elevation of conjugated bilirubin in female mice [102]. Elevated
the initial description of oatp1a1, over 20 additional members conjugated bilirubin levels in the double knockout mice was
of the oatp family have been described [99]. Amino acid phenotypically similar to the conjugated hyperbilirubinemia
sequences of these proteins have a high degree of homology. that has been described in patients with Rotor syndrome, a rare
In hepatocytes oatps are distributed on the basolateral disorder characterized by elevated plasma conjugated bilirubin
­(sinusoidal) plasma membrane. They are of similar size and with otherwise good health and normal routine liver function
have similar predicted membrane topologies and biochemical tests [108, 109]. These individuals have now been found to have
characteristics. simultaneous null mutations in the genes encoding OATP1B1
332 THE LIVER:  ORGANIC ANION TRANSPORT PROTEINS (OATPs)

and OATP1B3 [110]. It has been hypothesized that bilirubin Included among these are studies that implicate OATPs in clear­
glucuronides are effluxed from hepatocytes back into the circu­ ance of statins from the circulation. The importance of OATP
lation where they subsequently undergo reuptake mediated by polymorphisms on statin pharmacokinetics was epitomized in a
OATP1B1 or OATP1B3 [110]. In the absence of either one of study in which 85 patients out of 12 000 patients on simvastatin
these transporters, the other is able to mediate conjugated developed myopathy [125]. A genome‐wide association study
­bilirubin uptake. In the absence of both of these transporters, of these individuals revealed a strong association with a poly­
levels of conjugated bilirubin in the circulation rise as is seen in morphism in the gene encoding OATP1B1. Previous studies
Rotor syndrome patients and the corresponding knockout mouse showed that this valine to alanine polymorphism (Val174Ala)
model. Of note is the fact that the transporter(s) mediating encodes a protein that traffics poorly to the plasma membrane
uptake of unconjugated bilirubin has not as yet been identified and accumulates within cells [126]. It is a reasonable presump­
[111, 112]. tion that in these patients, reduced hepatocyte plasma membrane
levels of OATP1B1 resulted in reduced uptake of simvastatin
leading to elevated serum levels and a consequent increased
Regulation of non‐bile acid organic anion uptake by muscle cells causing the toxic effects. These data
transport point out the potential importance of transporter subcellular
There are a number of physiologic states in which organic anion trafficking on its transport activity. Importantly, other factors
transport may be perturbed. Many of these studies have been that lead to reduced transporter on the hepatocyte surface could
performed in animal models. These include liver regeneration in potentially result in toxicity even in the absence of a mutation in
rats resulting from two‐thirds partial hepatectomy in which the transporter itself.
influx of bilirubin and BSP, when determined independently of
reduced liver size, fall by approximately 50% within six hours Regulation of oatp subcellular distribution
and normalize by four days [113, 114], correlating with changes
in oatp1a1 mRNA and protein levels [115]. Similar modulation Seven of the eleven oatps that have been described in rat,
of transport and transporters is seen during development. For mouse, or human liver have PDZ consensus binding sites, as
example, BSP uptake is 70% lower in hepatocytes prepared defined by the sequences of their four C‐terminal amino acids
from three‐week‐old rats as compared to adults [116], consist­ (Table 27.1) [127]. These binding sites can mediate interaction
ent with finding of reduced oatp1a1 protein and mRNA expres­ of these oatps with one or more of the over 150 known PDZ
sion in liver for the first month after birth [116, 117]. In other domain‐containing proteins [128]. Interestingly, each of these
studies performed in rat hepatocytes, uptake of BSP was found PDZ domain‐containing proteins can have multiple binding
to be reduced in the presence of extracellular ATP [118]. domains and can form functionally important complexes with
Specifically, within minutes of exposure of cells to ATP, BSP their protein ligands [128]. The presence of a potential PDZ
uptake is reduced by 80% [118, 119]. Pharmacokinetic analysis consensus binding site on a protein, such as a member of the
revealed reduced Vmax and unchanged Km and this effect was oatp family, does not mean that this protein necessarily binds
found to be mediated by a purinergic receptor, resulting in phos­ to a PDZ domain‐containing protein, nor does it help in iden­
phorylation of oatp1a1 [119, 120]. Subsequent studies in trans­ tifying the potential binding protein. In particular the C‐termi­
fected cells and primary rat hepatocytes suggest that nal four amino acids of rat oatp1a1, KTKL, form a potential
phosphorylation of oatp1a1 results in its internalization and PDZ consensus binding sequence  [121, 127, 129]. Using a
consequent loss of transport activity [121]. synthetic peptide corresponding to the C‐terminal 16 amino
acids of oatp1a1 for affinity isolation, interacting proteins
from rat liver cytosol were purified. Protein mass fingerprint­
ing identified PDZK1 as the major interacting protein [127].
Pharmacogenomic studies
Using an otherwise identical synthetic peptide, but lacking the
A number of pharmacogenomic studies have identified single C‐terminal KTKL that comprises the PDZ consensus binding
nucleotide polymorphisms (SNPs) of OATPs that correlate with motif, no interacting proteins were resolved [127]. That
episodes of drug toxicity or altered liver uptake [122–124]. PDZK1 and oatp1a1 are bound to each other in vivo was shown

Table 27.1  Oatp family members found in rat, mouse, and human liver
NCBI accession C‐terminal Potential PDZ Plasma membrane
Species Original name New name number sequence consensus localization
Rat oatp1 oatp1a1 NM_017111 KTKL Yes Yes
Rat oatp2 oatp1a4 NM_131906 VTED No Yes
Rat oatp4 oatp1b2 NM_031650 ETPL Yes Yes
Rat oatp9 Oatp2b1 NM_080786 LQEL No ND
Mouse Oatp1 Oatp1a1 NM_013797 KTKL Yes Yes
Mouse Oatp2 Oatp1a4 NM_030687 KTKL Yes ND
Mouse Oatp4 Oatp1b2 NM_020495 ETPL Yes Yes
Human OATP‐A OATP1A2 NM_021094 KTKL Yes Yes
Human OATP‐C OATP1B1 NM_006446 ETHC No Yes
Human OATP8 OATP1B3 NM_019844 AAAN No Yes
Human OATP‐B OATP2B1 NM_007256 DSRV Yes Yes
27:  Basolateral Plasma Membrane Organic Anion Transporters 333

Figure 27.3  Working model for trafficking of endocytic vesicle‐associated oatp1a1 on microtubules. The C‐terminal 4 amino acids of oatp1a1
(KTKL) represent a PDZ binding consensus sequence that binds to PDZK1. PDZK1 has 4 independent PDZ binding sites and oatp1a1 binds to sites
1 and 3 as indicated, leaving sites 2 and 4 available to bind other proteins [127]. Interestingly, PDZK1 itself has a PDZ consensus binding motif at
its C‐terminus that can bind internally to PDZ domain 1 as well as binding other scaffold proteins such as EBP50 [132]. In the presence of PDZK1
endocytic vesicles containing oatp1a1 can interact with microtubules and move toward their plus end (cell surface) via activity of the motor mole­
cule kinesin‐1. Endocytic vesicles in the absence of PDZK1 or those that contain constructs of oatp1a1 lacking the terminal 4 amino acids and can
no longer bind PDZK1 associate with the molecular motor dynein that moves towards the minus end of microtubules (cell interior). Endocytic vesi­
cles containing oatp1a1 can also be associated with other cargo and we hypothesize that this can include oatps that lack a PDZ consensus binding
site and co‐traffic to and from the plasma membrane.

by their coimmunoprecipitation from rat liver lysates. Mouse movement of oatp1a1‐containing vesicles on microtubules
oatp1a1 has an 82% amino acid identity to rat oatp1a1 includ­ (Figure  27.3) [129]. Previous studies indicated that microtu­
ing the same KTKL amino acid sequence at its C‐terminus. bules in hepatocytes have their plus, rapidly growing ends near
Immunofluorescence studies performed in wild‐type and the cell surface, and minus ends within the cell at a microtubule
PDZK1 knockout mice revealed that plasma membrane distri­ organizing center [131].
bution of oatp1a1 was markedly reduced in the knockout mice Immunofluorescence studies performed in microchambers
[127]. Total expression of oatp1a1 in PDZK1 knockout mice showed the existence of an oatp1a1‐associated population of
was not reduced, as determined by Western blot, but it was endocytic vesicles prepared from wild‐type mouse liver [129].
located predominantly in intracellular vesicles on immunoflu­ These vesicles were also largely associated with PDZK1, bound
orescence analysis [127]. Corresponding to reduced plasma to polymerized microtubules in vitro, and moved approximately
membrane expression of oatp1a1 in the PDZK1 knockout equally toward the plus and minus microtubule ends upon addi­
mouse, the rate of plasma disappearance of 35S‐sulfobromoph­ tion of 50 μM ATP. A population of vesicles prepared from liv­
thalein (BSP), a ligand transported by oatp1a1, was reduced ers of PDZK1 knockout mice was also associated with oatp1a1.
by approximately 25% [127]. These vesicles bound to microtubules in vitro but upon ATP
To validate that interaction of oatp1a1 with PDZK1 is addition, their movement was highly biased toward the minus
required for its optimal trafficking to the plasma membrane, end, consistent with movement of vesicles to the cell interior.
studies were performed in HEK 293T cells transiently trans­ Further analysis showed that oatp‐associated vesicles were also
fected with GFP‐oatp1a1 with or without co‐transfection with associated with microtubule‐based motors. The identity of
PDZK1 [121]. These studies showed that oatp1a1 was present motors that were associated with these vesicles differed between
on the plasma membrane only in the presence of PDZK1 expres­ vesicles prepared from wild‐type vs PDZK1 knockout mice.
sion. Transfection with an oatp1a1 construct that did not code Kinesin‐1, a microtubule‐based plus end motor was highly asso­
for the terminal four amino acids revealed that it was no longer ciated with wild‐type vesicles, with little associated with
found on the cell surface even in the presence of PDZK1 [121]. ­vesicles prepared from PDZK1 knockout mice. Dynein, a
Trafficking to the cell surface was also reduced with a phospho­ microtubule‐based minus‐end motor, was largely associated
mimetic oatp1a1 construct, in which glutamic acid replaced the with PDZK1 knockout vesicles [129]. These studies suggest
serines at positions 634 and 635, irrespective of the presence of that PDZK1 on oatp1a1‐containing vesicles results in recruit­
PDZK1 [121]. This was related to increased internalization of ment of microtubule‐based motors that are not associated with
phosphomimetic oatp1a1 from the cell surface [121]. As these vesicles in its absence (Figure  27.3). Whether, in the
endocytic vesicles have been shown to traffic through cells on absence of PDZK1, oatp1a1 itself or other vesicle associated
­microtubules [41, 130], studies were performed to examine the proteins can recruit motors such as dynein is the subject of
hypothesis that subcellular trafficking of oatp1a1 requires ongoing studies.
334 THE LIVER:  REFERENCES

REFERENCES 25. Kramer, W. Transporters, Trojan horses and therapeutics: suitability of bile
acid and peptide transporters for drug delivery. Biol Chem, 2011;392:77–94.
26. Giacomini, K.M., Huang, S.M., Tweedie, D.J. et al. Membrane transporters
1. Anwer, M.S. and Hegner, D. Effect of Na+ on bile acid uptake by isolated rat
in drug development. Nat Rev Drug Discov, 2010;9:215–36.
hepatocytes. Hoppe‐Zeyler’s Z Physiol Chem, 1978;359:181–92.
27. Hebert, M.F., Townsend, R.W., Austin, S. et al. Concomitant cyclosporine
2. Hagenbuch, B., Lubbert, H., Stieger, B., and Meier, P.J. Expression of the
and micafungin pharmacokinetics in healthy volunteers. J Clin Pharmacol,
hepatocyte Na+/bile acid cotransporter in Xenopus laevis oocytes. J Biol
2005;45:954–60.
Chem, 1990;265:5357–60.
28. Yan, H., Zhong, G., Xu, G. et al. Sodium taurocholate cotransporting poly­
3. Hagenbuch, B., Stieger, B., Foguet, M. et al. Functional expression cloning
peptide is a functional receptor for human hepatitis B and D virus. Elife,
and characterization of the hepatocyte Na+/bile acid cotransport system. Proc
2012;1:1–28.
Natl Acad Sci U S A, 1991;88:10629–33.
29. Yan, H., Peng, B., He, W. et al. Molecular determinants of hepatitis B and d
4. Claro da, S.T., Polli, J.E., and Swaan, P.W. The solute carrier family 10
virus entry restriction in mouse sodium taurocholate cotransporting polypep­
(SLC10): beyond bile acid transport. Mol Aspects Med, 2013;34:252–69.
tide. J Virol, 2013;87:7977–91.
5. Doring, B., Lutteke, T., Geyer, J., and Petzinger, E. The SLC10 carrier fam­
30. Glebe, D. and Urban, S. Viral and cellular determinants involved in hepadna­
ily: transport functions and molecular structure. Curr Top Membr,
viral entry. World J Gastroenterol, 2007;13:22–38.
2012;70:105–68.
31. Gripon, P., Cannie, I., and Urban, S. Efficient inhibition of hepatitis B virus
6. Hagenbuch, B. and Dawson, P. The sodium bile salt cotransport family
infection by acylated peptides derived from the large viral surface protein.
SLC10. Pflugers Arch, 2004;447:566–70.
J Virol, 2005;79:1613–22.
7. Weinman, S.A. Electrogenicity of Na(+)‐coupled bile acid transporters. Yale
32. Meier, A., Mehrle, S., Weiss, T.S. et al. Myristoylated PreS1‐domain of the
J Biol Med, 1997;70:331–40.
hepatitis B virus L‐protein mediates specific binding to differentiated hepat­
8. Meier, P.J. and Stieger, B. Bile salt transporters. Annu Rev Physiol,
ocytes. Hepatology, 2013;58:31–42.
2002;64:635–61.
33. Petersen, J., Dandri, M., Mier, W. et al. Prevention of hepatitis B virus infec­
9. Stieger, B. The role of the sodium‐taurocholate cotransporting polypeptide
tion in vivo by entry inhibitors derived from the large envelope protein. Nat
(NTCP) and of the bile salt export pump (BSEP) in physiology and patho­
Biotechnol, 2008;26:335–41.
physiology of bile formation. Handb Exp Pharmacol, 2011;201:205–59.
34. Bogomolov, P., Alexandrov, A., Voronkova, N. et al. Treatment of chronic
10. Qiu, J.W., Deng, M., Cheng, Y. et  al. Sodium taurocholate cotransporting
hepatitis D with the entry inhibitor myrcludex B: first results of a phase Ib/
polypeptide (NTCP) deficiency: identification of a novel SLC10A1 mutation
IIa study. J Hepatol, 2016;65:490–8.
in two unrelated infants presenting with neonatal indirect hyperbilirubinemia
35. Xiao, F., McKeating, J.A., and Baumert, T.F. A bile acid transporter as a
and remarkable hypercholanemia. Oncotarget, 2017;8:106598–607.
candidate receptor for hepatitis B and D virus entry. J Hepatol,
11. Hagenbuch, B. and Stieger, B. The SLCO (former SLC21) superfamily of
2013;58:1246–8.
transporters. Mol Aspects Med, 2013;34:396–412.
36. Warner, N. and Locarnini, S. The new front‐line in hepatitis B/D research:
12. Csanaky, I.L., Lu, H., Zhang, Y. et al. Organic anion‐transporting polypep­
Identification and blocking of a functional receptor. Hepatology, 2013;58:9–12.
tide 1b2 (Oatp1b2) is important for the hepatic uptake of unconjugated bile
37. Schieck, A., Schulze, A., Gahler, C. et al. Hepatitis B virus hepatotropism is
acids: studies in Oatp1b2‐null mice. Hepatology, 2011;53:272–81.
mediated by specific receptor recognition in the liver and not restricted to
13. van de Steeg, E., Wagenaar, E., van der Kruijssen, C.M. et al. Organic anion
susceptible hosts. Hepatology, 2013;58:43–53.
transporting polypeptide 1a/1b‐knockout mice provide insights into hepatic
38. Lempp, F.A., Wiedtke, E., Qu, B. et al. Sodium taurocholate cotransporting
handling of bilirubin, bile acids, and drugs. J Clin Invest, 2010;120:2942–52.
polypeptide is the limiting host factor of hepatitis B virus infection in
14. Faber, K.N., Muller, M., and Jansen, P.L. Drug transport proteins in the liver.
macaque and pig hepatocytes. Hepatology, 2017;66:703–16.
Adv Drug Deliv Rev, 2003;55:107–24.
39. Stross, C., Helmer, A., Weissenberger, K. et  al. Protein kinase C induces
15. Yanni, S.B., Augustijns, P.F., Benjamin, D.K., Jr. et al. In vitro investigation
endocytosis of the sodium taurocholate cotransporting polypeptide. Am J
of the hepatobiliary disposition mechanisms of the antifungal agent
Physiol Gastrointest Liver Physiol, 2010;299:G320–8.
micafungin in humans and rats. Drug Metab Dispos, 2010;38:1848–56.
40. Muhlfeld, S., Domanova, O., Berlage, T. et al. Short‐term feedback regula­
16. Ho, R.H., Tirona, R.G., Leake, B.F. et al. Drug and bile acid transporters in
tion of bile salt uptake by bile salts in rodent liver. Hepatology,
rosuvastatin hepatic uptake: function, expression, and pharmacogenetics.
2012;56:2387–97.
Gastroenterology, 2006;130:1793–806.
41. Wang, X., Wang, P., Wang, W. et al. The Na(+)‐taurocholate cotransporting
17. Kullak‐Ublick, G.A., Glasa, J., Boker, C. et al. Chlorambucil‐taurocholate is
polypeptide traffics with the epidermal growth factor receptor. Traffic,
transported by bile acid carriers expressed in human hepatocellular carcino­
2016;17:230–44.
mas. Gastroenterology, 1997;113:1295–305.
42. Claudel, T., Zollner, G., Wagner, M., and Trauner, M. Role of nuclear recep­
18. Abel, J.J. and Rowntree, L.G. On the pharmacological action of some phtha­
tors for bile acid metabolism, bile secretion, cholestasis, and gallstone dis­
leins and their derivatives, with especial reference to their behavior as purga­
ease. Biochim Biophys Acta, 2011;1812:867–78.
tives. J Pharmacol Exp Ther, 1909;1:231–64.
43. Trauner, M. and Boyer, J.L. Bile salt transporters: molecular characteriza­
19. Hoekstra, L.T., de Graaf, W., Nibourg, G.A. et  al. Physiological and bio­
tion, function, and regulation. Physiol Rev, 2003;83:633–71.
chemical basis of clinical liver function tests: a review. Ann Surg,
44. Arrese, M., Trauner, M., Ananthanarayanan, M. et  al. Down‐regulation of
2013;257:27–36.
the Na+/taurocholate cotransporting polypeptide during pregnancy in the rat.
20. Stieger, B., Heger, M., de Graaf, W. et  al. The emerging role of transport
J Hepatol, 2003;38:148–55.
systems in liver function tests. Eur J Pharmacol, 2012;675:1–5.
45. Trauner, M., Arrese, M., Lee, H. et al. Endotoxin downregulates rat hepatic
21. Anwer, M.S. and Stieger, B. Sodium‐dependent bile salt transporters of the
ntcp gene expression via decreased activity of critical transcription factors.
SLC10A transporter family: more than solute transporters. Pflugers Arch,
J Clin Invest, 1998;101:2092–100.
2013;466:77–89.
46. Dawson, P.A., Lan, T., and Rao, A. Bile acid transporters. J Lipid Res,
22. Kouzuki, H., Suzuki, H., Stieger, B. et al. Characterization of the transport
2009;50:2340–57.
properties of organic anion transporting polypeptide 1 (oatp1) and Na(+)/
47. Claessen, J.H., Kundrat, L., and Ploegh, H.L. Protein quality control in the
taurocholate cotransporting polypeptide (Ntcp): comparative studies on the
ER: balancing the ubiquitin checkbook. Trends Cell Biol, 2012;22:22–32.
inhibitory effect of their possible substrates in hepatocytes and cDNA‐trans­
48. Kuhlkamp, T., Keitel, V., Helmer, A. et al. Degradation of the sodium tauro­
fected COS‐7 cells. J Pharmacol Exp Ther, 2000;292:505–11.
cholate cotransporting polypeptide (NTCP) by the ubiquitin–proteasome
23. Mita, S., Suzuki, H., Akita, H. et al. Inhibition of bile acid transport across
system. Biol Chem, 2005;386:1065–74.
Na+/taurocholate cotransporting polypeptide (SLC10A1) and bile salt export
49. Botham, K.M. and Suckling, K.E. The effect of dibutyryl cyclic AMP on the
pump (ABCB 11)‐coexpressing LLC‐PK1 cells by cholestasis‐inducing drugs.
uptake of taurocholic acid by isolated rat liver cells. Biochim Biophys Acta,
Drug Metab Dispos, 2006;34:1575–81.
1986;883:26–32.
24. McRae, M.P., Lowe, C.M., Tian, X. et  al. Ritonavir, saquinavir, and efa­
50. Grüne, S., Engelking, L.R., and Anwer, M.S. Role of intracellular calcium
virenz, but not nevirapine, inhibit bile acid transport in human and rat hepat­
and protein kinases in the activation of hepatic Na+/taurocholate cotransport
ocytes. J Pharmacol Exp Ther, 2006;318:1068–75.
by cyclic AMP. J Biol Chem, 1993;268:17734–41.
27:  Basolateral Plasma Membrane Organic Anion Transporters 335

51. Park, S.W., Schonhoff, C.M., Webster, C.R., and Anwer, M.S. Protein kinase 74. McConkey, M., Gillin, H., Webster, C.R., and Anwer, M.S. Cross‐talk
Cdelta differentially regulates cAMP‐dependent translocation of NTCP and between protein kinases Czeta and B in cyclic AMP‐mediated sodium tauro­
MRP2 to the plasma membrane. Am J Physiol Gastrointest Liver Physiol, cholate co‐transporting polypeptide translocation in hepatocytes. J Biol
2012;303:G657–65. Chem, 2004;279:20882–8.
52. Schonhoff, C.M., Thankey, K., Webster, C.R. et  al. Rab4 facilitates cyclic 75. Jiang, Y., Chen, C., Li, Z. et al. Characterization of the structure and function
adenosine monophosphate‐stimulated bile acid uptake and Na(+)‐taurocholate of a new mitogen‐activated protein kinase (p38beta). J Biol Chem,
cotransporting polypeptide translocation. Hepatology, 2008;48:1665–70. 1996;271:17920–6.
53. Dranoff, J.A., McClure, M., Burgstahler, A.D. et al. Short‐term regulation of 76. Beuers, U., Bilzer, M., Chittattu, A. et al. Tauroursodeoxycholic acid inserts
bile acid uptake by microfilament‐dependent translocation of ntcp to the the apical conjugate export pump, Mrp2, into canalicular membranes and
plasma membrane. Hepatology, 1999;30:223–9. stimulates organic anion secretion by protein kinase C‐dependent mecha­
54. Mukhopadhayay, S., Ananthanarayanan, M., Stieger, B. et  al. cAMP nisms in cholestatic rat liver. Hepatology, 2001;33:1206–16.
increases liver Na+‐taurocholate cotransport by translocating transporter to 77. Agola, J.O., Jim, P.A., Ward, H.H. et al. Rab GTPases as regulators of endo­
plasma membranes. Am J Physiol, 1997;273, G842–G848. cytosis, targets of disease and therapeutic opportunities. Clin Genet,
55. Schwenk, M., Schwarz, L.R., and Greim, H. Taurolithocholate inhibits tau­ 2011;80:305–18.
rocholate uptake by isolated hepatocytes at low concentrations. Naunyn 78. Sarkar, S., Bananis, E., Nath, S. et al. PKCzeta is required for microtubule‐based
Schmiedebergs Arch Pharmacol, 1977;298:175–9. motility of vesicles containing the ntcp transporter. Traffic, 2006;7:1078–91.
56. Schonhoff, C.M., Yamazaki, A., Hohenester, S. et al. PKC(epsilon)‐depend­ 79. Webster, C.R.L. and Anwer, M.S. Role of the PI3K/PKB signaling pathway
ent and ‐independent effects of taurolithocholate on PI3K/PKB pathway and in cAMP‐mediated translocation of rat liver Ntcp. Am J Physiol,
taurocholate uptake in HuH‐NTCP cell line. Am J Physiol Gastrointest Liver 1999;277:G1165–72.
Physiol, 2009;297:G1259–67. 80. Newton, A.C. Regulation of the ABC kinases by phosphorylation: protein
57. Prekeris, R., Klumperman, J., and Scheller, R.H. A Rab11/Rip11 protein kinase C as a paradigm. Biochem J, 2003;370:361–71.
complex regulates apical membrane trafficking via recycling endosomes. 81. Blumrich, M. and Petzinger, E. Membrane transport of conjugated and
Mol Cell, 2000;6:1437–48. unconjugated bile acids into hepatocytes is susceptible to SH‐blocking rea­
58. Ho, R.H., Leake, B.F., Roberts, R.L. et al. Ethnicity‐dependent polymorphism in gents. Biochim Biophys Acta, 1990;1029:1–12.
Na+‐taurocholate cotransporting polypeptide (SLC10A1) reveals a domain criti­ 82. Blumrich, M. and Petzinger, E. Two distinct types of SH‐groups are neces­
cal for bile acid substrate recognition. J Biol Chem, 2004;279:7213–22. sary for bumetanide and bile acid uptake into isolated rat hepatocytes.
59. Kosters, A. and Karpen, S.J. Bile acid transporters in health and disease. Biochim Biophys Acta, 1993;1149:278–84.
Xenobiotica, 2008;38:1043–71. 83. Hallen, S., Fryklund, J., and Sachs, G. Inhibition of the human sodium/bile
60. Anwer, M.S. Cellular Regulation of hepatic bile acid transport in health and acid cotransporters by side‐specific methanethiosulfonate sulfhydryl rea­
cholestasis. Hepatology, 2004;39:581–9. gents: substrate‐controlled accessibility of site of inactivation. Biochemistry,
61. Bouscarel, B., Kroll, S.D., and Fromm, H. Signal transduction and hepato­ 2000;39:6743–50.
cellular bile acid transport: cross talk between bile acids and second mes­ 84. Song, I.S., Lee, I.K., Chung, S.J. et al. Effect of nitric oxide on the sinusoidal
sengers. Gastroenterology, 1999;117:433–52. uptake of organic cations and anions by isolated hepatocytes. Arch Pharm
62. Rust, C., Bauchmuller, K., Fickert, P. et al. Phosphatidylinositol 3‐kinase‐ Res, 2002;25:984–8.
dependent signaling modulates taurochenodeoxycholic acid‐induced liver 85. Schonhoff, C.M., Ramasamy, U., and Anwer, M.S. Nitric oxide‐mediated
injury and cholestasis in perfused rat livers. Am J Physiol Gastrointest Liver inhibition of taurocholate uptake involves S‐nitrosylation of NTCP. Am J
Physiol, 2005;289:G88–94. Physiol Gastrointest Liver Physiol, 2011;300:G364–70.
63. Webster, C.R., Usechak, P., and Anwer, M.S. cAMP inhibits bile acid‐ 86. Gonzalez, R., Cruz, A., Ferrin, G. et al. Nitric oxide mimics transcriptional
induced apoptosis by blocking caspase activation and cytochrome c release. and post‐translational regulation during alpha‐tocopherol cytoprotection
Am J Physiol Gastrointest Liver Physiol, 2002;283:G727–38. against glycochenodeoxycholate‐induced cell death in hepatocytes. J Hepatol,
64. Gates, A., Hohenester, S., Anwer, M.S., and Webster, C.R. cAMP‐GEF cyto­ 2011;55:133–44.
protection by Src tyrosine kinase activation of phosphoinositide‐3‐kinase 87. Hess, D.T. and Stamler, J.S. Regulation by S‐nitrosylation of protein post­
p110 {beta}/{alpha} in rat hepatocytes. Am J Physiol Gastrointest Liver translational modification. J Biol Chem, 2012;287(7):4411–8.
Physiol, 2009;296:G764–74. 88. Schonhoff, C.M., Gaston, B., and Mannick, J.B. Nitrosylation of cytochrome
65. Cantley, L.C. he phosphoinositide 3‐kinase pathway. Science, 2002; 296: c during apoptosis. J Biol Chem, 2003;278:18265–70.
1655–7. 89. Ramasamy, U., Anwer, M.S., and Schonhoff, C.M. Cysteine 96 of Ntcp is
66. Toker, A. Protein kinases as mediators of phosphoinositide 3‐kinase signal­ responsible for NO‐mediated inhibition of taurocholate uptake. Am J Physiol
ing. Mol Pharmacol, 2000;57:652–8. Gastrointest Liver Physiol, 2013;305:G513–19.
67. Webster, C.R., Srinivasulu, U., Ananthanarayanan, M. et al. rotein kinase B/ 90. Geier, A., Fickert, P., and Trauner, M. Mechanisms of disease: mechanisms
Akt mediates cAMP‐ and cell swelling‐stimulated Na+/taurocholate cotrans­ and clinical implications of cholestasis in sepsis. Nat Clin Pract Gastroenterol
port and Ntcp translocation. J Biol Chem, 2002;277:28578–83. Hepatol, 2006;3:574–85.
68. Webster, C.R.L., Blanch, C.J., Philips, J., and Anwer, M.S. ell swelling‐ 91. Ischiropoulos, H. Protein tyrosine nitration—an update. Arch Biochem
induced translocation of rat liver Na+/taurocholate cotransport polypeptide is Biophys, 2009;484:117–21.
mediated via the phosphoinositide 3‐kinase signaling pathway. J Biol Chem, 92. Sun, A.Q., Arrese, M.A., Zeng, L. et al. The rat liver Na(+)/bile acid cotrans­
2000;275:29754–60. porter. Importance of the cytoplasmic tail to function and plasma membrane
69. Mukhopadhayay, S., Ananthanarayanan, M., Stieger, B. et al. Sodium tauro­ targeting. J Biol Chem, 2001;276:6825–33.
cholate cotransporting polypeptide is a serine, threonine phosphoprotein and 93. Leslie, E.M., Watkins, P.B., Kim, R.B., and Brouwer, K.L. Differential inhi­
is dephosphorylated by cyclic AMP. Hepatology, 1998;28:1629–36. bition of rat and human Na+‐dependent taurocholate cotransporting poly­
70. Mukhopadhayay, S., Webster, C.R.L., and Anwer, M.S. Role of protein peptide (NTCP/SLC10A1)by bosentan: a mechanism for species differences
phosphatase in cyclic AMP‐mediated stimulation of hepatic Na+/taurocho­ in hepatotoxicity. J Pharmacol Exp Ther, 2007;321:1170–8.
late cotransport. J Biol Chem, 1998;273:30039–45. 94. Scharschmidt, B.F., Waggoner, J.G., and Berk, P.D. Hepatic organic anion
71. Anwer, M.S., Gillin, H., Mukhopadhyay, S. et  al. Dephosphorylation of uptake in the rat. J Clin Invest, 1975;56:1280–92.
Ser‐226 facilitates plasma membrane retention of Ntcp. J Biol Chem, 95. Min, A.D., Johansen, K.L., Campbell, C.G., and Wolkoff, A.W. Role of chlo­
2005;280:33687–92. ride and intracellular pH on the activity of the rat hepatocyte organic anion
72. Webster, C.R.L., Blanch, C., and Anwer, M.S. Role of PP2B in cAMP‐ transporter. J Clin Invest, 1991;87:1496–502.
induced dephosphorylation and translocation of NTCP. Am J Physiol 96. Wolkoff, A.W., Samuelson, A.C., Johansen, K.L. et al. Influence of Cl‐ on
Gastrointest Liver Physiol, 2002;283:G44–50. organic anion transport in short‐term cultured rat hepatocytes and isolated
73. Schonhoff, C.M., Gillin, H., Webster, C.R., and Anwer, M.S. Protein kinase perfused rat liver. J Clin Invest, 1987;79:1259–68.
Cdelta mediates cyclic adenosine monophosphate‐stimulated translocation 97. Jacquemin, E., Hagenbuch, B., Stieger, B. et al. Expression of the hepatocel­
of sodium taurocholate cotransporting polypeptide and multidrug resistant lular chloride‐dependent sulfobromophthalein uptake system in Xenopus
associated protein 2 in rat hepatocytes. Hepatology, 2008;47:1309–16. laevis oocytes. J Clin Invest, 1991;88:2146–9.
336 THE LIVER:  REFERENCES

98. Jacquemin, E., Hagenbuch, B., Stieger, B. et al. Expression cloning of a rat 115. Gerloff, T., Geier, A., Stieger, B. et al. Differential expression of basolateral
liver Na(+)‐independent organic anion transporter. Proc Natl Acad Sci U S and canalicular organic anion transporters during regeneration of rat liver.
A, 1994;91:133–7. Gastroenterology, 1999;117:1408–15.
99. Hagenbuch, B. and Gui, C. Xenobiotic transporters of the human organic 116. Angeletti, R.H., Bergwerk, A.J., Novikoff, P.M., and Wolkoff, A.W.
anion transporting polypeptides (OATP) family. Xenobiotica, Dichotomous development of the organic anion transport protein in liver
2008;38:778–801. and choroid plexus. Am J Physiol, 1998;275:C882–7.
100. Zhao, W., Zitzow, J.D., Weaver, Y. et al. Organic anion transporting poly­ 117. Dubuisson, C., Cresteil, D., Desrochers, M. et al. Ontogenic expression of
peptides contribute to the disposition of perfluoroalkyl acids in humans and the Na(+)‐independent organic anion transporting polypeptide (oatp) in rat
rats. Toxicol Sci, 2017;156:84–95. liver and kidney. J Hepatol, 1996;25:932–40.
101. Chen, C., Stock, J.L., Liu, X. et al. Utility of a novel Oatp1b2 knockout 118. Campbell, C.G., Spray, D.C., and Wolkoff, A.W. Extracellular ATP4‐ mod­
mouse model for evaluating the role of Oatp1b2 in the hepatic uptake of ulates organic anion transport by rat hepatocytes. J Biol Chem,
model compounds. Drug Metab Dispos, 2008;36:1840–5. 1993;268:15399–404.
102. Lu, H., Choudhuri, S., Ogura, K. et  al. Characterization of organic 119. Glavy, J.S., Wu, S.M., Wang, P.J. et al. Down‐regulation by extracellular
anion transporting polypeptide 1b2‐null mice: essential role in hepatic ATP of rat hepatocyte organic anion transport is mediated by serine phos­
uptake/toxicity of phalloidin and microcystin‐LR. Toxicol Sci, 2008; phorylation of Oatp1. J Biol Chem, 2000;275:1479–84.
103:35–45. 120. Xiao, Y., Nieves, E., Angeletti, R.H. et al. Rat organic anion transporting
103. Zaher, H., Meyer zu Schwabedissen, H.E., Tirona, R.G. et al. Targeted dis­ protein 1A1 (Oatp1a1): purification and phosphopeptide assignment.
ruption of murine organic anion‐transporting polypeptide 1b2 (Oatp1b2/ Biochemistry, 2006;45:3357–69.
Slco1b2) significantly alters disposition of prototypical drug substrates 121. Choi, J.H., Murray, J.W., and Wolkoff, A.W. PDZK1 binding and serine
pravastatin and rifampin. Mol Pharmacol, 2008;74:320–9. phosphorylation regulate subcellular trafficking of organic anion transport
104. Gong, L., Aranibar, N., Han, Y.H. et al. Characterization of organic anion‐ protein 1a1. Am J Physiol Gastrointest Liver Physiol, 2011;300:G384–93.
transporting polypeptide (Oatp) 1a1 and 1a4 null mice reveals altered trans­ 122. Ho, R.H., Choi, L., Lee, W. et al. Effect of drug transporter genotypes on
port function and urinary metabolomic profiles. Toxicol Sci, 2011; pravastatin disposition in European‐ and African‐American participants.
122:587–97. Pharmacogenet Genomics, 2007;17:647–56.
105. Lee, H.H., Leake, B.F., Kim, R.B., and Ho, R.H. Contribution of organic 123. Miura, M., Satoh, S., Inoue, K. et al. Influence of SLCO1B1, 1B3, 2B1 and
anion‐transporting polypeptides 1A/1B to doxorubicin uptake and clear­ ABCC2 genetic polymorphisms on mycophenolic acid pharmacokinetics
ance. Mol Pharmacol, 2017;91:14–24. in Japanese renal transplant recipients. Eur J Clin Pharmacol,
106. Hata, S., Wang, P., Eftychiou, N. et al. Substrate specificities of rat oatp1 2007;63:1161–9.
and ntcp: implications for hepatic organic anion uptake. Am J Physiol 124. Gong, I.Y. and Kim, R.B. Impact of genetic variation in OATP transporters to
Gastrointest Liver Physiol, 2003;285:G829–39. drug disposition and response. Drug Metab Pharmacokinet, 2013;28:4–18.
107. Wang, D.Q., Tazuma, S., Cohen, D.E., and Carey, M.C. Feeding natural 125. Link, E., Parish, S., Armitage, J. et al. SLCO1B1 variants and statin‐induced
hydrophilic bile acids inhibits intestinal cholesterol absorption: studies in myopathy—a genomewide study. N Engl J Med, 2008;359:789–99.
the gallstone‐susceptible mouse. Am J Physiol Gastrointest Liver Physiol, 126. Tirona, R.G., Leake, B.F., Merino, G., and Kim, R.B. Polymorphisms in
2003;285:G494–502. OATP‐C: identification of multiple allelic variants associated with altered
108. Fu, D., Wakabayashi, Y., Lippincott‐Schwartz, J., and Arias, I.M. Bile acid transport activity among European‐ and African‐Americans. J Biol Chem,
stimulates hepatocyte polarization through a cAMP‐Epac‐MEK‐LKB1‐ 2001;276:35669–75.
AMPK pathway. Proc Natl Acad Sci U S A, 2011;108:1403–8. 127. Wang, P., Wang, J.J., Xiao, Y. et al. Interaction with PDZK1 is required for
109. Wolkoff, A. The hyperbilirubinemias, in Harrison’s Principles of Internal expression of organic anion transporting protein 1A1 on the hepatocyte sur­
Medicine, (eds. H.L. Jameson et al.). McGraw Hill, New York, 2018. face. J Biol Chem, 2005;280:30143–9.
110. van de Steeg, E., Stranecky, V., Hartmannova, H. et al. Complete OATP1B1 128. Ye, F. and Zhang, M. Structures and target recognition modes of PDZ domains:
and OATP1B3 deficiency causes human Rotor syndrome by interrupting con­ recurring themes and emerging pictures. Biochem J, 2013;455:1–14.
jugated bilirubin reuptake into the liver. J Clin Invest, 2012;122:519–28. 129. Wang, W.J., Murray, J.W., and Wolkoff, A.W. Oatp1a1 requires PDZK1 to
111. Wang, P., Kim, R.B., Chowdhury, J.R., and Wolkoff, A.W. The human traffic to the plasma membrane by selective recruitment of microtubule‐
organic anion transport protein SLC21A6 is not sufficient for bilirubin based motor proteins. Drug Metab Dispos, 2014;42:62–9.
transport. J Biol Chem, 2003;278:20695–9. 130. Mukhopadhyay, A., Quiroz, J.A., and Wolkoff, A.W. Rab1a regulates sort­
112. Cvorovic, J. and Passamonti, S. Membrane transporters for bilirubin and its ing of early endocytic vesicles. Am J Physiol Gastrointest Liver Physiol,
conjugates: a systematic review. Front Pharmacol, 2017;8:887. 2014;306:G412–24.
113. Oleaga, A., Gonzalez, J., and Esteller, A. Effects of two‐thirds hepatectomy 131. Novikoff, P.M., Cammer, M., Tao, L. et al. Three‐dimensional organization
on sulfobromophthalein handling by the rat liver. Comp Biochem Physiol A of rat hepatocyte cytoskeleton: relation to the asialoglycoprotein endocyto­
Comp Physiol, 1987;87:13–9. sis pathway. J Cell Sci, 1996;109(1):21–32.
114. Gartner, U., Stockert, R.J., Morell, A.G., and Wolkoff, A.W. Modulation of 132. LaLonde, D.P. and Bretscher, A. The scaffold protein PDZK1 undergoes a
the transport of bilirubin and asialoorosomucoid during liver regeneration. head‐to‐tail intramolecular association that negatively regulates its interac­
Hepatology, 1981;1:99–106. tion with EBP50. Biochemistry, 2009;48:2261–71.
Hepatic Nuclear Receptors
28 Raymond E. Soccio
University of Pennsylvania, Perelman School of Medicine, Department of Medicine, Division of
Endocrinology, Diabetes, and Metabolism, Institute for Diabetes, Obesity, and Metabolism, Philadelphia
Crescenz VA Medical Center, Philadelphia, PA, USA

INTRODUCTION historical artifacts, many NR names are uninformative (like liver X


receptor) or even misleading. For example, farnesoid X receptor
Nuclear receptors (NRs) are important regulators of develop- (FXR) binds with high affinity to bile acids not farnesol deriva-
ment and homeostasis, and many family members have vital tives, and despite attempts to rename it bile acid receptor (BAR)
roles in the liver. NRs are transcription factors whose function is [5] the name FXR has persisted. This chapter will generally refer
reflected in their bipartite structure: the DNA‐binding domain to to NR names in common use, and hepatic NRs will be broadly
directly bind regulatory DNA and the ligand‐binding domain to divided into five classes based on their functions and ligands:
sense hormones, metabolites, or drugs and modulate expression endocrine, xenobiotic, metabolic, circadian, and other (Table 28.1).
of nearby genes [1–3]. Thus, NRs directly sense the environ- The sixth class of NRs have undetectable expression in liver and
ment to regulate expression of target genes. Their ligands are are thus not covered here: GCNF is restricted to germ cells, TLX
small molecules falling into three classes: hormones which cir- to brain, and PNR to retina. Likewise, SF‐1 and DAX1 are
culate to affect distant organs, molecules generated locally by restricted to steroidogenic tissues, though their close relatives
normal metabolism, or exogenous compounds like drugs. LRH1 and SHP have key functions in liver.
Sensing and responding to hormones, metabolites, and drugs is The majority of the 48 human NRs are expressed in liver
an essential function of the liver to maintain homeostasis – the (Figure 28.1). Human mRNA expression across multiple tissues
constant internal environment  –  in the face of changes to the has been reported by several consortia, including the Human
external environment. One critical aspect of the external Protein Atlas [6] and Genotype‐Tissue Expression (GTEx) [7].
environment to maintain life is nutrient availability, so many Some NRs, such as the glucocorticoid receptor (GR), are
liver NRs directly or indirectly regulate energy metabolism in broadly expressed across all tissues, such that hepatic expres-
the fed or fasted state [2]. NRs are also notable for their pharma- sion mediates the ligand action in liver. Other NRs have expres-
cology, with “druggable” ligand binding domains such that 13% sion limited to liver and other tissues where the ligand acts, such
of all FDA‐approved drugs target NRs [1]. as FXR being mainly restricted to liver and other gastrointesti-
The 48 human NRs have been classified into subfamilies by nal tissue related to bile acids their enterohepatic circulation.
their sequence homology. This system is reflected by the names One NR, the xenobiotic receptor CAR, is expressed primarily in
(i.e. NR1A1) from the NR Nomenclature Committee [4], but most the liver with low or undetectable expression elsewhere, such
NRs’ names and official gene symbols still relate to their original that its main effects are hepatic. In many NR groups, one family
discoveries. The field of NR study began in the 1980s, with the member predominates in liver. For instance, the β isoform of
molecular cloning of several hormone receptors revealing a com- thyroid hormone receptor (THR) is the major liver form, and
mon structure [1, 3]. For classic endocrine receptors, names accu- only one of the three estrogen related receptors (ERRα) has
rately describe the hormones that serves as ligands: thyroid appreciable human liver expression. Finally, given that hepato-
hormone, steroid hormones, and vitamin D. Other nuclear recep- cytes constitute >90% of liver mass, this chapter mainly
tors were first identified as “orphans” lacking known ligands, addresses NR functions in hepatocytes, though some important
though ligands have since been identified for many [2]. Due to roles in non‐parenchymal liver cells are also discussed.

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
Class Group NRNC Name Gene HPA (TPM) GTEx (RPKM)
Thyroid Hormone NR1A2 THRβ Thyroid hormone receptor β THRB 11.6 2.7
Receptor NR1A1 THRα Thyroid hormone receptor α THRA 2.5 1.8
NR3C1 GR Glucocorticoid receptor NR3C1 20.6 4.9
3-ketosteroid NR3C2 MR Mineralocorticoid receptor NR3C2 4.4 1.4
Endocrine receptors NR3C3 PR Progesterone receptor PGR 0.1 0.0
NR3C4 AR Androgen receptor AR 21.6 8.3
NR3A1 ERα Estrogen receptor α ESR1 8.8 1.9
Estrogen Receptor
NR3A2 ERβ Estrogen receptor β ESR2 0.0 0.1
NR1I1 VDR Vitamin D receptor VDR 0.5 0.2
Vitamin D receptor-
NR1I2 PXR Pregnane X receptor NR1I2 20.0 23.3
Xenobiotic like
NR1I3 CAR Constitutive androstane receptor NR1I3 82.5 52.3
NR1H3 LXRα Liver X receptor α NR1H3 26.7 19.2
Liver X Receptor-like NR1H2 LXRβ Liver X receptor β NR1H2 11.0 14.3
NR1H4 FXR Farnesoid X receptor NR1H4 66.6 22.3
Peroxisome proliferator-activated
NR1C1 PPARα receptor α PPARA 14.7 11.7
Peroxisome
proliferator Peroxisome proliferator-activated
Activated receptor NR1C2 PPARβ/δ receptor δ PPARD 2.9 8.4
Peroxisome proliferator-activated
NR1C3 PPARγ receptor γ PPARG 8.6 2.0
NR1B1 RARα Retinoic acid receptor α RARA 11.8 9.8
Retinoic acid
Metabolic receptor NR1B2 RARβ Retinoic acid receptor β RARB 3.3 0.8
NR1B3 RARγ Retinoic acid receptor γ RARG 0.9 0.8
NR2B1 RXRα Retinoid X receptor α RXRA 7.2 22.7
Retinoid X receptor NR2B2 RXRβ Retinoid X receptor β RXRB 1.8 14.2
NR2B3 RXRγ Retinoid X receptor γ RXRG 0.0 0.1
NR5A2 LRH1 Liver receptor homolog 1 NR5A2 19.5 4.8
LRH1/SF1
NR5A1 SF1 Steroidogenic factor 1 NR5A1 0.0 0.0
NR0B2 SHP Small heterodimer partner NR0B2 48.8 46.9
Dosage-sensitive sex reversal,
SHP/DAX
adrenal hypoplasia critical region,
NR0B1 DAX1 on chromosome X, gene 1 NR0B1 0.0 0.0
NR1D1 REVERBα Rev-ErbAα NR1D1 6.7 14.1
Rev-erb
NR1D2 REVERBβ Rev-ErbAβ NR1D2 15.4 4.9
NF1F1 RORα RAR-related orphan receptor α RORA 14.5 4.6
RAR-related orphan
NF1F2 RORβ RAR-related orphan receptor β RORB 0.0 0.0
Circadian receptor
NF1F3 RORγ RAR-related orphan receptor γ RORC 43.0 18.4
NR3B1 ERRα Estrogen-related receptor α ESRRA 11.3 24.3
Estrogen Receptor
NR3B2 ERRβ Estrogen-related receptor β ESRRB 0.0 0.0
Related
NR3B3 ERRγ Estrogen-related receptor γ ESRRG 0.3 0.0
Hepatocyte nuclear NR2A1 HNF4α Hepatocyte nuclear factor 4 α HNF4A 51.0 39.5
factor 4 NR2A2 HNF4γ Hepatocyte nuclear factor 4 γ HNF4G 6.5 2.1
NR2C1 TR2 Testicular receptor 2 NR2C1 2.9 4.7
NR2C/TR
NR2C2 TR4 Testicular receptor 4 NR2C2 3.2 1.0
Chicken ovalbumin upstream
NR2F1 COUP-TF1 promoter-transcription factor I NR2F1 5.7 4.2
Other
NR2F/COUP Chicken ovalbumin upstream
NR2F2 COUP-TF2 promoter-transcription factor II NR2F2 16.6 6.4
NR2F6 EAR2 V-erbA-related NR2F6 12.2 33.9
NR4A1 NGF1B Nerve growth factor IB NR4A1 20.7 16.9
NR4A NR4A2 NURR1 Nuclear receptor related 1 NR4A2 5.3 6.2
NR4A3 NOR1 Neuron-derived orphan receptor 1 NR4A3 5.3 0.9
GCNF NR6A1 GCNF Germ cell nuclear factor NR6A1 0.5 0.9
Homologue of the Drosophila
Non-liver NR2E1 TLX tailless gene NR2E1 0.0 0.0
TLX/PNR
Photoreceptor cell-specific
NR2E3 PNR nuclear receptor NR2E3 0.0 0.0

Figure 28.1  48 Members of the human nuclear receptor superfamily. In this chapter, the 48 human NRs are divided into these six classes based
on their functions in liver. Based on sequence homology, the Nuclear Receptor Nomenclature Committee (NRNC) has divided NRs into 19 groups
and assigned standardized names. However, NRs are commonly called other names, and these abbreviated and full names are shown. The official
human gene symbols reflect either common or NRNC names. Gene expression profiling across human tissues has been reported by several consor-
tia including the Human Protein Atlas (HPA) and Genotype‐Tissue Expression (GTEx) [6]. The normalized expression of each human NR in liver
is thus quantified by RNA‐sequencing as either transcripts per million (TPM) or reads per kilobase per million mapped reads (RPKM). The color
scaling in these columns reflect hepatic expression levels (green: high, yellow: middle, red: low). NRs are listed in the order they are presented in
this chapter, with the 21 highlighted in yellow discussed most extensively due to their high expression and/or well‐studied roles in liver.
28:  Hepatic Nuclear Receptors 339

Type 1 Type 2 Type 3 Type 4


Homodimers with RXR heterodimers with Exclusively homodimers
Monomers
ligand-dependent ligand-dependent in nucleus with
(or Homodimers)
nuclear localization transcriptional activation uncertain ligand effects

LBD GR GR THR RXR HNF4 HNF4 ROR

DBD
DNA

Inverted Direct Direct Half Site


Repeat Repeat Repeat

Endocrine/Steroid: Endocrine: THR, VDR Other (NR2 family): Circadian:


GR, AR, ER Metabolic: LXR, FXR, HNF4, COUP-TF, TR Reverb, ROR, ERR
PPAR, RAR (RXR homodimers)
Xenobiotic:
CAR, PXR
Figure 28.2  Four mechanistic types of nuclear receptors (NRs). NRs have been classified into these four mechanistic types based on their dimeri-
zation, response elements, and ligand dependence. Notably, the four types correspond roughly to their functional classifications in liver. Type 1
receptors are the endocrine receptors for steroid hormones. Type 2 receptors include both the endocrine receptors for thyroid hormone and vitamin
D, as well as the four metabolic receptors with prominent roles in liver (LXRs, FXR, PPARs, and RARs). The endocrine RXR heterodimers are
“non‐permissive” such that RXR ligands are generally silent, while the metabolic ones are “permissive” such that they can also be activated by RXR
agonists. The xenobiotic receptors have features of both type 1 and type 2, the circadian receptors are generally type 4, and the other orphan recep-
tors of the NR2 family are type 3. LBD, ligand binding domain; DBD, DNA binding domain. Figure adapted from [3].

NRs have also been classified into mechanistic types based ENDOCRINE NUCLEAR HORMONE
on their DNA response elements, dimerization, and subcellu- RECEPTORS IN LIVER
lar localization in the absence of ligand (Table 28.1, Figure
28.2). Each NR monomer binds to a core 6 nucleotide consen-
sus sequence (AGGTCA or related sequences), thus NR
Thyroid hormone receptors (THRs)
dimers bind to response elements consisting of two such “half Thyroid hormones exert effects throughout the body, including
sites”. For different dimers, the two half sites have variable prominent actions on energy metabolism and the cardiovascular
orientation (direct or inverted) and spacing between them (1–5 system. The liver plays key roles in thyroid hormone physiol-
nucleotides), giving rise to names like direct repeat 1 (DR‐1) ogy, both as a target of local effects and a mediator of systemic
and inverted repeat 2 (IR‐2). Steroid hormone receptors like effects [8]. Two THR genes encode the THRα and THRβ iso-
GR are type 1 NRs. In the absence of ligand, type 1 NRs reside forms. While most tissues express both, THRβ is the major
predominantly in the cytosol associated with heat shock pro- receptor in liver (as well as brain), while THRα predominates in
teins (HSPs). Ligand binding triggers dissociation of HSPs, the cardiovascular system and bone.
homodimerization, and nuclear translocation to bind inverted In the liver, thyroid hormone regulates many target genes
repeat DNA response elements and activate target genes. Other with key roles in lipid metabolism [9]. Some transcriptional
endocrine NRs and the metabolic NRs are type 2, constitu- effects of thyroid hormone are direct via THRs, while oth-
tively nuclear and binding to various direct repeats as heterodi- ers may be indirect as THR can regulate other key transcrip-
mers with retinoid X receptors (RXRs). In the absence of tion factors in lipid metabolism including sterol regulatory
ligand, these RXR heterodimers associate with corepressor element binding proteins (SREBP1c regulating fatty acid
proteins and actively decrease expression of target genes – giv- synthesis and SREBP2 cholesterol synthesis), carbohy-
ing lower expression than even the basal level in the absence drate‐responsive element binding proteins (ChREBPs), and
of receptor. Ligand binding triggers a conformation change, metabolic NRs (like LXRs, PPARs, FXRs). Some THR tar-
with release of corepressors and binding of coactivators. This gets, such as the highly‐regulated THRSP (thyroid hormone
simplified “coregulator switch model” of type 2 NRs like THR responsive protein, also called Spot14), are involved in de
is well‐validated in vitro, though in vivo relevance remains to novo lipogenesis, the synthesis of fatty acids from glucose.
be proven [2]. Type 3 NRs, mainly the NR2 family members However, thyroid hormones also activate fatty acid oxida-
(HNF4s, COUP‐TFs, and TRs), are homodimers with consti- tion, and this catabolic effect predominates in hyperthy-
tutive nuclear localization and direct repeat binding. Finally, roidism to reduce liver fat. Thyroid hormones similarly
the circadian NRs (Rev‐erbs, RORs, ERRs) are type 4, which affect both cholesterol synthesis and clearance, with the lat-
can bind DNA as monomers to half sites and/or as dimers to ter apparently predominating, including effects on the LDL
direct repeats. receptor activity and bile acid synthesis via cholesterol
Table 28.1  Hepatic nuclear receptor (NR) ligands, dimerization, and response elements. For the five classes of hepatic nuclear receptor defined here, three key features are shown for each NR or
group of NR isoforms: ligands, dimerization status, and response elements. These features define the mechanistic type of NR (types 1–4, see Figure 28.2). In general, there are well‐defined ligands
for endocrine, xenobiotic, and metabolic NRs, with regulation by hormones, various drugs, and lipid metabolites, respectively. Most circadian and other NRs are still considered orphans (brackets
in table), either without a widely‐accepted ligand, or having a described ligand but without clear gene regulatory effects. The various NRs can function as monomers, homodimers, and/or heterodi-
mers with retinoid X receptors (RXRs). NRs bind to DNA motifs called response elements, consisting of a 6‐nucleotide half site, or two such sites in the same (direct repeat, DR) or opposite
(inverted repeat, IR) directions spaced by 1–5 nucleotides. Some NRs have well‐studied and specific direct target genes in liver, and examples are shown. However, many liver NRs directly or
indirectly regulate many often‐overlapping sets of genes, such that no specific example is shown. See text for details
Class Receptor(s) Ligand(s) Dimerization Response element(s) Type Example canonical liver target gene
Endocrine THRs T3 (active thyroid hormone) RXR heterodimer DR‐4 2 THRSP (thyroid hormone responsive protein)
GR Glucocorticoids Homodimer IR‐3 1 G6PC (Glucose 6‐phosphatase)
AR Androgens Homodimer IR‐3 1
ERs Estrogens Homodimer IR‐3 1
VDR Calcitriol (active vitamin D) RXR heterodimer DR‐3 2 CYP24A1 (vitamin D 24‐hydroxylase)
Xenobiotic CAR Many, including phenobarbital RXR heterodimer DR‐3 or DR‐4 1,2 CYP2B1 (phase 1 drug metabolism)
PXR Many, including rifampin RXR heterodimer DR‐3 or DR‐4 1,2 CYP3A4 (phase 1 drug metabolism)
Metabolic LXRs Oxysterols RXR heterodimer DR‐4 2 ABCG5/ABCG8 (cholesterol efflux)
FXR Bile acids RXR heterodimer IR‐1, DR‐5 2 ABCB11 (Bile salt export pump)
PPARs Fatty acid derivatives RXR heterodimer DR‐1 2 FGF21 (liver‐derived hormone)
RARs All‐trans retinoic acid (ATRA) RXR heterodimer DR‐5, DR‐2, others 2 CRABP1/2 (cellular retinoic acid binding proteins)
RXRs 9‐cis retinoic acid Heterodimer, homodimer DR‐1 (homodimer) 2,3
LRH1 (Phospholipids) Monomer, SHP heterodimer Half site 4 CYP7A1 (bile acid synthesis)
SHP (Orphan) LRH1, others N/A N/A
Circadian Reverbs (Heme) Monomers or homodimers Half site or DR‐2 4 BMAL1 (core clock gene)
RORs (Orphan) Monomers Half site 4 BMAL1 (core clock gene)
ERRs (Orphan) Monomers or homodimers Half site 4
Other NR2A/HNF4s (Linoleic acid) Homodimers DR‐1 3
NR2C/TRs (Orphan) Homodimers Various DRs 3
NF2F/COUP‐TFs (Orphan) Homodimers Various DRs 3
NR4As (Orphan) RXR heterodimers, monomers DR‐5 or half site 2,4

c28.indd 340 03-11-2019 19:50:15


28:  Hepatic Nuclear Receptors 341

7α‐hydroxylase (CYP7A1). Hypothyroidism increases glucose‐6‐phosphatase (G6PC), which is required to liberate


serum cholesterol (particularly in low density lipoproteins) free glucose. Glucocorticoids also stimulate hepatic synthesis
and triglycerides, and also causes liver steatosis. Conversely, of glycogen, fatty acids, and cholesterol, particularly in the
thyroid hormone is an effective treatment for dyslipidemia fed state in the presence of insulin. Indeed, hepatic steatosis is
but limited by adverse cardiac effects (tachycardia) and observed with glucocorticoid treatment, and this effect may
decreased bone mass. Selective activation of liver THRs, be due to direct cell‐autonomous effects on hepatocyte GR as
either by TRβ specific agonists or selective hepatic delivery, well as indirect effects upon adipose tissue, where glucocorti-
is thus a promising therapeutic approach to dyslipidemia coids activate lipolysis releasing fatty acids for uptake in liver
and non‐alcoholic fatty liver disease (NAFLD) and non‐ [13]. Overall, glucocorticoid excess (Cushing’s syndrome,
alcoholic steatohepatitis (NASH). either iatrogenic or endogenous due to overproduction), is
In addition to the liver being a target of thyroid hormone, it associated with metabolic disease including obesity, insulin
alters systemic thyroid hormone levels by two main mecha- resistance, and hyperglycemia, some of which relate to GR
nisms: deiodination and synthesis of binding proteins. The effects in liver.
thyroid gland, in response to the hypothalamic–pituitary axis, In addition to gene activation, widespread gene repression is
releases the hormones T4 (thyroxine, with four iodines) and T3 also noted in response to glucocorticoids, particularly of inflam-
(3,5,3′‐triiodothyronine, with one less iodine in the 5′ posi- matory genes. As opposed to direct DNA binding by GR to acti-
tion of its outer ring). THRs bind predominantly to T3, such vate genes, this gene repression is thought to result from
that T4 is essentially a prohormone acted upon by deiodinase “transrepression,” whereby GR is brought to DNA indirectly by
enzymes in tissues to generate T3. Importantly, the liver tethering to inhibit activation by prominent inflammatory tran-
expresses type 1 deiodinase enzyme (DIO1), which can deio- scription factors like AP‐1 and NF‐κB. Therefore, an elusive
dinate the outer ring to generate either active T3 or the inactive goal in pharmacology has been developing glucocorticoid ana-
reverse T3 (rT3), as well as the inner ring to generate other logs that stimulate transrepression of inflammatory genes with-
inactive derivatives [10]. While DIO1 contributes to circulat- out the adverse metabolic effects of activating GR in liver,
ing levels of active T3, its main function in liver appears to be muscle, and fat [14].
thyroid hormone inactivation. In mouse models, DIO1 is a Cortisol is the main glucocorticoid in humans, produced by
target gene activated by thyroid hormone in liver, such that the zona fasciculata of the adrenal cortex in response to a hypo-
liver TRβ‐deficiency decreases DIO1, while thyrotoxicosis thalamic–pituitary axis. 11β‐hydroxysteroid dehydrogenase
increases hepatic DIO1. Therefore, the liver mediates nega- (HSD) enzymes interconvert cortisol and inactive cortisone,
tive feedback whereby excess thyroid hormone, via TRβ, with 11β‐HSD2 primarily inactivating cortisol (notably in renal
stimulates its own inactivation by DIO1. cells to protect the mineralocorticoid receptor from cortisol) and
The liver is also the source of all three circulating thyroid 11β‐HSD1 primarily reactivating cortisone. The liver is the
hormone binding proteins: thyroxine‐binding globulin (TBG), main site of 11β‐HSD1 expression to activate cortisone to corti-
transthyretin (TTR, also called prealbumin), and albumin [11]. sol, and this enzyme has been implicated strongly in metabolic
While free thyroid hormone is the active fraction taken up by diseases like NAFLD [13]. Liver also expresses enzymes that
cells, more than 99% of circulating thyroid hormone is protein‐ inactivate cortisol including 5α‐ and 5β‐reductases and 3α‐HSD,
bound, and the two forms are in equilibrium. Steroid NR and thyroid hormone increases hepatic cortisol clearance by
ligands are well‐known to alter TBG, the major binding pro- activating these pathways. Cortisol also circulates predomi-
tein, and thus affect the total versus free fractions of serum T4: nantly bound to a liver‐derived carrier protein, cortisol binding
androgens and glucocorticoids decrease TBG, while estrogens globulin (CBG). Therefore, similar to the situation for thyroid
increase TBG (though indirectly by affecting glycosylation). hormone, the liver is a direct target for glucocorticoid action on
Often these effects on binding proteins only affect total thyroid glucose and lipid metabolism via GR, and it also plays a central
hormone levels but not the active free levels, such that patients role in determining systemic glucocorticoid effects via cortisol
remain clinically euthyroid, though selective effects on thyroid metabolism and CBG production.
hormone activity are plausible. Liver diseases like cirrhosis or
acute hepatitis can thus affect thyroid hormone binding pro-
Mineralocorticoid receptor (MR)
teins or deiodination [8].
Expression of MR is detectable in liver, though it is uncertain
whether this is in hepatocytes, versus in endothelial cells or
Glucocorticoid receptor (GR) macrophages where MR expression has been studied [15].
GR is present in nearly every cell type, and glucocorticoids exert MR is closely related to GR, and in the absence of 11β‐HSD2
pleiotropic effects on the immune and cardiovascular systems, as to inactivate cortisol, MR can be similarly activated by corti-
well as metabolism. Pharmacologic glucocorticoids are primar- sol or the main mineralocorticoid aldosterone. Aldosterone is
ily used as anti‐inflammatory and immunosuppressive agents, produced by the zona glomerulosa of the adrenal cortex,
but metabolic side‐effects are common. The name glucocorticoid under control of the renin–angiotensin system, and it binds
stems from their effects on carbohydrate metabolism to maintain MR in kidney to activate genes like the epithelial sodium
blood glucose in the fasted state, most prominently via stimula- channel (ENaC) to regulate sodium and potassium homeostasis
tion of gluconeogenesis in hepatocytes [12]. Thus, direct GR and blood pressure. MR also has roles in the cardiovascular
target genes in liver include phosphoenolpyruvate kinase system and gut, but little is known about any function for
(PCK1), encoding the rate‐limiting gluconeogenic enzyme, and hepatic MR.
342 THE LIVER:  XENOBIOTIC SENSING RECEPTORS: CAR AND PXR

Estrogen and androgen sex steroid receptors role of sex hormones and their receptors in liver energy
(ER and AR) ­homeostasis is complex and an area of active investigation.

Sex hormone receptors are widely expressed and not limited to


reproductive tissues. Like GR and MR, they are type 1 NRs with Vitamin D receptor (VDR)
ligand‐induced nuclear localization. AR and ERα are expressed in Unlike the steroid receptors, VDR is an RXR heterodimer type
liver in both sexes, while ERβ and PR have little to no hepatic 2 NR like THR. VDR is activated by the hormone calcitriol, or
expression. Sex differences are observed in many liver diseases, 1,25‐dihydroxyvitamin D. The precursor vitamin D3, or chole-
with male preponderance in steatosis, hepatitis, cirrhosis, and calciferol, is obtained from the diet or synthesized in sun‐
hepatocellular carcinoma [16]. Of course, other factors besides sex exposed skin. 25‐hydroxylation of vitamin D3 occurs in liver,
steroids drive many differences, including behaviors related to via the P450 enzyme encoded by the CYP2R1 gene. However,
viral infection and alcohol use. Sexual dimorphism has also been CYP2R1 is not regulated by VDR or other inputs, so constitutive
found in the liver gene expression, with sex‐specific expression of activity in liver results in inactive 25‐hydroxyvitamin D as the
more than 1000 genes including some encoding drug‐metaboliz- primary circulating form reflecting overall vitamin D levels.
ing enzymes. Some of these differences had been attributed to The final activation to calcitriol occurs via 1α‐hydroxylation in
direct cell‐autonomous effects of ERα and AR in liver, but more the kidney, a step under tight regulation by parathyroid hormone
recent studies suggest brain control of this dimorphism via sex‐ and phosphate levels. Conversely, calcitriol is inactivated by 24‐
specific temporal patterns of growth hormone release [17]. hydroxylation, and the CYP24A1 gene encoding this 24‐hydroxy-
Estrogen is the primary female sex hormone, synthesized in lase is highly regulated by VDR. Thus, calcitriol exerts negative
the ovary by the combined action of theca and granulosa cells, feedback on itself by VDR activation of CYP24A1.
the latter expressing aromatase enzyme. In males, much lower VDR is expressed in most tissues, though overall hepatic
amounts of estrogen are generated by the testes and peripheral expression is extremely low compared to its levels in parathy-
aromatase‐expressing tissues including fat. Clinical observa- roid, kidney, and GI tract [6]. Indeed, hepatocytes have little
tions show that modifying estrogen levels, either increasing (i.e. to  no detectable VDR, but levels are much higher in non‐
hormone replacement in post‐menopausal women) or decreas- parenchymal cells of liver particularly hepatic stellate cells
ing (suppression therapy in breast cancer), affects serum lipid (HSCs), where VDR regulates CYP24A1 and other genes [25].
profiles, generally with estrogen lowering LDL cholesterol and Importantly, HSCs mediate liver fibrosis in response to injury
raising HDL cholesterol and triglycerides [18]. Furthermore, and pathology, and an antifibrotic role of VDR in HSCs is
mouse models with estrogen deficiency show hepatic steatosis emerging [26]. VDR activation inhibits the main profibrotic
that is rescued by estrogen treatment, yet such studies are unable pathway in HSCs mediated by TGFβ, and VDR knockout mice
to distinguish liver autonomous versus whole body effects of show spontaneous liver fibrosis. Thus, there is potential for
estrogen. More recent liver‐specific knockout mice have shown VDR‐targeted therapies in fibrotic liver diseases, especially if
a role of hepatic ERα in glucose and lipid metabolism, with liver‐specific ligands are developed to avoid hypercalcemia due
hyperglycemia and liver steatosis [19]. Liver ERα has also been to excess VDR activity in bone, gut, and kidney.
linked to cholesterol and lipoprotein metabolism, via a func-
tional interaction with LXRα (see Liver X and farnesoid X
receptors (LXR and FXR). Indeed, estrogen can ameliorate the
fatty liver induced by LXR agonists, in a manner requiring XENOBIOTIC SENSING RECEPTORS:
hepatic ERα [20]. Furthermore, ERs can even bind oxysterols CAR AND PXR
that are classically LXR agonists [21].
Androgens are synthesized in the testes and the zona reticula- CAR and PXR are VDR‐related NRs that sense xenobiotics,
ris of the adrenal cortex, with the latter being the main source of foreign substances like drugs and toxins that are not naturally
low levels in females. Androgen excess in females is part of the produced by the organism but require clearance. The names of
polycystic ovarian syndrome (PCOS), with metabolic manifesta- both NRs reflect historical identification of ligands, though
tions including obesity, insulin resistance, and NAFLD, though it numerous chemical compounds have now been shown to serve
is unlikely that hepatic sex steroid receptors play a significant with variable affinities as agonists, antagonists, partial agonists,
role in PCOS pathophysiology. The effect of androgens on liver and inverse agonists [27]. The structures of PXR and CAR
steatosis is complex, with inconsistent and opposite effects in reveal large ligand binding pockets that can accommodate such
various studies [16]. Liver‐specific AR knockout causes hepatic large and diverse ligands, with many being dual activators of
steatosis in male but not female mice [22], though the mecha- both CAR and PXR. In research, the model agonist ligands for
nism is unclear. Similar to ER, AR has been proposed to regulate CAR and PXR are the anticonvulsant phenobarbital and the
cholesterol metabolism by cross‐talk with LXR [23]. antibiotic rifampicin, respectively. Both NRs activate genes
Like for thyroid and glucocorticoid hormones, the liver is also involved in various stages of the detoxification response: phase
a site for metabolism and clearance of sex steroids, and the syn- 1 modification (cytochrome P450 enzymes), phase 2 conjuga-
thesis of sex hormone binding globulin (SHBG). Hepatic SHBG tion (glutathione S‐transferase [GST], UDP‐glucuronosyl-
expression is increased by estrogens and decreased by andro- transferases [UGTs], and sulfotransferases [SULTs]), and
gens. Notably, in contrast to TBG and CBG, SHBG levels do phase 3 transport/efflux (organic anion‐transporting poly-
appear to alter levels of free sex hormones in blood, and SHBG peptides [OATPs] and multidrug resistance proteins (MDRs]).
is an emerging biomarker in metabolic disease [24]. Overall, the Classic targets for CAR are CYP2B genes, and for PXR CYP3A4,
28:  Hepatic Nuclear Receptors 343

which encodes the most abundant human liver P450 responsible with “reverse cholesterol transport” to liver, where cholesterol is
for metabolizing over a third of clinically used drugs [28]. taken up then secreted directly into bile or metabolized to bile
CAR is expressed almost exclusively in liver, while PXR is acids. Pharmacologic LXR agonists thus have anti‐atheroscle-
also highly expressed in the gastrointestinal tract [6]. The litera- rotic effects in multiple studies, and LXRs have also been pro-
ture on CAR mechanism is complex, suggesting features of posed to have insulin‐sensitizing and anti‐inflammatory roles
both type 1 and type 2 NRs. Like type 1 NRs, CAR can be [33]. LXRs also have a prominent role in stimulating hepatic de
retained in the cytosol by HSPs and the cytoplasmic CAR reten- novo lipogenesis, such that liver steatosis and hypertriglyceri-
tion protein (CCRP) [29], and it is reported not only to homodi- demia have limited the therapeutic application of LXR agonists
merize but also to form an RXR heterodimer like the closely [34]. This effect is mediated indirectly as LXR/RXR is
related type 2 NR VDR. However, unlike other RXR heterodi- potent activator of SREBP‐1c, the master regulator of de novo
mers that repress transcription in the absence of ligand, CAR/ lipogenesis [35].
RXR is reported to have constitutive ligand‐independent activa- FXR expression is limited to liver, gut, kidney, and steroido-
tion [30]. PXR is also an RXR heterodimer with apparent regu- genic tissues [6]. Only FXRα exists in humans, as FXRβ is a
lation by nuclear localization [31]. pseudogene in primates (and rodents thus have 49 NR genes
In addition to being master regulators of the xenobiotic compared to 48 in humans). Various bile acids serve as FXR
response and drug metabolism, it has become clear that CAR ligands, with the strongest natural agonist being the hydropho-
and PXR also have roles binding endobiotics and functioning in bic chenodeoxycholic acid (CDCA). Less potent agonists
normal physiology [32]. Reported endogenous ligands include include deoxycholic acid and lithocholic acid, while hydrophilic
metabolites of bile acids, steroids, bilirubin, and lipids. CAR bile acids like ursodeoxycholic acid and murocholic acids do
may indirectly affect metabolic rate, as phase 2 SULT and UGT not activate FXR. Other sterol derivatives are also reported as
enzymes are involved in the inactivation of thyroid hormones. weak FXR agonists of uncertain relevance. Canonical FXR tar-
By this mechanism, CAR antagonism might be desirable, yet get genes in liver include the atypical NR SHP (discussed below
other studies suggest instead that CAR activation improves met- in the section on LRH1 and SHP) and the canalicular bile salt
abolic health. Furthermore, there are conflicting results about export pump (ABCB11). In the distal small intestine, FXR tar-
whether CAR activates or represses key pathways like gluco- gets include the ileal bile acid binding protein (I‐BABP) and,
neogenesis and de novo lipogenesis. Both CAR and PXR have importantly, the hormone FGF19 (FGF15 in rodents), which
been proposed to reduce the activation of glucose and fat syn- provides negative feedback to hepatic bile acid synthesis. FXR
thesis by other transcription factors like CREB, via an indirect functions beyond bile acid metabolism have also been described
mechanism akin to transrepression. Overall, modulation of [36]. In both intestine and liver, anti‐inflammatory effects of
CAR and PXR (in clinical studies with ligands, in knockout FXR have been reported, making FXR a candidate target in
mice, in cell models, and so on) has been consistently linked to inflammatory bowel disease and hepatitis. In the gut, FXR and
disturbances in energy homeostasis, but results have been mixed bile acids have complex crosstalk with the intestinal microbi-
and conflicted such that the precise metabolic role of xenobiotic ome, and in liver FXR has effects on fatty acid and glucose
receptors remains quite uncertain, particularly in humans. metabolism. In contrast to the activation of SREBP1c and lipo-
genesis by LXR, FXR has the opposite via SHP‐mediated
repression. Indeed, FXR has been proposed to integrate liver
energy balance in the fed state, with PPARα as its counterpart in
METABOLIC NUCLEAR RECEPTORS the fasting state [37]. Bile acids and their derivatives are cur-
rently strong candidates for therapeutics, via activation of FXR
Liver X and farnesoid X receptors (LXR and/or the cell surface G protein coupled bile acid receptor
(GPBAR1, or TGR5). In particular, the FXR agonist obeticholic
and FXR): sensing sterols and bile acids
acid is approved for use in primary biliary cholangitis, and it is
The important effects of LXRs and FXR on hepatic lipid and in phase 3 trials for NASH [38].
bile acid metabolism are described in detail in other chapters
and will be summarized here. LXRs and FXR are RXR heter-
odimer type 2 NRs, and both pathways are thought to be active Peroxisome proliferator activated receptors
in the fed state in liver [2].
There are two isoforms of LXR, with LXRβ considered ubiq-
(PPARs): sensing lipids
uitous and LXRα highest in liver, fat, gut, and immune cells, yet The three PPAR nuclear receptors sense cellular lipids and regu-
the mRNAs for both are present in most tissues including liver late key transcriptional programs in lipid metabolism. The name
[6]. LXR agonists are oxysterols, oxygenated derivatives of PPAR relates to the effects of PPARα agonists on rodent liver,
cholesterol including several specifically generated by cellular with induction of peroxisome proliferation, hepatomegaly, and
hydroxylase enzymes (i.e. 24S‐, 25‐, and 27‐hydroxycholes- hepatocellular carcinoma. Notably, none of these adverse effects
terol), as well as other potent ligands of uncertain physiological occur with PPARα agonism in human liver, nor with ligands for
significance (i.e. 24S,25‐epoxycholesterol 22R‐ and 20S‐ the other two PPARs. All three PPARs are type 2 NRs that func-
hydroxylcholesterol). LXRs coordinately activate a program of tion as RXR heterodimers to bind DR‐1 response elements. The
gene expression that removes cholesterol from peripheral cells PPARs are expressed across multiple tissues, though PPARα is
(such as macrophages in atherosclerotic lesions) via ABC1A1‐ highest in liver and PPARγ in adipose tissue [6]. The precise
mediated efflux to nascent HDL (APOA1 secreted by liver), endogenous ligands for PPARs remain uncertain, but there is
344 THE LIVER:  METABOLIC NUCLEAR RECEPTORS

general agreement that they are various fatty acids and their agonist saroglitazar is used in India to treat diabetic dyslipi-
metabolites, such as arachidonic acids and other polyunsatu- demia [47], while there is great interest in the dual PPARα/δ
rated fatty acids, or eicosanoids like 15‐hydroxyicosatetraenoic agonist elafibrinor in late stage development for NASH [48].
acids (15‐HETE) and prostaglandins [39]. The discussion above focused on PPARs in hepatocytes, but
As the predominant PPAR in liver, PPARα functions as a they may also function in other liver cell types like HSCs related
master regulator in hepatic lipid metabolism [40]. In the fasting to liver fibrosis [49].
state, liver takes up free fatty acids released by lipolysis in adi-
pose tissue, and liver PPARα activates genes required for hepatic
fatty acid oxidation (FAO) and ketogenesis. FAO is also indi-
Retinoic acid receptors (RARs): sensing
rectly required for gluconeogenesis in fasting, as the source of retinoic acid
cellular ATP to fuel glucose production and of mitochondrial RARs are activated by metabolites of vitamin A, the most potent
acetyl‐CoA, which activates pyruvate carboxylase (the first step being all‐trans retinoic acid (ATRA). Vitamin A is a fat‐soluble
in gluconeogenesis). Therefore, mice deficient in PPARα appear vitamin with diverse roles in development and physiology, and
normal but respond poorly to fasting, with elevated serum free ~90% of total body stores are retinyl esters in lipid droplets of
fatty acids and severe hepatic steatosis, but defective ketogene- HSCs. Metabolism of retinoids is complex and covered in detail
sis and gluconeogenesis. PPAR target genes are also involved in in other chapters. In liver, many classic target genes of RARs
lipoprotein metabolism, such that PPARα fibrate drugs are used function in retinoid metabolism, including binding proteins and
to lower serum triglycerides. Another key transcriptional target modifying enzymes.
of PPARα is FGF21, a liver‐derived hormone that increases in RARα is the most abundant RAR isoform in whole liver [6],
the fasted state and exerts effects throughout the body [41]. though the RARβ and RARγ are present and could be enriched
PPARγ has two protein isoforms generated by alternative in HSCs. Connections have been noted between vitamin A and
transcriptional initiation and splicing of the same gene: widely‐ bile acid metabolism, with complex crosstalk among RAR,
expressed PPARγ1 and adipose‐selective PPARγ2, having an FXR, RXR and their ligands, and bile acids required for absorp-
additional 30 amino acids at the N‐terminus. Adipocytes express tion of fat soluble vitamins [50]. Liver diseases with cholestasis
both PPARγ1 and PPARγ2, and PPARγ is a master regulator of are associated with vitamin A deficiency and disrupted retinoid
fat cell differentiation. PPARγ agonist thiazolidinediones metabolism, and ATRA combined with bile acid therapy has
(TZDs) are used clinically in type 2 diabetes to lower blood glu- shown some promise in treating cholestasis in animal models
cose, and they function primarily in adipose tissue as insulin‐ and a human pilot study [51].
sensitizers [39]. PPARγ1 is expressed in macrophages, thus During liver injury and fibrosis, HSCs transdifferentiate into
widespread low levels of PPARγ in some tissues may reflect myofibroblasts with features of fibroblasts and smooth muscle
resident macrophages. While mostly PPARγ1 is present in nor- cells, which lose vitamin A storage capacity such that vitamin A
mal liver, multiple rodent models with liver steatosis show metabolism is disturbed in NAFLD [52]. Furthermore, the top
selective induction of the PPARγ2 isoform, consistent with a genetic variant associated with human liver steatosis is
model whereby hepatic PPARγ is lipogenic in liver like it is in PNPLA3‐I148M, and while the exact function of PNPLA3 is
adipocytes [42]. However, while liver autonomous effects of unknown it has retinyl ester hydrolase activity (in addition to
PPARγ may increase triglyceride storage, PPARγ agonist treat- triglyceride lipase activity) [53]. Therefore, vitamin A metabo-
ment of intact animals generally reduces liver steatosis. This lites also have potential roles in metabolic liver diseases, and
may reflect the predominant effect of TZDs increasing lipid RAR agonists have shown favorable effects in mouse models of
storage in adipose tissue, thus resulting in “lipid steal” from NAFLD [54].
liver and other tissues [43]. Notably, multiple human trials have
shown that the PPARγ agonist pioglitazone improves liver his-
tology in NAFLD, and current guidelines support consideration Retinoid X receptors (RXRs) and nuclear
of pioglitazone in patients with biopsy‐proven NASH with or
receptor heterodimerization
without diabetes [44].
PPARδ (also called PPARβ) is widely‐expressed but best Of the three RXR isoforms, RXRα expression predominates in
studied in highly oxidative tissues like skeletal and cardiac mus- liver, with lower levels of RXRβ and very low or undetectable
cle. Studies of hepatic PPARδ are more limited and often con- RXRγ [6]. The best‐studied RXR ligand is 9‐cis retinoic acid,
flicting, but they certainly suggest a role in regulating glucose though endogenous levels of this and other potential RXR
and lipid metabolism, overlapping with PPARα and PPARγ ligands in tissues (including liver) are a point of contention. All
[42]. However, it is clear from the phenotype of PPARα‐deficient three RXR isoforms can form homodimers (or heterodimers
mice in fasting that neither PPARδ or PPARγ can rescue the key with each other) to bind DR‐1 motifs, like the other type 3 NRs
role of PPARα in hepatic lipid metabolism. in the NR2 family. However, the physiological roles of these
There is great interest in drugs targeting PPARs for their met- RXR homodimers, as well as the redundancy versus specific
abolic and anti‐inflammatory effects. The clinical use of fibrate function for each isoform, remain largely unknown.
PPARα and TZD PPARγ agonists is described above, and RXRs are best known as the obligate heterodimeric partners
PPARδ agonists have been proposed as exercise mimetics due to with the endocrine, metabolic, and xenobiotic NRs described
their effects on muscle metabolism [45]. There are dual PPAR above. The potential benefits of heterodimerization are subject
agonists and pan‐PPAR agonists, as well as efforts to develop to much teleological speculation, and RXR heterodimerization
selective and tissue‐specific drugs [46]. The dual PPARα/γ has been described as the “big bang” in the evolution of the NR
28:  Hepatic Nuclear Receptors 345

superfamily, as well as in the study of NRs [2]. Most RXR het- CIRCADIAN NUCLEAR RECEPTORS
erodimers are highly dependent on the ligand for the RXR part- IN LIVER
ner, while RXR ligands have variable effects. RXR heterodimers
are classified as “permissive” if RXR ligands generally activate,
or “non‐permissive” if RXR ligands appear silent [2]. Notably,
Rev‐erbs: heme‐binding circadian nuclear
the non‐permissive group includes endocrine hormone recep- receptors
tors like TR and VDR which respond mainly to their cognate The two Rev‐erb NRs, of which Rev‐erbα is better studied, were
hormones, while the permissive group includes lipid‐sensing originally orphan receptors but have since been shown to bind
NRs like LXRs, FXR, and PPARs. For these permissive heter- heme and regulate circadian rhythms [62]. Twenty‐four‐hour
odimers, synergistic activation has been observed with the com- diurnal rhythms are fundamental in biology, synchronizing
bination of both ligands compared to either alone, and behavior and physiology with the day–night cycle. The elegant
pharmacological pan‐RXR agonists called rexinoids have been molecular underpinnings of this clock have been worked out in
developed to activate this entire group [55]. In general, ligands great detail [63]. The CLOCK/BMAL1 transcription factor is
of a single NR can have unintended and pleiotropic effects due the key activator of circadian genes, also activating the period
to diverse and tissue‐selective gene activation, and this problem and cryptochrome (PER and CRY) factors which accumulate
could be compounded by activating all RXR heterodimers. and exert negative feedback on CLOCK/BMAL1. PER and
CRY (the “negative arm”) then decrease allowing re‐accumula-
Liver receptor homolog 1 (LRH1) and small tion of CLOCK/BMAL1 (the “positive arm”) in a 24‐hour
heterodimer partner (SHP) cycle, and this is the core circadian clock. Rev‐erbs are also acti-
vated by CLOCK/BMAL and thus have drastic diurnal rhythms
LRH1 is widely expressed with highest levels in liver, pancreas, in their mRNA and protein levels, and this is not just an output
and intestine [6]. It was originally an orphan, but crystal struc- of the clock but a key input to the feedback system [64]. The
tures revealed a large ligand binding pocket occupied by a phos- canonical target gene of Rev‐erbs is BMAL1 itself, and this tran-
pholipid. Mutagenesis and other studies support a model of scriptional repression serves as a second negative feedback
ligand‐independent transcriptional activation that can be (“stabilizing loop”) on CLOCK/BMAL1 to maintain its 24‐hour
enhanced by phospholipid binding [56]. It is debated whether rhythm (Figure 28.3a). The central clock in the hypothalamus is
phospholipid ligands simply serve a constitutive structural role, entrained by light, and mice lacking whole body Rev‐erbα and
or actively regulate LRH1 in normal physiology. Consistent Rev‐erbβ have profound disruptions in circadian behavior [65].
with the latter model of LRH1 as a phospholipid sensor, LRH1 There are also autonomous peripheral clocks operating in most
was recently shown to have a key role in hepatic lipid storage tissues and cells, and Rev‐erbs are well‐studied as elements of
and phospholipid composition [57]. the hepatic clock. Rev‐erbs regulate many key genes in liver
A key aspect of LRH1 function in liver is its binding to metabolism, and liver‐specific loss of both Rev‐erbs results in
SHP, an orphan NR with a similar tissue expression pattern. marked hepatic steatosis [66].
SHP belongs to the distinctive NR0B group of NRs lacking Uniquely among NRs, Rev‐erbs lack the C‐terminal α‐helix
DNA binding domains, and is thus brought to DNA indirectly 12, which is required for coactivator binding, and thus serve as
by heterodimerization with other NRs (or even binding to transcriptional repressors by recruiting the NCoR/HDAC3 core-
non‐NR transcription factors), where it has a repressive effect pressor complex. As opposed to type 2 NRs in which ligand
on transcription [58]. SHP expression is highly regulated, and stimulates corepressor release, heme binding to Rev‐erbα stabi-
activated by many NRs (particularly FXR) to mediate poten- lizes the interaction with NCoR to increase its repressive activ-
tial negative feedback loops. LRH1 is one of the best‐studied ity. It remains uncertain whether heme physiologically
binding partners for SHP, such that the two function together modulates Rev‐erbs to control heme metabolism, or whether
to form a non‐traditional heterodimer with LRH1 providing heme binding is constitutive and structural (similar to the situa-
DNA binding and SHP providing repressive activity. A simi- tion with phospholipids and LRH1 described in the section on
lar situation occurs in steroidogenic tissues (adrenal cortex, LRH1 and SHP). Rev‐erbs are type 4 NRs that bind as homodi-
testis, ovaries) with the LRH1‐related SF1 repressed by the mers or monomers to DNA, though only dimeric Rev‐erb can
SHP‐related DAX1. recruit NCoR to actively repress genes. However, monomeric
LRH1 regulates expression of bile acid biosynthetic Rev‐erb can repress genes by competing for the same half site
enzymes CYP7A1 and CYP8B1. It was originally proposed DNA elements with RAR‐related orphan receptors (RORs),
that LRH1/SHP repression of CYP7A1, with SHP induced by which are activators as described below. Repression of circadian
bile acids via FXR, was responsible for the negative feedback genes like BMAL1 in all tissues appears to involve direct Rev‐
of bile acids on their own synthesis [59]. Subsequent studies erb binding to DNA and displacement of ROR (Figure 28.3b),
showed that neither LRH1 or SHP was required for this nega- while repression of metabolic genes in liver instead often
tive feedback, indicating other possibly redundant mecha- involves indirect binding via tethering to other factors [67].
nisms, but loss of either NR in liver markedly alters bile acid Circadian regulation of metabolism may serve to synchronize
homeostasis [60, 61]. In addition to bile acid homeostasis, expression of key gene products with anticipated daily periods
hepatic LRH1 and SHP have also been proposed to regulate of rest and sleep versus activity and eating. This is a subject of
many other genes involved in cholesterol, lipid, and glucose great interest, as circadian disruptions such as shift work are
metabolism (similar to LXRs, PPARs, and other NRs associated with metabolic diseases, and alterations in circadian
described above) [56, 58]. gene expression occur in steatotic livers [68]. In liver and other
346 THE LIVER:  OTHER HEPATIC NUCLEAR RECEPTORS

(a) (b)
Positive Arm
ROR Rev-erb ROR
CORE BMAL1

PER CLOCK BMAL1

CRY CLOCK Rev-erb

Response
Stabilizing Element
Negative Arm
Loop

Figure 28.3  Nuclear receptors in the circadian clock. (a) There is a well‐described transcriptional–translational loop that maintains 24‐hour diur-
nal rhythms. The core clock has a positive arm mediated by BMAL1/CLOCK and a negative arm mediated by PER/CRY. This is stabilized via regu-
lation of BMAL1 by two opposing nuclear receptors groups, the Rev‐erbs and RORs. Rev‐erbs are activated by BMAL1, but then act as transcription
repressors of BMAL1 in a negative feedback loop. (b) At the BMAL1 promoter response element, the Rev‐erb repressors compete for binding with
the ROR activators. Figure adapted from [69].

tissues, Rev‐erbs provide a key link between circadian and met- NR, and it is abundant in human liver, while expression of ERRβ
abolic pathways, such that therapeutics targeting Rev‐erb are and ERRγ is low or undetectable [6]. However, ERRγ is more
candidates to affect both processes. abundant in mouse liver, where it has been implicated in meta-
bolic pathways like gluconeogenesis [72].
ERRs have emerged as key regulators of fatty acid oxida-
RAR‐related orphan receptors (RORs)
tion and mitochondrial respiration, and many of these func-
RORs are closely related to Rev‐erbs, also regulating circadian tions are due to recruitment of PCG1 coactivators. Indeed,
and metabolic genes in liver. RORs bind to DNA as monomers PCG1α is well‐known to stimulate oxidative phosphoryla-
only, occupying half sites rather than repeats, and are thought to tion and mitochondrial biogenesis, and this activity requires
function solely as transcriptional activators even in the absence ERRα [73]. Therefore, ERR activation is a potential avenue
of ligand. In fact, ligand binding may even decrease the basal to increase energy expenditure. However, this is best studied
activation activity (inverse agonism) [69]. The identities of nat- in highly oxidative tissues like skeletal and cardiac muscle,
ural ROR ligands are controversial, with reports of binding to yet in adipocytes ERRα can also bind the corepressor
various metabolites including oxysterols (like LXRs), vitamin RIP‐140 to stimulate lipid storage [74]. ERRα‐deficient
D metabolites (like VDR), and retinoids (like RARs). RORβ is mice showed decreased body weight and resistance to high
restricted to the central nervous system, while both RORα and fat diet‐induced obesity, but these mice lack the NR in all
RORγ are more widely expressed including high expression in tissues [75]. The specific role of ERRs in liver is uncertain,
liver. Prominent roles of these RORs have also been found in and while ERR activation in other tissues may be desirable,
immune cells related to autoimmune disease, particularly the it appears from mouse models that hepatic ERR inhibition
RORγt splice variant in the TH17 class of CD4 T cells [69]. improves glucose metabolism [76]. ERRs clearly play
In liver, both RORα and RORγ have diurnal rhythms in their important roles in mitochondrial energy metabolism, but
expression levels. As part of the stabilizing circadian loop their roles as circadian factors in liver are only beginning to
described above, the transcriptional activator RORs compete be deciphered.
with repressor Rev‐erbs for binding upstream of BMAL1, thus
having the opposing effects on the expression of this core clock
gene (Figure 28.3b) [70]. Similar competition occurs at other
genes, such that RORs also regulate metabolic genes in liver
OTHER HEPATIC NUCLEAR RECEPTORS
[71], and drugs targeting RORs are likewise being developed
but remain far from clinical application. Hepatocyte nuclear factor 4 (HNF4)
HNF4s are NR2As expressed mainly in liver, GI tract, kidney,
Estrogen related receptors (ERRs) and pancreatic beta cells. Two HNF4 isoforms are encoded by
different genes, with HNF4γ predominating in small intestine
and mitochondria
such that liver studies focus on HNF4α. The name HNF reflects
In mouse liver, marked circadian rhythms in expression have the discovery of a diverse and unrelated group of transcription
been shown for all three ERRs [64]. In contrast to Rev‐erbs and factors as liver proteins with DNA binding activity, but HNF4s
RORs, this appears not to be an input to the core clock, but are the only NRs among these (HNF1s are POU‐homeodomain
rather an output such that ERR targets also have diurnal rhythms. proteins, HNF3s are winged helix forkhead box proteins, and
ERRs are named based on homology to estrogen receptors, but HNF6s are ONECUT proteins). HNF4s bind DNA as homodi-
they do not bind estrogen or other steroids and remain orphan mers to DR‐1 response elements [77]. HNF4s are considered
receptors. ERRs are type 4 NRs that can bind DNA as mono- orphan nuclear receptors, though reversible binding of linoleic
mers or homodimers, and may even form heterodimers with two acid has been reported without apparent effects on transcrip-
different ERR isoforms. ERRα was the first described orphan tional activity [78].
28:  Hepatic Nuclear Receptors 347

In adult liver, HNF4α is thought to maintain hepatocyte dif- The COUP name is historical as these proteins were purified
ferentiated gene expression patterns. Mice completely lacking based on binding to a chicken regulatory DNA. NR2F family
HNF4α show embryonic lethality, likely due to the requirement members form homodimers, but heterodimers are also reported
for HNF4α in liver development [79]. Liver‐specific loss of with each other and with RXR [93]. These NRs bind to direct
HNF4α in adult livers causes hepatic steatosis, with increase repeats of various spacing, and they function primarily as tran-
serum bile acids but reduced cholesterol and triglycerides, all scriptional repressors. They are considered orphan receptors,
attributable to reduced hepatic expression of key genes in lipid though their ligand binding domains are similar to RARs and
and bile acid metabolism [80]. Consistent with this is a small RXRs, and retinoic acid is a reported ligand COUP‐TFII [94].
study in people with heterozygous mutations in HNF4α, which COUP‐TFII is the best‐studied NR2F and strongly implicated in
causes maturity onset diabetes of the young type 1 (MODY1) many developmental processes. In contrast, COUP‐TFI and EAR2
due to effects in beta cells. While MODY1 does not have a clear are primarily studied in the nervous and immune systems respec-
liver phenotype, subjects with a MODY1 mutation had reduced tively, and little to nothing is reported about their roles in liver.
serum levels of several liver‐derived apolipoproteins, consistent While COUP‐TFII deletion in adult mice does not give a discern-
with HNF4 maintaining expression of hepatocyte genes [81]. able phenotype, roles in energy metabolism are nonetheless emerg-
Even though HNF4 is mainly thought to maintain ligand‐inde- ing [95]. For instance, hepatic COUP‐TFII was recently shown to
pendent constitutive expression of these genes, it could cooper- be up‐regulated by fasting and play a role in gluconeogenesis and
ate with other cofactors or NRs to mediate regulated expression fatty acid oxidation [96]. Consistent with this, COUP‐TFII has
[82]. One example is the recent demonstration that HNF4α and complex physical and functional interactions with HNF4α, PPARα,
another transcription factor (PROX1) co‐recruit the HDAC3 and particularly GR to regulate metabolic genes [95].
corepressor complex to lipid metabolic genes [83], and another
is the drug‐metabolizing CYP3A4 gene which requires HNF4α
The NR4A subfamily
for its activation by xenobiotic NRs [84]. Other examples of
HNF4 crosstalk have been described for genes involved in bile The final group of hepatic NRs is the three closely related
acid, sterol, fatty acid, and glucose metabolism, raising the pos- NR4As, also called nerve growth factor IB (NGFIB, NUR77 or
sibility that HNF4 may yet emerge as a drug target [85]. NR4A1), nuclear receptor related 1 (NURR1, or NR4A2), and
neuron‐derived orphan receptor 1 (NOR1, or NR4A3). While
The testicular receptor (TR) NR2C subfamily two of these historical names relate to neurons, the three NR4As
are widely expressed including in liver [6]. All have been
Related to HNF4s are two NR2C family members also known reported to bind DNA as monomers, and NR4A1 and NR4A2 as
as testicular receptors: TR2 is NR2C1 and TR4 (also called RXR heterodimers [97]. No ligands are known or expected to be
TAK1) is NR2C2. Despite the TR name, both are widely found, as they are thought to be ligand‐independent NRs based
expressed including in the liver [6]. Mice deficient in NR2C1 on the NR4A2 crystal structure showing a constitutively active
have normal testis and spermatogenesis, mice deficient in conformation with hydrophobic side‐chains filling the ligand‐
NR2C2 have reduced sperm and other systemic problems, and binding pocket. Rather than ligands, NR4A family members are
mice lacking both factors die in early embryonic development regulated by their expression levels. They are immediate early
[86]. Both are considered orphan receptors, though NR2C2 has response genes rapidly induced by environmental changes, and
been reported to bind retinoids or fatty acids [87], consistent implicated in diverse cellular processes [98].
with homology to RXRs and HNF4s. The two NR2Cs may In liver, all three NR4As are induced by cAMP, which is
homodimerize or heterodimerize with one another and repress downstream of glucagon and catecholamine signaling in fasting
transcription. Genome‐wide binding profiles of NR2C2 in four [99]. These hormones trigger hepatic gluconeogenesis, and
human cell lines, including HepG2 liver carcinoma cells, sug- NR4A members were strongly implicated in this response.
gested roles in fundamental cellular processes like RNA metab- Overexpression of NR4A1 in mouse liver could mimic fasting
olism and protein translation, rather than specific metabolic and activate gluconeogenic genes, while a dominant negative
pathways [88]. However, in mouse liver, NR2C2 has been NR4A1 (that antagonizes all three NR4As) could inhibit gluco-
implicated in regulation of lipid metabolism particularly neogenic genes and lower glucose in a diabetic model. In addi-
stearoyl‐CoA desaturase 1 (SCD1) [89]. NR2C2 has also been tion to this role in gluconeogenesis, NR4A1 has also been
shown to be cAMP‐induced during fasting with a potential role implicated in hepatic lipid metabolism, so NR4A1 and it family
in hepatic gluconeogenesis [90]. Finally, NR2C2‐deficient mice members thus could emerge as therapeutic targets [100].
are protected against high fat diet‐induced metabolic disease
[91]. Similar to HNF4s, crosstalk between NR2Cs and other
NRs has also been proposed, particularly those which also bind
DR‐1 response elements like PPARs [92]. OTHER ISSUES IN HEPATIC NUCLEAR
RECEPTOR BIOLOGY
NR2F chicken ovalbumin upstream promoter‐
While there are 48 NR genes, the NR proteome is even more
transcription factors (COUP‐TFs) diverse. Most NR genes generate multiple mRNAs due to alter-
Also in the NR2 family related to HNF4s, TRs, and RXRs are the native transcriptional initiation and splicing (as described above
three NR2Fs or COUP‐TF family (COUP‐TFI, COUP‐TFII, and for PPARγ), and the proteins are subject to various reported
EAR2), all of which are widely expressed including in liver [6]. post‐translational modifications (phosphorylation, acetylation,
348 THE LIVER:  REFERENCES

SUMOylation, etc.) [101]. The relevance of these in liver is gen- FXR, LRH1, and SHP could treat cholestatic diseases. Many
erally poorly understood and beyond the scope here. NRs exert anti‐inflammatory effects that could be relevant in
A recent and profound advance in the study of NRs is the hepatitis. However, most hepatic NRs have clear roles in energy
advent of next generation sequencing approaches to identify their metabolism, particularly lipid homeostasis. It is notable that,
DNA binding sites genome‐wide (cistromes), such as chromatin while there are no approved therapeutics for NAFLD, one rec-
immunoprecipitation followed by deep sequencing (ChIP‐seq). ommended for off‐label use in NASH is the PPARγ agonist
Whereas previous efforts had identified only dozens of NR bind- pioglitazone [44], and two agents most advanced in phase 3
ing sites, mainly at response elements in promoters of well‐stud- ­trials specifically for NASH are other metabolic NR ligands:
ied targets, ChIP‐seq typically identifies tens of thousands of obeticholic acid targeting FXR [38], and elafibrinor targeting
binding sites throughout the genome and near most expressed PPARα/δ [48].
genes. As described above, both direct binding and indirect teth- Human genetics has also implicated various NRs in disease
ering is observed. Other unbiased methods like gene expression [104]. Rare PPARγ mutations cause familial partial lipodystro-
profiling (transcriptomes, for instance in response to NR ligands, phy with severe liver steatosis, while FXR mutations cause
knockdown, or overexpression) have also drastically increased ­progressive familial intrahepatic cholestasis. However, coding
the number of potential targets genes, but not to the same extent. mutations in NRs causing loss‐of‐function are expected to be
Comparisons of cistromes and transcriptomes make it clear that rare given the requirement of many NRs for normal develop-
not all binding events have apparent functional consequences on ment and homeostasis. More common may be non‐coding
gene expression. Furthermore, regulated transcripts represent a genetic variants in NR binding sites that affect regulation of spe-
mixture of direct and indirect targets, regulated in both directions cific target genes, rather than global NR activity. Such noncod-
(such that genes can be indirectly downregulated by ligands for ing regulatory variants would be predicted to have subtle effects
activating NR, and upregulated by a repressive NR). Advances in on complex disease risk rather than causing Mendelian syn-
genomic and computational methods have thus opened new win- dromes. Indeed, in human genome‐wide association studies
dows into the complexity of NR function [2]. (GWAS), over 95% of causal genetic variants that are associated
Most regulatory elements occupied by NRs do not simply con- with phenotype fall in noncoding regions, and these are thought
sist of that single NR response element motif, but also binding to function by affecting regulatory DNA. Human single nucleo-
motifs for other hepatic transcription factors including other NRs. tide polymorphisms (SNPs) that alter response elements for
The composite nature of these cis‐acting sites theoretically allows NRs, as described for HNF4 in liver [85] and PPARγ in adipose
distinct multiprotein complexes at different elements to fine‐tune tissue [105], could determine differential target gene regulation
regulation based on crosstalk from multiple inputs. This complex among individuals. Such variants would be relevant to precision
transcriptional regulation by crosstalk at composite elements is medicine approaches for individualized disease prediction and
only beginning to be understood. Given the abundance of impor- treatments.
tant hepatic NRs and their frequent co‐occupancy of regulatory
elements at target genes, the liver represents an excellent physio-
logical system to investigate these important issues.
This chapter focused on NRs themselves, but there are also CONCLUSION
hundreds of NR coregulators, which do not bind DNA directly
but are recruited by transcription factors. They serve to modulate Liver NRs play key roles in homeostasis, mediating many
the rate of transcription of nearby genes by various mechanisms, essential hepatic functions. NRs regulate the metabolism of glu-
such as altering chromatin structure (SWI/SNF), histone modifi- cose, fatty acids, cholesterol, and bile acids, as well as the
cations (acetyl transferase [HAT] and deacylatase [HDAC] enzy- detoxification and clearance of xenobiotics. By sensing hor-
matic activity), recruiting the basal transcription machinery with mones, metabolites, and drugs, many NRs allow the liver to
RNA polymerase 2, and facilitating transcriptional elongation respond appropriately to environmental changes. Other circa-
versus pausing. A well‐studied co‐repressor of many NR target dian NRs allow the liver to anticipate such changes based on
genes is the NCoR/HDAC3 complex [102], while prominent co‐ diurnal rhythms. Existing and novel drugs targeting NRs will
activators include the histone acetyltransferase p300/CBP, the play an increasing role in the treatment of liver diseases.
steroid receptor coactivators (SRCs) family, and PGC1s [103].
Like NRs, manipulations of their co‐regulators can likewise affect
liver metabolic homeostasis, but targeting of co‐regulators thera- REFERENCES
peutically is even more complex. It as an area of active investiga-
tion to elucidate mechanisms by which NRs and co‐regulators Note: Important primary references were regretfully omitted due to constraints of
function at the genome to regulate transcription. space. Many of the excellent review articles below include relevant citations of
primary literature.

1. Lazar, M.A. Maturing of the nuclear receptor family. J Clin Invest, 2017;
127(4):1123–5.
NUCLEAR RECEPTORS IN LIVER DISEASE 2. Evans, R.M., Mangelsdorf, D.J. Nuclear receptors, RXR, and the big bang.
Cell, 2014:157(1):255–66.
3. Mangelsdorf, D.J., Thummel, C., Beato, M. et al. The nuclear receptor super-
NRs have many potential roles in liver disease, as noted above. family: the second decade. Cell, 1995;83(6):835–9.
Targeting NRs like VDR and RARs in HSCs may have antifibrotic 4. Nuclear Receptors Nomenclature Committee. A unified nomenclature system
roles in cirrhosis, while targeting bile acid‐related NRs like for the nuclear receptor superfamily. Cell, 1999;97(2):161–3.
28:  Hepatic Nuclear Receptors 349

5. Wang, H., Chen, J., Hollister, K., Sowers, L.C., and Forman, B.M. 32. Mackowiak, B., Hodge, J., Stern, S., and Wang, H. The roles of xenobiotic
Endogenous bile acids are ligands for the nuclear receptor FXR/BAR. Mol receptors: beyond chemical disposition. Drug Metab Dispos Biol Fate Chem,
Cell, 1999;3(5):543–53. 2018;46(9):1361–71.
6. The Human Protein Atlas. https://www.proteinatlas.org/ (Accessed Feb 26, 33. Ma, Z., Deng, C., Hu, W. et al. Liver X receptors and their agonists: targeting
2019). for cholesterol homeostasis and cardiovascular diseases. Curr Issues Mol
7. Melé, M., Ferreira, P.G., Reverter, F. et al. Human genomics. The human tran- Biol, 2017;22:41–64.
scriptome across tissues and individuals. Science, 2015;348(6235):660–5. 34. Fessler, M.B. The challenges and promise of targeting the liver X receptors
8. Malik, R. and Hodgson, H. The relationship between the thyroid gland and for treatment of inflammatory disease. Pharmacol Ther, 2018;181:1–12.
the liver. QJM Int J Med, 2002;95(9):559–69. 35. Schultz, J.R., Tu, H., Luk, A. et al. Role of LXRs in control of lipogenesis.
9. Sinha, R.A., Singh, B.K., and Yen, P.M. Direct effects of thyroid hormones Genes Dev, 2000;14(22):2831–8.
on hepatic lipid metabolism. Nat Rev Endocrinol, 2018;14(5):259–69. 36. Lee, F.Y., Lee, H., Hubbert, M.L., Edwards, P.A., and Zhang, Y. FXR, a mul-
10. Maia, A.L., Goemann, I.M., Meyer, E.L.S., and Wajner, S.M. Deiodinases: tipurpose nuclear receptor. Trends Biochem Sci, 2006;31(10):572–80.
the balance of thyroid hormone: type 1 iodothyronine deiodinase in human 37. Preidis, G.A., Kim, K.H., and Moore, D.D. Nutrient‐sensing nuclear recep-
physiology and disease. J Endocrinol, 2011;209(3):283–97. tors PPARα and FXR control liver energy balance. J Clin Invest,
11. Schussler, G.C. The thyroxine‐binding proteins. Thyroid, 2000;10(2):141–9. 2017;127(4):1193–201.
12. Kuo, T., McQueen A., Chen T‐C., and Wang J‐C. Regulation of glucose home- 38. Wang, H., He, Q., Wang, G., Xu, X., and Hao, H. FXR modulators for entero-
ostasis by glucocorticoids, in: Glucocorticoid Signaling: From Molecules to hepatic and metabolic diseases. Expert Opin Ther Pat, 2018;28(11):765–82.
Mice to Man, (eds. J‐C. Wang and C. Harris), Springer New York, New York, 39. Soccio, R.E., Chen, E.R., and Lazar, M.A. Thiazolidinediones and the promise
NY, 2015, pp. 99–126. doi.org/10.1007/978‐1‐4939‐2895‐8_5. of insulin sensitization in type 2 diabetes. Cell Metab, 2014;20(4):573–91.
13. Woods, C.P., Hazlehurst, J.M., and Tomlinson, J.W. Glucocorticoids and non‐ 40. Kersten, S. Integrated physiology and systems biology of PPARα. Mol
alcoholic fatty liver disease. J Steroid Biochem Mol Biol, 2015;154:94–103. Metab, 2014;3(4):354–71.
14. Newton, R. and Holden, N.S. Separating transrepression and transactivation: 41. Vernia, S., Cavanagh‐Kyros, J., Garcia‐Haro, L. et al. The PPARα‐FGF21
a distressing divorce for the glucocorticoid receptor? Mol Pharmacol, hormone axis contributes to metabolic regulation by the hepatic JNK signal-
2007;72(4):799–809. ing pathway. Cell Metab, 2014;20(3):512–25.
15. Cole, T.J. and Young, M.J. 30 years of the mineralocorticoid receptor: min- 42. Chen, J., Montagner, A., Tan, N.S., and Wahli, W. Insights into the role of
eralocorticoid receptor null mice: informing cell‐type‐specific roles. J PPARβ/δ in NAFLD. Int J Mol Sci, 2018;19(7).
Endocrinol, 2017;234(1):T83–92. 43. Ye, J.‐M., Dzamko, N., Cleasby, M.E. et al. Direct demonstration of lipid
16. Ma, W‐L., Lai, H‐C., Yeh, S., Cai, X., and Chang, C. Androgen receptor sequestration as a mechanism by which rosiglitazone prevents fatty‐acid‐
roles in hepatocellular carcinoma, fatty liver, cirrhosis and hepatitis. Endocr induced insulin resistance in the rat: comparison with metformin.
Relat Cancer, 2014;21(3):R165–182. Diabetologia, 2004;47(7):1306–13.
17. Brie, B., Ramirez, M.C., De Winne C. et al. Brain control of sexually dimor- 44. Chalasani, N., Younossi, Z., Lavine, J.E. et al. The diagnosis and manage-
phic liver function and disease: the endocrine connection. Cell Mol ment of non‐alcoholic fatty liver disease: Practice guidance from the
Neurobiol, 2019;39(2):169–80. American Association for the Study of Liver Diseases. Hepatol Baltim Md,
18. Barros, R.P.A. and Gustafsson, J.‐Å. Estrogen receptors and the metabolic 2018;67(1):328–57.
network. Cell Metab, 2011;14(3):289–99. 45. Fan, W., Waizenegger, W., Lin, C.S. et al. PPARδ Promotes running endur-
19. Qiu, S., Vazquez, J.T., Boulger, E. et al. Hepatic estrogen receptor α is criti- ance by preserving glucose. Cell Metab, 2017;25(5):1186–93.
cal for regulation of gluconeogenesis and lipid metabolism in males. Sci Rep, 46. Heald, M. and Cawthorne, M.A. Dual acting and pan‐PPAR activators as
2017;7(1):1661. potential anti‐diabetic therapies. Handb Exp Pharmacol, 2011;
20. Han, S., Komatsu, Y., Murayama, A. et al. Estrogen receptor ligands amelio- (203):35–51.
rate fatty liver through a nonclassical estrogen receptor/liver X receptor path- 47. Joshi, S.R. Saroglitazar for the treatment of dyslipidemia in diabetic patients.
way in mice. Hepatology, 2014;59(5):1791–802. Expert Opin Pharmacother, 2015;16(4):597–606.
21. Umetani, M. and Shaul, P.W. 27‐Hydroxycholesterol: the first identified 48. Ratziu, V., Harrison, S.A., Francque, S. et al. Elafibranor, an agonist of the
endogenous SERM. Trends Endocrinol Metab TEM, 2011;22(4):130–5. peroxisome proliferator‐activated receptor‐α and ‐δ, induces resolution of
22. Lin, H‐Y., Yu, I‐C., Wang, R‐S. et al. Increased hepatic steatosis and insulin nonalcoholic steatohepatitis without fibrosis worsening. Gastroenterology,
resistance in mice lacking hepatic androgen receptor. Hepatology, 2016;150(5):1147–1159.e5.
2008;47(6):1924–35. 49. Zardi, E.M., Navarini, L., Sambataro, G. et al. Hepatic PPARs: their role in liver
23. Krycer, J.R. and Brown, A.J. Cross‐talk between the androgen receptor and physiology, fibrosis and treatment. Curr Med Chem,. 2013;20(27):3370–96.
the liver x receptor implications for cholesterol homeostasis. J Biol 50. Saeed, A., Hoekstra, M., Hoeke, M.O., Heegsma, J., and Faber, K.N. The
Chem,2011;286(23):20637–47. interrelationship between bile acid and vitamin A homeostasis. Biochim
24. Thaler, M.A., Seifert‐Klauss, V. and Luppa, P.B. The biomarker sex hor- Biophys Acta Mol Cell Biol Lipids, 2017;1862(5):496–512.
mone‐binding globulin ‐ from established applications to emerging trends in 51. Assis, D.N., Abdelghany, O., Cai, S.‐Y. et al. Combination therapy of all‐
clinical medicine. Best Pract Res Clin Endocrinol Metab, 2015;29(5):749–60. trans retinoic acid with ursodeoxycholic acid in patients with primary scle-
25. Gascon‐Barré, M., Demers, C., Mirshahi, A., Néron S., Zalzal, S., and rosing cholangitis: a human pilot study. J Clin Gastroenterol, 2017;
Nanci, A. The normal liver harbors the vitamin D nuclear receptor in non- 51(2):e11–6.
parenchymal and biliary epithelial cells. Hepatology, 2003;37(5):1034–42. 52. Saeed, A., Dullaart, R.P.F., Schreuder, T.C.M.A., Blokzijl, H., and Faber,
26. Ding, N., Liddle, C., Evans, R.M., and Downes, M. Hepatic actions of K.N. Disturbed vitamin A metabolism in non‐alcoholic fatty liver disease
Vitamin D receptor ligands: an unexpected solution to chronic liver disease? (NAFLD). Nutrients, 2017;10(1).
Expert Rev Clin Pharmacol, 2013;6(6):597–9. 53. Trépo, E., Romeo, S., Zucman‐Rossi, J., and Nahon, P. PNPLA3 gene in
27. Yan, J. and Xie, W. A brief history of the discovery of PXR and CAR as liver diseases. J Hepatol, 2016;65(2):399–412.
xenobiotic receptors. Acta Pharm Sin B, 2016;6(5):450–2. 54. Trasino, S.E., Tang X‐H., Jessurun, J., and Gudas, L.J. A retinoic acid recep-
28. Zanger, U.M., Turpeinen, M., Klein, K., Schwab, M. Functional pharmaco- tor β2 agonist reduces hepatic stellate cell activation in non‐alcoholic fatty
genetics/genomics of human cytochromes P450 involved in drug biotrans- liver disease. J Mol Med Berl Ger, 2016;94(10):1143–51.
formation. Anal Bioanal Chem, 2008;392(6):1093–108. 55. Wagner, C.E., Jurutka, P.W., Marshall, P.A., and Heck, M.C. Retinoid X
29. Timsit, Y.E. and Negishi, M. Coordinated regulation of nuclear receptor receptor selective agonists and their synthetic methods. Curr Top Med Chem,
CAR by CCRP/DNAJC7, HSP70 and the ubiquitin–proteasome system. 2017;17(6):742–67.
PloS One, 2014;9(5):e96092. 56. Stein, S. and Schoonjans, K. Molecular basis for the regulation of the nuclear
30. Suino, K., Peng, L., Reynolds, R. et  al. The nuclear xenobiotic receptor receptor LRH‐1. Curr Opin Cell Biol, 2015;33:26–34.
CAR: structural determinants of constitutive activation and heterodimeriza- 57. Miranda, D.A., Krause, W.C., Cazenave‐Gassiot, A. et al. LRH‐1 regulates
tion. Mol Cell, 2004;16(6):893–905. hepatic lipid homeostasis and maintains arachidonoyl phospholipid pools
31. Squires, E.J., Sueyoshi, T., and Negishi, M. Cytoplasmic localization of critical for phospholipid diversity. JCI Insight, 2018;3(5).
pregnane X receptor and ligand‐dependent nuclear translocation in mouse 58. Zhang, Y., Hagedorn, C.H., and Wang, L. Role of nuclear receptor SHP in
liver. J Biol Chem, 2004;279(47):49307–14. metabolism and cancer. Biochim Biophys Acta, 2011;1812(8):893–908.
350 THE LIVER:  REFERENCES

59. Goodwin, B., Jones, S.A., Price, R.R. et  al. A regulatory cascade of the 82. Gonzalez, F.J. Regulation of hepatocyte nuclear factor 4 alpha‐mediated
nuclear receptors FXR, SHP‐1, and LRH‐1 represses bile acid biosynthesis. transcription. Drug Metab Pharmacokinet, 2008;23(1):2–7.
Mol Cell, 2000;6(3):517–26. 83. Armour, S.M., Remsberg, J.R., Damle, M. et al. An HDAC3‐PROX1 core-
60. Lee, Y‐K., Schmidt, D.R., Cummins, C.L. et al. Liver receptor homolog‐1 pressor module acts on HNF4α to control hepatic triglycerides. Nat
regulates bile acid homeostasis but is not essential for feedback regulation of Commun, 2017;8(1):549.
bile acid synthesis. Mol Endocrinol, 2008;22(6):1345–56. 84. Tirona, R.G., Lee, W., Leake, B.F. et  al. The orphan nuclear receptor
61. Kerr, T.A., Saeki, S., Schneider, M. et  al. Loss of nuclear receptor SHP HNF4alpha determines PXR‐ and CAR‐mediated xenobiotic induction of
impairs but does not eliminate negative feedback regulation of bile acid syn- CYP3A4. Nat Med, 2003;9(2):220–4.
thesis. Dev Cell, 2002;2(6):713–20. 85. Hwang‐Verslues, W.W. and Sladek, F.M. HNF4α‐‐role in drug metabolism
62. Yin, L., Wu, N., and Lazar, M.A. Nuclear receptor Rev‐erbalpha: a heme and potential drug target? Curr Opin Pharmacol, 2010;10(6):698–705.
receptor that coordinates circadian rhythm and metabolism. Nucl Recept 86. Shyr, C.‐R., Kang, H.‐Y., Tsai, M.‐Y. et  al. Roles of testicular orphan
Signal, 2010;8:e001. nuclear receptors 2 and 4 in early embryonic development and embryonic
63. Huang, R.‐C. The discoveries of molecular mechanisms for the circadian stem cells. Endocrinology, 2009;150(5):2454–62.
rhythm: the 2017 Nobel Prize in physiology or medicine. Biomed J, 87. Zhou, X.E., Suino‐Powell, K.M., Xu, Y. et al. The orphan nuclear receptor
2018;41(1):5–8. TR4 is a vitamin A‐activated nuclear receptor. J Biol Chem,
64. Yang, X., Downes, M., Yu, R.T. et al. Nuclear receptor expression links the 2011;286(4):2877–85.
circadian clock to metabolism. Cell, 2006;126(4):801–10. 88. O’Geen H., Lin, Y.‐H., Xu, X. et al. Genome‐wide binding of the orphan
65. Cho, H., Zhao, X., Hatori, M. et al. Regulation of circadian behaviour and nuclear receptor TR4 suggests its general role in fundamental biological
metabolism by REV‐ERB‐α and REV‐ERB‐β. Nature, processes. BMC Genomics, 2010;11:689.
2012;485(7396):123–7. 89. Kim, E., Liu, N.‐C., Yu, I.‐C. et  al. Metformin inhibits nuclear receptor
66. Bugge, A., Feng, D., Everett, L.J. et al. Rev‐erbα and Rev‐erbβ coordinately TR4‐mediated hepatic stearoyl‐CoA desaturase 1 gene expression with
protect the circadian clock and normal metabolic function. Genes Dev, altered insulin sensitivity. Diabetes, 2011;60(5):1493–503.
2012;26(7):657–67. 90. Liu, N.‐C., Lin, W.‐J., Yu, I.‐C. et  al. Activation of TR4 orphan nuclear
67. Fang, B. and Lazar, M.A. Dissecting the Rev‐erbα cistrome and the mecha- receptor gene promoter by cAMP/PKA and C/EBP signaling. Endocrine,
nisms controlling circadian transcription in liver. Cold Spring Harb Symp 2009;36(2):211–7.
Quant Biol, 2015;80:233–8. 91. Kang, H.S., Okamoto, K., Kim Y‐S. et al. Nuclear orphan receptor TAK1/
68. Gerhart‐Hines, Z. and Lazar, M.A. Rev‐erbα and the circadian transcriptional TR4‐deficient mice are protected against obesity‐linked inflammation,
regulation of metabolism. Diabetes Obes Metab, 2015;17 Suppl 1:12–6. hepatic steatosis, and insulin resistance. Diabetes, 2011;60(1):177–88.
69. Solt, L.A. and Burris, T.P. Action of RORs and their ligands in (patho)physi- 92. Lin, S.‐J., Yang, D.‐R., Yang, G. et al. TR2 and TR4 orphan nuclear recep-
ology. Trends Endocrinol Metab, 2012;23(12):619–27. tors: an overview. Curr Top Dev Biol, 2017;125:357–73.
70. Sato, T.K., Panda, S., Miraglia, L.J. et  al. A functional genomics strategy 93. Avram, D., Ishmael, J.E., Nevrivy, D.J. et  al. Heterodimeric interactions
reveals Rora as a component of the mammalian circadian clock. Neuron, between chicken ovalbumin upstream promoter‐transcription factor family
2004;43(4):527–37. members ARP1 and ear2. J Biol Chem, 1999;274(20):14331–6.
71. Zhang, Y., Papazyan, R., Damle, M. et al. The hepatic circadian clock fine‐ 94. Kruse, S.W., Suino‐Powell, K., Zhou, X.E. et al. Identification of COUP‐
tunes the lipogenic response to feeding through RORα/γ. Genes Dev, TFII orphan nuclear receptor as a retinoic acid‐activated receptor. PLoS
2017;31(12):1202–11. Biol, 2008;6(9):e227.
72. Kim, D‐K., Ryu, D., Koh, M. et al. Orphan nuclear receptor estrogen‐related 95. Ashraf, U.M., Sanchez, E.R. and Kumarasamy, S. COUP‐TFII revisited: Its
receptor γ (ERRγ) is key regulator of hepatic gluconeogenesis. J Biol Chem, role in metabolic gene regulation. Steroids, 2019;141:63–9.
2012;287(26):21628–39. 96. Planchais, J., Boutant, M., Fauveau, V. et al. The role of chicken ovalbumin
73. Schreiber, S.N., Emter, R., Hock, M.B. et al. The estrogen‐related receptor upstream promoter transcription factor II in the regulation of hepatic fatty
alpha (ERRalpha) functions in PPARgamma coactivator 1alpha acid oxidation and gluconeogenesis in newborn mice. Am J Physiol
(PGC‐1alpha)‐induced mitochondrial biogenesis. Proc Natl Acad Sci U S A, Endocrinol Metab. 2015;308(10):E868–78.
2004;101(17):6472–7. 97. Kurakula, K., Koenis, D.S., van Tiel, C.M., and de Vries, C.J.M. NR4A
74. Hummasti, S. and Tontonoz, P. Adopting new orphans into the family of nuclear receptors are orphans but not lonesome. Biochim Biophys Acta,
metabolic regulators. Mol Endocrinol Baltim Md, 2008;22(8):1743–53. 2014;1843(11):2543–55.
75. Luo,, J., Sladek, R., Carrier, J., Bader J‐A., Richard, D., and Giguère, V. 98. Maxwell, M.A. and Muscat GE.O. The NR4A subgroup: immediate early
Reduced fat mass in mice lacking orphan nuclear receptor estrogen‐related response genes with pleiotropic physiological roles. Nucl Recept Signal,
receptor alpha. Mol Cell Biol, 2003;23(22):7947–56. 2006;4:e002.
76. Audet‐walsh, É. and Giguére, V. The multiple universes of estrogen‐related 99. Pei, L., Waki, H., Vaitheesvaran, B., Wilpitz, D.C., Kurland, I.J., and
receptor α and γ in metabolic control and related diseases. Acta Pharmacol Tontonoz, P. NR4A orphan nuclear receptors are transcriptional regulators
Sin, 2015;36(1):51–61. of hepatic glucose metabolism. Nat Med, 2006;12(9):1048–55.
77. Fang, B., Mane‐Padros, D., Bolotin, E., Jiang, T., and Sladek, F.M. 100. Zhang, L., Wang, Q., Liu, W., Liu, F., Ji, A., and Li, Y. The orphan nuclear
Identification of a binding motif specific to HNF4 by comparative analysis receptor 4A1: a potential new therapeutic target for metabolic diseases. J
of multiple nuclear receptors. Nucleic Acids Res, 2012;40(12):5343–56. Diabetes Res, 2018;2018:9363461.
78. Yuan, X., Ta, T.C., Lin, M. et  al. Identification of an endogenous ligand 101. Berrabah, W., Aumercier, P., Lefebvre, P., and Staels, B. Control of nuclear
bound to a native orphan nuclear receptor. PloS One, 2009;4(5):e5609. receptor activities in metabolism by post‐translational modifications. FEBS
79. Li, J., Ning, G., and Duncan, S.A. Mammalian hepatocyte differentiation Lett, 2011;585(11):1640–50.
requires the transcription factor HNF‐4α. Genes Dev, 2000;14(4):464–74. 102. Emmett, M.J. and Lazar, M.A. Integrative regulation of physiology by his-
80. Hayhurst, G.P., Lee, Y.H., Lambert, G., Ward, J.M., and Gonzalez, F.J. tone deacetylase 3. Nat Rev Mol Cell Biol, 2019;20(2):102–15.
Hepatocyte nuclear factor 4alpha (nuclear receptor 2A1) is essential for 103. Dasgupta, S., Lonard, D.M., and O’Malley, B.W. Nuclear receptor coacti-
maintenance of hepatic gene expression and lipid homeostasis. Mol Cell vators: master regulators of human health and disease. Annu Rev Med,
Biol, 2001;21(4):1393–403. 2014;65:279–92.
81. Shih, D.Q., Dansky, H.M., Fleisher, M., Assmann, G., Fajans, S.S., and 104. Achermann, J.C., Schwabe, J., Fairall, L., and Chatterjee, K. Genetic disor-
Stoffel, M. Genotype/phenotype relationships in HNF‐4alpha/MODY1: ders of nuclear receptors. J Clin Invest, 127(4):1181–92.
haploinsufficiency is associated with reduced apolipoprotein (AII), apolipo- 105. Soccio, R.E., Chen, E.R., Rajapurkar, S.R. et al. Genetic variation deter-
protein (CIII), lipoprotein(a), and triglyceride levels. Diabetes, 2000; mines PPARγ function and anti‐diabetic drug response in vivo. Cell.
49(5):832–7. 2015;162(1):33–44.
Molecular Cholestasis
29 Paul Gissen1 and Richard J. Thompson2
1
UCL Great Ormond Street Institute of Child Health and Great Ormond Street Hospital for Children,
London, UK
2
Institute of Liver Studies, King’s College London, UK

The molecular bases of cholestasis have been unraveled through Disorders of primary BA synthesis and
the detailed investigation of individual patients and cohort conjugation presenting with infantile
studies. The most obvious examples have resulted from
­
genomic and molecular biology investigation. Major scientific
cholestasis
advances that participated in this increase included elucidation Deficiencies of enzymes involved in primary BA synthesis
of the cellular mechanisms of synthesis and secretion of vari- cause some instances of neonatal cholestasis in which serum
ous constituents of bile, facilitated by improved genomic meth- gamma‐glutamyl transpeptidase (GGT) values do not rise,
ods that have led to identification of genes mutated in patients despite conjugated hyperbilirubinemia (see Table  29.1). As at
with several familial forms of cholestasis. Discovery of the least 16 enzymes are involved in conversion of cholesterol into
genetic causes in patients with severe clinical manifestations primary BA, disorders of primary BA synthesis are not uncom-
has gradually led to an appreciation of the scope of the phe- mon [4]. Bile acid synthesis disorders can be diagnosed by iden-
nomena associated with abnormalities in individual genes. tification of changes in BA composition within serum and urine
More recently, the aim of clinical research has been to elucidate (lack of primary BA, high concentrations of intermediary
how genes mutated in severe forms of childhood cholestasis metabolites). Such disorders are important to recognize because
contribute to multifactorial, later onset, liver diseases. In this many can be easily treated by supplementing the diet with pri-
chapter we discuss the molecular and clinical features of the mary BA, which can be lifesaving. The fed BA, through the
heritable forms of cholestasis. action of the nuclear receptor FXR, downregulate BA synthesis,
and fewer toxic intermediate species are generated [5].

INHERITED FORMS OF INTRAHEPATIC Disorders associated with abnormalities


CHOLESTASIS of hepatocyte function
Transmembrane transporter proteins are important for normal bile
Cholic acid (CA), chenodeoxycholic acid (CDCA), the primary secretion, therefore transporter defects play a central role in the
bile acids (BAs), are synthesized in the liver from cholesterol by pathogenesis of cholestasis. Many of the proteins crucial to bile‐
a number of steps occurring in several organelles, including component transport belong to the family of ATP‐binding cassette
­peroxisomes. Failure of primary BA synthesis leads to accumu- (ABC) transporters. Some of these defects manifest in early life
lations of toxic intermediate metabolites, with hepatocellular and others may become apparent only in certain situations such as
damage and cholestasis [1]. For this reason, cholestasis can be a pregnancy or after ingestion of specific pharmaceutical com-
feature of single enzyme defects and peroxisomal biogenesis pounds. Defects in different transports are considered below and
disorders. Some defects of cholesterol biogenesis, such as 3β‐ summarized in Table 29.2.
hydroxysteroid Δ7‐reductase deficiency, also may cause choles- Bile constituents (BA, bilirubin, cholesterol, phospholipids,
tasis due to production of abnormal BA [2, 3]. and other products of metabolism) are secreted into biliary

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
352 THE LIVER:  INHERITED FORMS OF INTRAHEPATIC CHOLESTASIS

Table 29.1  Disorders of primary BA biosynthesis


Response to treatment with
Enzyme defect Clinical presentation primary BA Reference
CYP27A1 deficiency Neonatal cholestasis in some patients; juvenile Good [6, 7]
(cerebrotendinous xanthomatosis) cataracts, tendinous xanthomata, and progressive
neuropsychiatric signs including cerebellar
ataxia, spastic paraparesis, and dementia
3β‐Hydroxy‐C27‐steroid Cholestasis in infancy, hepatomegaly, variable Good [8–12]
oxidoreductase deficiency course. Commonest BA synthesis disorder
Δ4‐3‐oxosteroid‐5β‐reductase Early neonatal cholestasis, rapid progression of Good [13]
deficiency liver disease
Oxysterol 7α‐hydroxylase Infantile cholestasis progressing to cirrhosis and Poor, requiring liver [14, 15]
deficiency liver failure (single patient identified). Same gene transplantation (single known
(SPG5A) mutated in hereditary spastic paraplegia patient, CDCA given)
2‐methylacyl‐CoA racemase Infantile mild cholestasis; adult sensorimotor Good, may require restriction [16–18]
deficiency neuropathy of phytanic and pristanic acids
Bile acid‐CoA ligase deficiency Neonatal cholestasis, growth failure, fat‐soluble May require conjugated BA [1, 19]
vitamin deficiency
Bile acid‐CoA: Amino acid N‐ Growth failure and coagulopathy of fat‐soluble May require conjugated BA [20]
acyltransferase deficiency (familial vitamin deficiency, without jaundice
hypercholanemia)

NA, not applicable; CA, cholic acid; CDCA, chenodeoxycholic acid; GGT, gamma glutamyl transpeptidase.

c­analiculi in an energy‐dependent manner. Transmembrane clear [33–37]. FIC1 appears to translocate aminophospholipids
transporter proteins mediate the secretory function of hepato- such as phosphatidylserine (PS) from the outer to the inner
cytes and biliary epithelial cells. Genes encoding some of these ­leaflet of the plasma membrane bilayer [36].
transporters were found to be mutated in progressive familial In mice homozygous for an Atp8b1 missense substitution
intrahepatic cholestasis (PFIC), a clinical syndrome manifest by mutation (mimicking clinically severe FIC1 deficiency), PS,
nonremitting conjugated hyperbilirubinemia, often with low or cholesterol, and ectoenzymes were easily lost from canalicular
normal serum GGT values and liver scarring; typically requir- membranes during increased BA flux [38]. Membrane instabil-
ing liver transplantation before adulthood. Many patients with ity in FIC1 deficiency may impair bile salt transport and
PFIC also have reduced concentrations of primary BA in bile decrease hepatocyte resistance to bile salts. FIC1 also may, via
[21, 22]. Mutations in two genes, ATP8B1 and ABCB11, were FXR, upregulate transcription of ABCB11 [37].
initially found to cause ­respectively FIC1 and bile salt export Patients with severe FIC1 deficiency present in the first year
pump (BSEP) deficiencies. A spectrum of disease severity is of life with pruritus and cholestatic jaundice [31, 35]. Extrahepatic
recognized in association with mutation in ATP8B1 and manifestations include diarrhea, episodes of pancreatitis, fat‐
ABCB11. Whilst some patients have a classical PFIC pheno- soluble vitamin deficiencies secondary to malabsorption, sensori-
type, others have cholestasis that completely resolves between neural deafness, delay in growth and puberty, and elevated sweat
bouts, a phenotype sometimes termed “benign” recurrent intra- chloride concentration [39–43]. Although cholestasis resolves
hepatic cholestasis [23, 24]. However, patients with apparently after liver transplantation, marked hepatic steatosis is common in
benign disease may progress to PFIC after some years [25, 26]. the graft. Severe diarrhea and growth restriction persist [35]. The
In truth, therefore, both FIC1 and BSEP deficiencies are spectra, etiology of these extrahepatic symptoms is unclear; patients have
and late or relapsing onset does not always predict a benign been shown to have normal pancreatic secretion [44].
clinical course. Heterozygotes for ATP8B1 and ABCB11 muta- Liver biopsy at presentation in patients with FIC1 deficiency
tions are predisposed to intrahepatic cholestasis of pregnancy shows predominantly centrilobular cholestasis that is canalicu-
(ICP) [27, 28]. Intrahepatic cholestasis of pregnancy, a third‐tri- lar rather than hepatocellular. Swelling and multinucleation of
mester disorder, is characterized by pruritus and elevated serum hepatocytes is not common, neither is portal‐tract cholestasis.
concentrations of BA [29] and causes fetal disease, with fetal With progression of disease, fibrosis and portal–portal bridging
distress, premature birth, and stillbirth [29]. GGT values are develops. Unlike canalicular bile of patients with BSEP defi-
normal in 70% of patients with ICP, implying that at least two ciency, that of FIC1 patients is coarsely granular on transmis-
different processes can lead to ICP [30]. sion electron microscopy [22]. This may reflect shedding of
canalicular wall into lumen, as may deficiency of ectoenzyme
expression along canaliculi [38].
FIC1 deficiency
In one large study ATP8B1 mutations were detected in 30%
Previously, clinically severe FIC1 deficiency was often termed of PFIC (39/130) and 40% (20/50) of patients with relapsing
Byler disease; Byler is the family name of the extended Amish cholestasis. Genotype–phenotype correlation included a greater
kindred first described with this condition [31]. ATP8B1 encodes likelihood of missense substitutions in milder disease than in
FIC1, a member of the type 4 subfamily of P‐type ATPases [32], PFIC, with nonsense mutations only in heterozygous form in
and hence is not an ABC transporter. FIC1 is present in the the former phenotype; an exception, a homozygous frameshift
­apical membrane of many epithelial cells, including hepatocytes mutation, left most of ATP8B1 intact, with loss of only the ter-
and enterocytes. Its role in cholestasis is only now becoming minal 11 amino acid residues [45].
29:  Molecular Cholestasis 353

BSEP deficiency anion transporter). It exports anionic glutathione, glucuronate,


or sulphated conjugates (including bilirubin) from hepatocytes
Bile salt export pump, encoded by ABCB11, is the protein
into canaliculi [61, 62]. MRP2 is expressed on apical mem-
responsible for the ATP‐dependent transport of salts of primary
branes of epithelia other than hepatocytes (proximal renal
BA across the canalicular membrane [46]. BSEP is a member of
tubules, gallbladder, small intestine, bronchi, and placenta) [63,
the ABC family of proteins and also a member of the P glyco-
64]. MRP2 is now regarded as a crucial efflux pump, eliminat-
protein/multidrug resistance (MDR/ABCB) subfamily of trans-
ing conjugates of various toxins and carcinogens from hepato-
porters [47]. As with FIC1 disease, phenotypes of patients with
cytes into bile, from kidney proximal tubular epithelium into
mutations in ABCB11 represent a continuum between mild and
urine, and from intestinal enterocytes into chyme and stool [65].
severe disease [24, 26, 48–50].
Mutations in ABCC2 result in Dubin–Johnson syndrome, a con-
Liver biopsies from patients with severe BSEP deficiency dis-
dition characterized by episodes of cholestatic jaundice without
play a neonatal hepatitis picture at presentation [22]. Canalicular
other clinical biochemistry indication of hepatobiliary injury
bile on ultrastructural study may appear dense, amorphous, or
[66]. Microscopy of liver shows intrahepatocyte deposits of
finely filamentous, but is not loosely and coarsely granular as in
dark brown pigment but no other abnormalities. Patients may
FIC1 disease [22, 51]. Clinical features include intermittent or
present at any age; choleretics such as ursodeoxycholic acid
persistent cholestatic jaundice, pruritus, and growth failure, but
(UDCA) and phenobarbitone are recommended only in particu-
extrahepatic features such as diarrhea and pancreatitis typically
larly severe neonatal cases  [67]. The relatively mild, or even
are absent [35]. Genotype–phenotype correlation in a large
completely absent, clinical phenotype despite absence of MRP2
cohort of patients with severe BSEP deficiency found less severe
may be explained by upregulation of other transporters such as
disease when missense mutations, even in heterozygous form,
MRP3; this compensates for loss of MRP2 as an anion efflux
were present, and suggested that truncating mutations conferred
pump [68]. Abcc2‐deficient rats and mice may be useful in stud-
increased risk of hepatobiliary malignancy [50].
ying pharmacokinetics of various drugs and drug elimination
Concordant with this were the findings [52, 53] that the range
mechanisms [69]. A partial phenocopy of Dubin–Johnson syn-
of severity of clinical phenotype in patients with ABCB11 muta-
drome exists in mice lacking radixin, a cytoskeletal protein
tions can be explained by intracellular localization and differ-
found in canalicular membranes, without which poorly anchored
ences in stability of mutated BSEP and in retention of
Abcc2 may not function optimally [70].
taurocholate transport function. Mutations associated with mild
Rotor syndrome, phenotypically very similar to Dubin–
BSEP deficiency were trafficked normally to the apical domain
Johnson syndrome, also follows an autosomal recessive mode of
of the plasma membrane, where they retained some transport
inheritance. No intrahepatocyte pigment deposits are found in
activity, whilst most mutant BSEP forms associated with severe
Rotor syndrome patients, however, and Dubin–Johnson and
disease were not trafficked to the apical membrane domain and
Rotor syndromes can also be distinguished by studies of bromo-
lacked most transport activity [52].
sulfophthalein and coproporphyrin excretion [71]. Allelism with
Like ATP8B1, carriers of ABCB11 mutations are prone to devel-
Dubin–Johnson syndrome has been excluded in Rotor syndrome
oping ICP [54, 55]. Cholestasis associated with the administration
[72]; instead it has been suggested that a combined impairment
of hormonal agents such as oral contraceptives also is more frequent
of hepatocellular uptake transporters is lacking [73].
in carriers of ABCB11 mutations; inhibition of BSEP by estrogen
and progestogen metabolites may be contributory [56, 57]. ABCB11
mutations, unlike ATP8B1 mutations, predispose to gallstones. This MDR3 deficiency
complication is particularly common with milder disease; an inci-
This disorder results from mutations in ABCB4, which encodes
dence as high as 7/11 was found in benign recurrent intrahepatic
the MDR3, or multidrug resistance protein 3, a P‐glycoprotein
cholestasis‐2 (BRIC‐2) [24]. Although ABCB11 lies in the murine
that flops phospholipids from the internal to external leaflet of
Lith1 major susceptibility locus for cholelithiasis, a study of 810
the canalicular membrane [74, 75]. In mice Abcb4 encodes the
German patients and 718 controls showed no association between
orthologous protein mdr2, and homozygous mutants are a good
ABCB11 mutation and cholelithiasis [58].
model for the study of pathogenesis and testing of therapies in
In mice, knockout of Abcb11 resulted in only mild cholesta-
human MDR3 deficiency [76, 77].
sis unless CA was fed [59]. Higher excretion of BA than
MDR3 deficiency results in impaired translocation of phos-
expected was traced to complementation of Abcb11 by Abcb1,
phatidylcholine (PC) into bile and is associated with a wide
which may transport salts of BA in mice [60].
spectrum of phenotypes ranging from neonatal cholestasis to
Hepatocellular carcinoma seems to be a particularly common
biliary cirrhosis in adults [38, 78–80]. PC in bile is an essential
feature of BSEP deficiency. Complete absence of the protein is
component of the mixed micelles into which salts of BA are
associated with the highest incidence. The same patients are
emulsified, reducing BA detergent activity. Deficiency of
also at the greatest risk of inhibitory anti‐BSEP antibodies after
MDR3 permits BA unchaperoned by PC to damage hepatocytes
liver transplantation. This phenomenon has not been described
and cholangiocytes [75, 78, 81]. MDR3 disease is usually char-
in similar conditions, and can be problematic.
acterized by high serum GGT values and, on biopsy, ductular
reaction and fibrosis, a picture similar to that seen in Abcb4−/−
Dubin–Johnson syndrome
mice [74]. Liver disease progresses to fibrosis and cirrhosis,
The protein encoded by ABCC2, multidrug‐associated protein sometimes with ductopenia, and may require liver transplanta-
2 (MRP2), is a member of the ABC transporter superfamily and tion. Response to UDCA may be seen [79, 80, 82], particularly
also known as the cMOAT (canalicular multispecific organic in milder disease. Obligate carriers of ABCB4 mutations
354 THE LIVER:  INHERITED FORMS OF INTRAHEPATIC CHOLESTASIS

(parents of infants with ABCB4 mutations ablating MDR3 func- patients with a homozygous mutation in BAAT predicted to
tion and liver disease presenting as neonatal hepatitis [NH]) are yield the amino acid residue substitution p.M76V missense
prone to ICP and gallstones [81]. ABCB4 mutations similarly substitution in BAAT, which conjugates BA with glycine
have been found in a subgroup of patients with ICP and and  taurine [20]. The p.M67V mutation was associated in
high  GGT values, and some common ABCB4 polymorphisms dose‐dependent fashion with decreased BA conjugation.
­conduce to development of ICP [83–85]. Certain ABCB4 poly- Furthermore, 6 of 16 clinically‐affected individuals homozy-
morphisms also may mark adverse prognosis in primary biliary gous for a mutation in either TJP2 or BAAT also were
cholangitis [86]. ­heterozygous for a mutation in the second gene [20], thus
Low phospholipid‐associated cholelithiasis (LPAC) is a form ­suggesting oligogenic inheritance in FHCA.
of gallstone disease occurring in younger patients which is asso- Whilst FHCA is associated with seemingly mild missense
ciated with ABCB4 mutations, recurs after cholecystectomy, variation in TJP2, more damaging variants have been associated
and sometimes to respond well to UDCA [87]. Homozygous with a quite different phenotype [94]. In patients with a PFIC
missense mutations or heterozygous nonsense or frameshift phenotype, similar to FIC1 deficiency, protein truncating on
ABCB4 mutations may be found in LPAC patients [88]. UDCA both alleles of TJP2 have been identified. These patients have
may act by upregulating residual expression of MDR3 at the early onset progressive disease requiring liver transplantation.
canalicular membrane, exerting a protective effect against In several cases hepatocellular carcinomata have been described
hydrophobic endogenous BA, and increasing the pool of protec- [95]. As TJP2 is widely expressed in other epithelia it should
tive hydrophilic BA [87]. Hepatolithiasis with intractable chol- not be surprising that extrahepatic manifestations have become
angitis may, however, be associated with MDR3 disease in some evident. It is probably more surprising that they are not more
patients [89, 90]. severe, especially as the homozygous knockout mouse is embry-
As in BSEP disease, MDR3 disease is associated with drug‐ onic lethal.
induced cholestasis (DIC) following administration of amoxi-
cillin, clavulanic acid, and risperidone [91]. Not enough data are MYO5B associated cholestasis
available to explain the pathophysiology of DIC associated with
MYO5B is an atypical myosin involved in intracellular vesicu-
the mutations and polymorphisms so far identified in these
lar transport. As such it is clearly important in apical membrane
genes. Although inhibition of ABCB4 expression has been sug-
assembly [96–99]. The first patients described with mutations in
gested, no evidence of this effect has yet been found [92].
the MYO5B gene (encoding MYO5B) were those with microvil-
lous atrophy (MVA) also known as microvillous inclusion dis-
TJP and BAAT deficiencies ease (MVID) [100–103]. This is a severe form of intestinal
failure, requiring parenteral nutrition (PN). Whilst intestinal
Two forms of familial hypercholanemia (FHCA) were
failure associated with liver disease is common in children
described in Amish families. These involved mutations in
requiring PN from infancy, it was clear that cholestasis was a
TJP2, encoding tight junction protein 2 (TJP2), and BAAT,
particular feature of these children [103]. Reduced expression
encoding bile acid‐CoA:amino acid N‐acyltransferase (BAAT)
of BSEP in the apical membrane of hepatocytes in patients with
[20]. Affected individuals usually have markedly increased
MYO5B associated MVA has been demonstrated. More recently
serum primary BA concentrations, severe pruritus, fat malab-
patients have been described with little or no intestinal disease,
sorption leading to coagulopathy and rickets, and failure to
but instead isolated cholestasis, with mutation in MYO5B. No
thrive [93]. While abnormal BA species are not present in
genotype–phenotype correlations are evident, and it is not yet
serum, in one form of FHCA, that owing to lesions in BAAT,
clear why intestinal failure can predominate in some patients
BA are inadequately conjugated [20, 93]. Patients intermit-
and not others.
tently have episodes of hepatitis, and liver biopsy may find a
nonspecific pattern of portal and lobular inflammation with
Neonatal sclerosing cholangitis
spotty hepatocyte necrosis and mild ductular proliferation [93].
Serum BA concentrations fall and pruritus improves with Sclerosing cholangitis, of neonatal onset (NSC), was first
UDCA treatment [93]. The clinical course is unpredictable and described many years ago. There was a strong suggestion that
symptoms may resolve spontaneously [20]. many cases had a genetic etiology. A recent study found that a
The molecular pathology of familial hypercholanemia is quarter of such patients harbored mutations in the DCDC2 gene
complex. A homozygous TJP2 missense mutation predicted to [104, 105]. This gene encodes a member of the doublecortin
yield the amino acid residue substitution p.V48A was identi- family of proteins, and has previously been implicated in dys-
fied in 11 out of 17 subjects [20]. TJP2 is a tight junction‐­ lexia and shown to be mutated in nephronophthisis [106, 107].
associated protein and biophysical studies suggested that the The protein appears to be important in the flagellary transport
altered protein conformation caused by the p.V48A mutation body, itself a component of cilial assembly. DCDC2‐associated
may lead to increased paracellular permeability, with reflux disease is therefore a ciliopathy. The majority of patients have
of  BA into plasma. Morphological abnormalities of tight rapidly progressive liver disease, requiring early transplanta-
junctions were identified in patients’ liver tissues [20].
­ tion. Although other patients have been described previously
Homozygosity for this mutation was also found in three unaf- with a renal phenotype, many NSC have normal renal function
fected siblings, suggesting lack of penetrance for this variant. and imaging. Several such patients have undergone uneventful
Hence, the investigators searched for further genes conferring liver transplantation with use of calcineurin inhibitors, without
susceptibility to development of cholestasis and identified five obvious complications.
29:  Molecular Cholestasis 355

Table 29.2  Disorders associated with abnormalities of hepatocyte function


Diseases and genes Typical presentations and clinical features
FIC1 deficiency 1. Infantile cholestasis, pruritus, normal GGT; bland, canalicular cholestasis on liver biopsy at
Mutations in ATP8B1 presentation, advancing to cirrhosis. Bile coarse and loosely granular on transmission electron
microscopy. Diarrhea, episodes of pancreatitis, malabsorption, sensorineural deafness, delay in growth
and puberty, short stature, elevated sweat chloride concentrations. Severe mutations on both alleles.
2. Periodic bouts of cholestasis, intense pruritus. Classically cholestasis resolves between the relapses,
but may progress. Milder mutations, usually on both alleles.
3. Cholestasis of pregnancy. Resolution of symptoms after delivery of the baby. Usually associated with
monoallelic mutation.
BSEP deficiency 1. Infantile cholestasis, pruritus, normal GGT. Giant‐cell hepatitis on liver biopsy at presentation,
Mutations in ABCB11 advancing to cirrhosis. Growth failure. Increased risk of hepatocellular carcinoma and
cholangiocarcinoma. Severe mutations on both alleles.
2. Periodic bouts of cholestasis, intense pruritus. Classically cholestasis resolves between the relapses,
but may progress. Milder mutations, usually on both alleles.
3. Cholestasis of pregnancy. Usually symptoms resolve after delivery of the baby. Usually associated
with monoallelic mutation.
4. Cholestasis induced by the administration of hormonal contraceptives, and other drugs. Usually
associated with monoallelic mutation.
Dubin–Johnson syndrome Recurrent episodes of cholestatic jaundice without liver disease. Mutations on both alleles
Mutations in ABCC2
MDR3 deficiency 1. Early onset cholestatic jaundice with high GGT. Severe mutations on both alleles.
Mutations in ABCB4 2. Cholangiopathy, similar to “small duct primary sclerosing cholangitis”. Rapidity of progression
dependent on degree of functional loss. May be associated with severe mutation on one allele, or
milder mutations on both.
3. Low phospholipid‐associated cholelithiasis, with hepatolithiasis a well as cholecystolithiasis. Usually
monoalleleic mutation.
4. Cholestasis of pregnancy with raised GGT. Usually associated with monoallelic mutations, but can be
first presentation of more severe disease.
5. Cholestasis induced by the administration of hormonal contraceptives and other drugs. Usually
associated with monoallelic mutation.
TJP2 deficiency 1. Early onset, normal GGT cholestasis. Bland, canalicular cholestasis on liver biopsy at presentation,
Mutations in TJP2 advancing rapidly to cirrhosis. Severe mutations on both alleles.
2. Episodic cholestasis, often associated with drug administration. Milder mutations on both alleles.
3. Hypercholanaemia. No liver disease. Usually with specific missense change on both alleles.
MYO5B deficiency 1. Microvillous inclusion disease, causing intestinal failure, often with particularly severe cholestasis.
Mutations in MYO5B Mutations of both alleles.
2. Isolated cholestasis, with variable age of onset and duration. Mutations of both alleles.
Neonatal sclerosing cholangitis Early onset cholangiopathy with rapid progression in most cases. Renal involvement in some cases. Severe
Mutations in DCDC2 in 25% of cases mutations on both alleles.

Syndromic forms of inherited cholestasis clinical manifestations that characterize ALGS in JAG1 muta-
tion‐positive relatives of classical ALGS patients revealed that
Multisystem disorders with a prominent feature of cholestasis liver and cardiac disease exhibited 31% and 41% penetrance,
are delineated below and summarized in Table 29.3. respectively, among mutation carriers [112]. This suggested that
other factors, such as genetic modifier loci, in addition to JAG1
Alagille syndrome mutations determine the development of clinical features in this
Alagille syndrome (ALGS) is an autosomal dominant disorder condition. Complete knockout of Jag1 was embryonically lethal
which affects development of multiple organ systems, including in mice and haploinsufficiency for murine Jag1 did not cause a
cardiovascular, musculoskeletal, gastrointestinal, central nerv- complete ALGS‐like phenotype. A knockout of Notch2 pro-
ous, and renal [108]. A clinical diagnosis of ALGS is made when duced an ALGS mouse model; heterozygous mutations in
intrahepatic bile duct paucity is associated with three NOTCH2 were identified in several JAG1 mutation‐­negative
­distinguishing clinical features such as cardiac disease, abnormal patients with ALGS [120–122]. Further studies in various Notch2
vertebral bodies, typical ocular abnormalities, and characteristic and Jag1 mouse knockouts have extended ­knowledge of the cru-
facial features [109]. Hepatocellular carcinoma is a recognized cial role of this pathway, in particular with respect to gene dos-
complication, occurring in children as young as four years old age, in biliary morphogenesis [123].
[110, 111]. The incidence of ALGS is estimated between 1 in
30 000 to 1 in 70 000 [112]. Mutations in JAG1 are found in more
Arthrogryposis, renal dysfunction and cholestasis syndrome
than 90% of patients with ALGS [113, 114]. JAG1 encodes a
cell‐surface protein, JAG1, a ligand for NOTCH receptors. The Arthrogryposis, renal dysfunction, and cholestasis (ARC)
notch signaling pathway, including both JAGGED1 and syndrome is an autosomal recessive multisystem disorder
NOTCH2, is important in morphogenesis and is conserved in which commonly presents with NH. Patients usually have
evolution [115–119]. Clinical ­features in ALGS vary dramati- associated neurogenic arthrogryposis and renal tubular acido-
cally and intrafamilial differences among penetrance of various sis; however, as these features may be mild, they are easily
clinical features are substantial. In fact, detailed study of the missed early in evaluation. ARC syndrome is caused by
356 THE LIVER:  INHERITED FORMS OF INTRAHEPATIC CHOLESTASIS

mutations in either VPS33B, which encodes VPS33B, or CLAUDIN‐1 deficiency (NISCH syndrome)
VIPAS39 encoding VIPAR [124, 125]. VPS33B and VIPAR
Neonatal ichthyosis and sclerosing cholangitis syndrome
form a stable protein complex with a function in intracellular
(NISCH) is a rare monogenic form of sclerosing cholangitis
protein trafficking and biogenesis of various lysosome‐related
associated with recessive mutations in CLDN1 encoding the
organelles such as keratinocyte lamellar bodies or platelet
tight junction protein claudin‐1 [138, 139]. Sixteen patients and
alpha granules [126–130].
three different mutations have been described so far in families
VPS33B and VIPAR are homologous to VPS33A and
from Morocco and Switzerland [140]. There is significant clini-
VPS16 proteins that form parts of the HOPS and CORVET
cal heterogeneity in patients with NISCH. Whilst some may have
complexes involved in membrane fusion in endocytic pathway
early normal gGT, typically children develop neonatal cholesta-
and biogenesis of early and late endosomes and autophago-
sis with high gGT and hepatomegaly [141]. Liver biopsy find-
somes. Thus, it is hypothesized that VPS33B and VIPAR com-
ings include extensive fibrosis without fatty infiltration or
plexes may be involved in a similar process to HOPS but in
ductular proliferation and a mixture of extracellular and intracel-
different intracellular trafficking pathways. For example, it
lular cholestasis. Transparietal cholangiography may show scle-
was found that integrin beta subunit directly interacted with
rosing cholangitis and portal hypertension after the age of five.
VPS33B and Vps33b knockout mice had abnormalities in
Interestingly, in some patients resolution of liver disease may
αIIbβ3‐mediated fibrinogen endocytosis.
occur with symptomatic treatment [140]. The cutaneous features
ARC patients’ liver biopsy features include bile duct paucity,
of the syndrome include dry and hard skin with large scales par-
lipofuscin deposition, and giant‐cell transformation of hepato-
ticularly prominent on the abdomen and limbs [141]. In addition,
cytes [131]. A striking feature of ARC syndrome is abnormal
patients have substantial alopecia. Other reported clinical mani-
localization of specific apical membrane proteins (e.g. alanyl
festations included intracytoplasmic vacuoles in peripheral blood
aminopeptidase/CD13, dipeptidyl peptidase/CD26, and carci-
eosinophils and enamel dysplasia. Interestingly, all cutaneous
noembryonic antigen / CD66) along biliary canaliculi and in
phenomena subsided after liver transplantation.
renal tubular epithelium [124, 132]. Liver disease in ARC is
Primary tight junction abnormalities also have been identi-
associated with normal gGT, and gGT in hepatocytes also is
fied in a form of familial hypercholanemia (see TJP and BAAT
abnormally sited in ARC [124]. Impaired delivery of apical
deficiencies), with secondary abnormalities in animal models of
membrane proteins to their destination could explain renal tubu-
cholestatic disease [142–146]. Tight junctions are essential to
lar leak of various molecules including glucose, amino acids,
maintain the separation between tissue layers and electrochemi-
and bicarbonate and might underlie the failure of serum gGT
cal gradients across epithelial cell monolayers. Thus, intact tight
values to rise despite cholestasis. Vps33b knockout mice dis-
junction structure is important to prevent bile leaking from bil-
played abnormal localization of multiple ABC transporters
iary canaliculi into blood [147]. Claudin‐1 is expressed at high
including BSEP and ABCG8 but not MDR2 suggesting involve-
levels in mouse liver and kidney; however, a Claudin‐1 deficient
ment of VPS33B‐VIPAR complex in the Rab11a‐dependent
mouse was unsuitable for the investigation of cholestasis as skin
protein trafficking route in the hepatocytes [98, 103, 133]. Wang
abnormalities led to death on the first postnatal day with severe
and colleagues showed that VPS33B is downregulated in the
dehydration [148].
samples from human and mouse hepatocellular carcinoma and
Vps33b liver‐specific knockout mice developed hepatocellular
GRACILE syndrome
carcinoma following progressive hepatitis and fibrosis [134].
In addition to mild dysmorphism, hypotonia (more than 90% Several genetically distinct disorders affecting the mitochon-
of ARC syndrome patients), and developmental delay, abnor- drial respiratory chain can cause liver disease with or without
malities (hypoplasia or absence) of the corpus callosum occur in associated renal tubular abnormalities [149]. GRACILE syn-
~20% [135]. It is now recognized that sensorineural deafness drome (growth retardation, aminoaciduria, cholestasis, iron
can be identified in the majority of ARC patients. Other symp- overload, lactic acidosis, and early death) is a recessively inher-
toms include hypothyroidism and congenital cardiac defects. ited lethal disease to date found only in Finnish populations and
Ichthyosis is seen in most patients and recurrent febrile illnesses known to be caused only by homozygous inheritance of a mis-
occur in many, although the cause of the fevers is often not sense 232A>G (S78G) mutation in BSC1L [150, 151]. BCS1L
found. Interestingly, Akbar and colleagues found that VPS33B encodes a mitochondrial inner‐membrane protein necessary for
is necessary for clearance of endosomes containing internalized the assembly of mitochondrial respiratory chain complex III.
pattern‐recognition receptors following microbial recognition Typically, patients with GRACILE have fulminant lactic acido-
[136]. VPS33B deficiency led to enhanced signaling and expres- sis soon after birth, with nonspecific aminoaciduria and choles-
sion of inflammatory mediators which may explain the febrile tasis. Prominent siderosis is noted in the hepatocytes and
episodes in ARC. reticuloendothelial system but typical histopathologic features
Absence of platelet alpha granules can be detected on elec- of mitochondrial dysfunction such as microvesicular steatosis
tron microscopy; platelets appear hypogranular and large in are not found. GRACILE syndrome patients have no neurologi-
approximately 25% of patients with ARC [127, 128, 137]. An cal abnormalities or dysmorphic features and normal respiratory
increased risk of severe bleeding after liver and kidney biopsies chain function, including normal complex III activity.
is probably due to abnormal platelet aggregation. Severe failure Other mutations described in BSC1L cause a variety of phe-
to thrive is a prominent feature of ARC syndrome. Death occurs notypes including neurological disease, generalized mitochon-
before the first birthday in most patients, usually from sepsis, driopathies and, most recently, Bjørnstad syndrome, a form of
severe dehydration, and acidosis [135]. sensorineural hearing loss associated with pili torti [152–156].
29:  Molecular Cholestasis 357

Severity of effects of BSC1L mutation appears correlated with deficiency alone may not be sufficient to produce a CTLN2‐
increased generation of reactive oxygen species [154]. A recent like phenotype in mice. These observations are in parallel with
review showed that whilst the collection of symptoms in the variable age of onset and incomplete penetrance of CTLN2,
GRACILE is specific to the 232A>G (S78G) mutation, other where involvement of additional environmental or genetic
mutations in BCS1L can cause liver disease. The early diagno- ­factors is suspected.
sis is important as data shows that a ketogenic diet may improve
hepatic symptoms [157]. CIRH1A deficiency (North American Indian
childhood cirrhosis, NAIC)
SLC25A13 disease (citrullinemia type II, neonatal NAIC is a form of neonatal cholestasis that progresses to
intrahepatic cholestasis caused by citrin cirrhosis. It has been recognized only in the Ojibway‐Cree
­
deficiency, NICCD) ­population from northwestern Quebec [166, 167]. Children may
SLC25A13 (CITRIN) deficiency was initially described in initially present with transient neonatal jaundice which pro-
adult‐onset citrullinemia type II (CTLN2), an autosomal gresses to cirrhosis. Management of the liver disease requires
recessive disorder with variable age of presentation character- liver transplantation in childhood or early adulthood. Patients
ized by intermittent encephalopathy with hyperammonemia have elevated gGT levels and the characteristic histological
[158–160]. CTLN2 is due to decreased activity of hepatic appearance of this condition, with portal‐tract fibrosis, may
argininosuccinate synthase, which leads to hyperammonemia, resemble that of extrahepatic biliary atresia. A missense muta-
coma, and may cause death due to acute metabolic decompen- tion in CIRH1A, which encodes CIRHIN, was identified in
sation within a few years of onset [160]. More recently muta- these children [167]. It was found that CIRHIN is a ribonucleo-
tions in SLC25A13 were discovered in patients with an protein that is required for maturation of the 18S rRNA of the
inherited form of NH, now known as neonatal intrahepatic ribosome. The NAIC mutation is likely to affect the function of
cholestasis caused by CITRIN deficiency (NICCD). Although CIRHIN in ribosome biogenesis and pre‐ribosome assembly by
initially described predominantly in Japanese patients it was disrupting interaction with NOL11 [168, 169].
then recognized in patients from neighboring Korea, Taiwan,
and China. We now appreciate that this disorder occurs in all Lymphoedema‐cholestasis syndrome
ethnic groups and is important to diagnose as specialized
Lymphoedema‐cholestasis syndrome (LCS, also known as
management may not only cause clinical improvement but
Aagenaes syndrome) was originally described in Norwegian
could prevent the onset of the adult form of the disease [161,
patients [170], and refers to an autosomal recessive disorder
162]. A summary of the clinical characteristics of NICCD in
characterized by severe neonatal cholestasis with an elevated
­
75 patients reported intrahepatic cholestasis with raised gGT,
serum gGT. Although in most patients investigated liver disease
acholic stools, and poor weight gain in most. A NH syndrome
progressed to cirrhosis, some requiring liver transplantation in
was associated with hypocoagulability, hypoglycemia, distur-
childhood, more than 50% of patients survive into adulthood [171,
bances of liver synthetic function, galactosuria, and, rarely,
172]. In addition, patients develop severe lymphoedema, which
hyperammonemia. A typical but transient feature in NICCD
may be manifest at birth or appear later in childhood [173].
was found to be abnormal plasma amino acid concentrations
Norwegian families with LCS share genetic linkage to LCS1 locus
[161]. Most patients had raised citrulline, methionine, argi-
on chromosome 15 [173], however, the causal gene mutations
nine, and threonine to serine ratio values. Raised tyrosine and
have not yet been identified. Mutations in CCBE1 (collagen and
lysine values are also seen in some patients [163]. Timely
calcium‐binding EGF domain‐1) were also reported in pediatric
diagnosis of NICCD is important in view of specific dietary
and adult patients with LCS from different ethnic backgrounds
treatment (increased protein intake, initial treatment with a
[174, 175]. CCBE1 is known to be important for embryonic lym-
galactose‐free diet, and avoidance of high carbohydrate
phangiogenesis and mutations in CCBE1 were identified in cases
intake). The standard dietary management of NH could be
of lymphatic dysplasia and hydrops fetalis [176, 177].
partially toxic in this condition, precipitating deterioration of
liver disease and even requiring transplantation [161]. NICCD
is typically not severe although some cases required liver Other inherited conditions associated
transplantation [163]. The proportion of patients who pro-
gress from NICCD to CTLN2 is unknown. An additional
with cholestasis
­benefit of accurate diagnosis may be in the introduction of A number of other conditions not mentioned here but thor-
cancer surveillance and early intervention as hepatocellular oughly reviewed elsewhere may present with neonatal and
carcinoma and cholangiocarcinoma are known to complicate infantile cholestasis. The list of the disorders includes alpha‐1
CITRIN deficiency [164]. antitrypsin deficiency, Niemann–Pick disease type C, galacto-
Citrin is a liver‐specific mitochondrial aspartate‐glutamate semia, hereditary fructose deficiency, fatty acid oxidation
carrier. Investigation of the Slc25a13 knockout (Ctrn−/−) mouse defects, cystic fibrosis, and indeed many other infective or
model revealed markedly decreased activities in aspartate immune‐mediated conditions in which hepatocellular injury
transport and in the malate‐aspartate mitochondrial shuttle. leads nonspecifically to impairment of pathways involved in
Deficits in ureagenesis from ammonia and in gluconeogenesis handling of the constituents of bile [7, 178]. Each of the condi-
from lactate also could be demonstrated [165]. Ctrn−/− mice tions mentioned above has specific distinguishing features and
failed to show CTLN2‐like symptoms; this suggests that citrin baseline metabolic investigations should help in diagnosis.
358 THE LIVER:  INFANTILE CHOLESTATIC SYNDROMES WITHOUT OBVIOUS GENETIC CAUSE

Table 29.3  Syndromic forms of cholestasis


Clinical
syndromes Hepatic manifestations Extrahepatic manifestations Genetic defect
Alagille Neonatal giant‐cell hepatitis with high GGT in Cardiovascular: typically peripheral pulmonary stenosis, JAG1 or NOTCH2
syndrome some patients. Chronic cholestasis with bile duct various other congenital cardiac malformations. Monoallelic mutation,
loss in most patients with fully penetrant Renal: possibly more common in NICCD. Various forms mostly with complete
phenotype. of dysplasia, cystic kidneys, renal tubular insufficiency loss of function
Hepatocellular carcinoma may occur as early as (hematuria, proteinuria), renal tubular acidosis.
four years of age Ocular: posterior embryotoxon.
Facial: pointed chin, straight nose, deep set eyes, broad
forehead.
Skeletal: butterfly vertebrae
NISCH Neonatal cholestasis with high gGT and Dry and scaly skin. Ichthyosis most prominent on CLDN1
hepatomegaly. Liver biopsy shows extensive fibrosis abdomen and limbs. Hypotrichosis and alopecia. Mutation of both alleles
and cholangiography shows sclerosing cholangitis. Intracytoplasmic vacuoles in peripheral blood eosinophils.
Portal hypertension detected after age five years Significant improvement after liver transplantation
NICCD NH with high GGT. Liver biopsy features of Intermittent hyperammonemia, confusion, coma, SLC25A13 (CITRIN)
Citrullinemia hepatocyte microvesicular steatosis and cholestasis. cerebral edema occurs in the adult form of CITRIN Mutation of both alleles
type II Patients may develop hepatocellular carcinoma deficiency
(adult onset)
GRACILE Hepatocellular cholestasis and giant‐cell changes, Growth retardation, lactic acidosis, aminoaciduria, and BCSL1
syndrome progressing to intralobular fibrosis; prominent iron deposition in hepatocytes and reticuloendothelial Mutation of both alleles
siderosis. Bile duct paucity may be seen system. Pancreatic fibrosis. Death in infancy
ARC syndrome Neonatal cholestasis with low GGT, abnormal Neuromuscular: neurogenic arthrogryposis. The severity VPS33B or VIPAS39
bile excretion on hepatobiliary scan with may vary from unilateral talipes to multiple joint Mutation of both alleles
trimethylbromoimino‐diacetic acid (TBIDA) involvement. Dysplasia of corpus callosum in at least 20%.
scan, lipofuscin granule accumulation in Renal: renal fanconi syndrome and nephrogenic diabetes
hepatocytes and interlobular bile duct hypoplasia. insipidus. Both can vary in severity.
Patients typically have very severe failure to Hematology: deficiency of alpha granules resulting in
thrive hypogranular appearance of platelets. Abnormal
platelet aggregation.
Skin: ichthyosis, cutis laxa.
Dysmorphism: low set ears, hirsutism, proximal
insertion of thumbs
Lymphoedema‐ Transient neonatal cholestasis with high GGT. Peripheral lymphoedema LCS1 locus on
cholestasis Liver disease progresses to cirrhosis in most. chromosome 15q26.1
syndrome Transient episodes of cholestasis in adulthood CCBE1 mutations, on
both alleles, in patients
without linkage to LCS1
NAIC Transient neonatal cholestasis progressing to Not known CIRH1A
cirrhosis. High GGT. Liver transplantation is Mutation of both alleles
required in childhood/early adulthood.
Histopathological features of severe
cholangiopathy

INFANTILE CHOLESTATIC SYNDROMES girls. Concentrations of GGT activity in serum rise in EHBA.
WITHOUT OBVIOUS GENETIC CAUSE EHBA is characterized by complete fibrotic obliteration of the
lumen of all or part of the extrahepatic biliary tree within the first
Cholestatic jaundice is the commonest presentation of liver three postnatal months [183]. Luminal obliteration and fibrosis of
­disease in infancy, affecting approximately 1 in every 2500 intrahepatic bile ducts also may be seen [184]. Approximately
infants. Cholestasis has multiple etiologies with only subtle 20% of patients with biliary atresia also have at least one other
clinical differences among various diseases. Obstruction to bile major congenital anomaly. This finding suggests that genetically
flow – whether mechanical, as in extrahepatic biliary atresia, or determined defects underlie some of these cases and that the same
functional, as in PFIC  –  increases intrahepatocyte bile acid genes are involved in regulation of development of both the biliary
­concentration and causes damage which, if not corrected, leads tract and other organs [185, 186]. Polysplenia syndrome (polys-
to cell death. Jaundice, acholic stools, hepatomegaly, and the plenia, midline liver, interrupted inferior vena cava, situs inversus,
consequences of fat malabsorption are among the manifesta- preduodenal portal vein, and malrotation of the intestine) in
tions, joint or several, of cholestasis [179, 180]. particular is present in 10% of all children with biliary atresia
The most populated categories of cholestatic jaundice in the [187, 188]. Abnormal situs coupled with bile‐duct disease sug-
first months of life, historically accounting for up to 70% of gests that genes involved in shaping the laterality of thoracic
cases [179, 180], are the syndromes of extrahepatic biliary atresia and abdominal organs are involved in bile duct development.
(EHBA) and NH. In most patients with either syndrome the cause Numerous authors have suggested virus‐induced mecha-
was then unknown, although some cases were already linked to nisms in the etiopathogenesis of EHBA. The rationale for this
genetic, infectious, or other environmental factors [181, 182]. is reviewed elsewhere [189, 190].
EHBA is the most frequent form of neonatal cholestasis, occur- Interestingly, EHBA has been reported as a rare association
ring in approximately 1 in 10 000 births, and is more common in of primary ciliary dyskinesia [191]. Indeed, the inv mouse, with
29:  Molecular Cholestasis 359

a defect in the murine orthologue of INVS, has EHBA associ-   6. Sedel, F., Tourbah, A., Fontaine, B. et al. Leukoencephalopathies associated
ated with situs inversus [192, 193]. INVS mutations were iden- with inborn errors of metabolism in adults. J Inherit Metab Dis, 2008;31(3):
295–307.
tified as one of the causes of nephronophthisis, an autosomal   7. Clayton, P.T., Verrips, A., Sistermans, E., Mann, A., Mieli‐Vergani, G., and
recessive syndrome of cystic kidney disease often associated Wevers, R. Mutations in the sterol 27‐hydroxylase gene (CYP27A) cause
with congenital hepatic fibrosis and cholestasis [194, 195]. hepatitis of infancy as well as cerebrotendinous xanthomatosis. J Inherit
Many other inherited renal cystic disorders are associated with Metab Dis, 2002;25(6):501–13.
  8. Clayton, P.T., Leonard, J.V., Lawson, A.M. et al. Familial giant cell hepatitis
hepatic fibrosis, which suggests analogous pathways for the
associated with synthesis of 3 beta, 7 alpha‐dihydroxy‐ and 3 beta,7 alpha,
development of tubular structures in the liver and kidneys [196, 12 alpha‐trihydroxy‐5‐cholenoic acids. J Clin Invest, 1987;79(4):1031–8.
197]. Primary cilia act not only to move fluid: apart from  9. Ichimiya, H., Egestad, B., Nazer, H., Baginski, E.S., Clayton, P.T., and
mucociliary clearance, cilia also take part in pattern formation Sjövall, J. Bile acids and bile alcohols in a child with hepatic 3 beta‐hydroxy‐
during embryonic development, in left–right axis orientation delta 5‐C27‐steroid dehydrogenase deficiency: effects of chenodeoxycholic
acid treatment. J Lipid Res, 1991;32(5):829–41.
and in retinal photoreception [198, 197]. A particularly impor-
10. Buchmann, M.S., Kvittingen, E.A., Nazer, H. et al. Lack of 3 beta‐hydroxy‐
tant role in development of bile ducts and renal tubules is delta 5‐C27‐steroid dehydrogenase/isomerase in fibroblasts from a child
reserved for primary cilia; most of the known genes inactivated with urinary excretion of 3 beta‐hydroxy‐delta 5‐bile acids. A new inborn
in hepatorenal cystic diseases are associated with primary cilia error of metabolism. J Clin Invest, 1990;86(6):2034–7.
function [199]. 11. Jacquemin, E., Setchell, K.D., O’Connell N.C. et al. A new cause of progres-
sive intrahepatic cholestasis: 3 beta‐hydroxy‐C27‐steroid dehydrogenase/
Defects in other molecular processes, such as tight junction isomerase deficiency. J Pediatr, 1994;125(3):379–84.
protein complex formation and intercellular signaling during 12. Cheng, J.B., Jacquemin, E., Gerhardt, M. et al. Molecular genetics of 3beta‐
bile‐duct morphogenesis, may also be among contributors to hydroxy‐delta5‐C27‐steroid oxidoreductase deficiency in 16 patients with
EHBA. Deficiency of the proteins involved in these processes loss of bile acid synthesis and liver disease. J Clin Endocrinol Metab,
causes specific forms of neonatal sclerosing cholangitis (absence 2003;88(4):1833–41.
13. Setchell, K.D., Suchy, F.J., Welsh, M.B., Zimmer‐Nechemias, L., Heubi, J.,
of claudin‐1 in NISCH syndrome, [138]) and paucity of inter- and Balistreri W.F. Delta 4–3‐oxosteroid 5 beta‐reductase deficiency
lobular bile ducts (haploinsufficiency for JAG1/NOTCH2 in described in identical twins with neonatal hepatitis. A new inborn error in
Alagille syndrome [113, 122]. Assessment of synthesis of dys- bile acid synthesis. J Clin Invest, 1988;82(6):2148–57.
functional forms of these proteins in patients with EHBA 14. Setchell, K.D., Schwarz, M., O’Connell N.C. et al. Identification of a new
inborn error in bile acid synthesis: mutation of the oxysterol 7alpha‐hydrox-
remains to be undertaken. A description of villin deficiency
ylase gene causes severe neonatal liver disease. J Clin Invest, 1998;102(9):
associated with EHBA‐like disease [200] has not been followed 1690–703.
by a report of a genetic correlate. 15. Tsaousidou, M.K., Ouahchi, K., Warner, T.T. et  al. Sequence alterations
The proportion of patients with a final diagnosis of NH is within CYP7B1 implicate defective cholesterol homeostasis in motor‐neu-
getting smaller, as a result of the identification of the diseases ron degeneration. Am J Hum Genet, 2008;82(2):510–5.
17. Ferdinandusse, S., Rusch, H., van Lint, A.E., Dacremont, G., Wanders, R.J.,
described above. NH is evidently a syndrome, or clinicopatho-
and Vreken, P. Stereochemistry of the peroxisomal branched‐chain fatty acid
logic diagnosis of last resort. NH is diagnosed if cholestasis alpha‐ and beta‐oxidation systems in patients suffering from different per-
cannot be assigned to infection, genetic/metabolic defect, or
­ oxisomal disorders. J Lipid Res, 2002;43(3):438–44.
mechanical impediment to bile flow and if microscopy of liver 18. Van Veldhoven, P.P., Meyhi, E., Squires, R.H. et al. Fibroblast studies docu-
finds widespread giant‐cell transformation of hepatocytes. Giant‐ menting a case of peroxisomal 2‐methylacyl‐CoA racemase deficiency: pos-
sible link between racemase deficiency and malabsorption and vitamin K
cell change of hepatocytes is common, however, to many choles- deficiency. J Clin Invest Eur J Clin Invest, 2001;31(8):714–22.
tatic disorders in neonates, including EHBA. Most instances of 19. Setchell, K.D., Heubi, J.E., O’Connell N.C., Hofmann, A.F., and Levine J.A.
NH resolve without sequelae; they may represent constitutive Absence of bile acid conjugation: a new inborn error of bile acid metabolism.
dysfunction of bile‐handling mechanisms unmasked by perinatal Hepatol, 1996;24(4):371A.
stress, a hypothesis that awaits experimental assessment. 20. Carlton, V.E., Harris, B.Z., Puffenberger, E.G., Batta, A.K., Knisely, A.S.,
Robinson, D.L. et al. Complex inheritance of familial hypercholanemia with
Previously, 20% of cases of idiopathic NH that are progressive, associated mutations in TJP2 and BAAT. Nat Genet, 2003;34(1):91–6.
appeared familial, and had a poor prognosis [184]. This group in 21. Tazawa, Y., Yamada, M., Nakagawa, M., Konno, T., and Tada, K. Bile acid
particular has been eroded by the discoveries described above, profiles in siblings with progressive intrahepatic cholestasis: absence of bil-
combined with the widespread use of genetic testing [201–203]. iary chenodeoxycholate. J Pediatr Gastroenterol Nutr, 1985;4(1):32–7.
22. Bull, L.N., Carlton, V.E., Stricker, N.L. et  al. Genetic and morphological
findings in progressive familial intrahepatic cholestasis (Byler disease
[PFIC‐1] and Byler syndrome): evidence for heterogeneity. Hepatol, 1997;
26(1):155–64.
REFERENCES 23. Houwen, R.H., Baharloo, S., Blankenship, K. et al. Genome screening by
searching for shared segments: mapping a gene for benign recurrent intrahe-
1. Heubi, J.E., Setchell, K.D., and Bove, K.E. Inborn errors of bile acid metabo- patic cholestasis. Nat Genet, 1994;8(4):380–6.
lism. Semin Liver Dis, 2007;27(3):282–94. 24. van Mil, S.W., van der Woerd, W.L., van der Brugge, G. et al. Benign recur-
2. Rossi, M., Vajro, P., Iorio, R. et al. Characterization of liver involvement in rent intrahepatic cholestasis type 2 is caused by mutations in ABCB11.
defects of cholesterol biosynthesis: long‐term follow‐up and review. Am Gastroenterol, 2004;127(2):379–84.
J Med Genet A, 2005;132A(2):144–51. 25. van Ooteghem, N.A., Klomp, L.W., van Berge‐Henegouwen, G.P., and
3. Steinberg, S.J., Dodt, G., Raymond, G.V., Braverman, N.E., Moser, A.B., and Houwen R.H. Benign recurrent intrahepatic cholestasis progressing to pro-
Moser H.W. Peroxisome biogenesis disorders. Biochim Biophys Acta, gressive familial intrahepatic cholestasis: low GGT cholestasis is a clinical
2006;1763(12):1733–48. continuum. J Hepatol, 2002;36(3):439–43.
4. Russell, D.W. The enzymes, regulation, and genetics of bile acid synthesis. 26. Takahashi, A., Hasegawa, M., Sumazaki, R. et al. Gradual improvement of
Annu Rev Biochem, 2003;72:137–74. liver function after administration of ursodeoxycholic acid in an infant with
5. Hofmann, A.F. and Hagey, L.R. Bile acids: chemistry, pathochemistry, biology, a novel ABCB11 gene mutation with phenotypic continuum between BRIC2
pathobiology, and therapeutics. Cell Mol Life Sci, 2008;65(16):2461–83. and PFIC2. Eur J Gastroenterol Hepatol, 2007;19(11):942–6.
360 THE LIVER:  REFERENCES

27. Meier, Y., Zodan, T., Lang, C. et al. Increased susceptibility for intrahepatic of the bile salt export pump (Bsep/Abcb11) correlate with severity of chole-
cholestasis of pregnancy and contraceptive‐induced cholestasis in carriers of static diseases. Am J Physiol Cell Physiol, 2007;293(5):C1709–16.
the 1331T>C polymorphism in the bile salt export pump. World J Gastroenterol, 54. Dixon, P.H., Wadsworth, C.A., Chambers, J. et al. A comprehensive analysis
2008;14(1):38–45. of common genetic variation around six candidate loci for intrahepatic chol-
28. Müllenbach, R., Bennett, A., Tetlow, N. et al. ATP8B1 mutations in British estasis of pregnancy. Am J Gastroenterol, 2014;109(1):76–84.
cases with intrahepatic cholestasis of pregnancy. 1. Gut, 2005;54(6):829–34. 55. Dixon, P.H., Sambrotta, M., Chambers, J. et al. An expanded role for hete-
29. Lammert, F., Marschall, H.U., Glantz, A., and Matern, S. Intrahepatic chol- rozygous mutations of ABCB4, ABCB11, ATP8B1, ABCC2 and TJP2 in
estasis of pregnancy: molecular pathogenesis, diagnosis and management. J intrahepatic cholestasis of pregnancy. Sci Rep, 2017;7(1):11823.
Hepatol, 2000;33(6):1012–21. 56. Vallejo, M., Briz, O., Serrano, M.A., Monte, M.J., and Marin J.J. Potential
30. Milkiewicz, P., Gallagher, R., Chambers, J., Eggington, E., Weaver, J., and Elias, role of trans‐inhibition of the bile salt export pump by progesterone metabo-
E. Obstetric cholestasis with elevated gamma glutamyl transpeptidase: incidence, lites in the etiopathogenesis of intrahepatic cholestasis of pregnancy. J
presentation and treatment. J Gastroenterol Hepatol, 2003;18(11):1283–6. Hepatol, 2006;44(6):1150–7.
31. Clayton, R.J., Iber, F.L., Ruebner, B.H., and McKusick V.A. Byler disease. 57. Iwanaga, T., Nakakariya, M., Yabuuchi, H., Maeda, T., and Tamai, I.
Fatal familial intrahepatic cholestasis in an Amish kindred. Am J Dis Child, Involvement of bile salt export pump in flutamide‐induced cholestatic hepa-
1969;117(1):112–24. titis. Biol Pharm Bull, 2007;30(4):739–44.
32. Bull, L.N., van Eijk, M.J., Pawlikowska, L. et al. A gene encoding a P‐type 58. Schafmayer, C., Tepel, J., Franke, A. et al. Investigation of the Lith1 candi-
ATPase mutated in two forms of hereditary cholestasis. Nat Genet, date genes ABCB11 and LXRA in human gallstone disease. Hepatol,
1998;18(3):219–24. 2006;44(3):650–7.
33. Eppens, E.F., van Mil, S.W., de Vree, J.M. et al. FIC1, the protein affected in 59. Wang, R., Lam, P., Liu, L. et al. Severe cholestasis induced by cholic acid
two forms of hereditary cholestasis, is localized in the cholangiocyte and the feeding in knockout mice of sister of P‐glycoprotein. Hepatol, 2003;
canalicular membrane of the hepatocyte. J Hepatol, 2001;35(4):436–43. 38(6):1489–99.
34. Ujhazy, P., Ortiz, D., Misra, S. et al. Familial intrahepatic cholestasis 1: stud- 60. Lam, P., Wang, R., and Ling, V. Bile acid transport in sister of P‐glycoprotein
ies of localization and function. Hepatol, 2001;34(4 Pt 1):768–75. (ABCB11) knockout mice. Biochemistry, 2005;44(37):12598–605.
35. van Mil, S.W., Houwen, R.H., and Klomp L.W. Genetics of familial intrahe- 61. Paulusma, C.C., Bosma, P.J., Zaman, G.J. et al. Congenital jaundice in rats
patic cholestasis syndromes. J Med Genet, 2005;42(6):449–63. with a mutation in a multidrug resistance‐associated protein gene. Science,
36. Paulusma, C.C., Folmer, D.E., Ho‐Mok, K.S. et  al. ATP8B1 requires an 1996;271(5252):1126–8.
accessory protein for endoplasmic reticulum exit and plasma membrane lipid 62. Ito, K., Suzuki, H., Hirohashi, T., Kume, K., Shimizu, T., and Sugiyama, Y.
flippase activity. Hepatol, 2008;47(1):268–78. Molecular cloning of canalicular multispecific organic anion transporter
37. Frankenberg, T., Miloh, T., Chen, F.Y. et al. The membrane protein ATPase defective in EHBR. Am J Physiol, 1997;272(1 Pt 1):G16–22.
class I type 8B member 1 signals through protein kinase C zeta to activate the 63. Paulusma, C.C., Kool, M., Bosma, P.J. et al. A mutation in the human cana-
farnesoid X receptor. Hepatol, 2008;44(1):195–204. licular multispecific organic anion transporter gene causes the Dubin–
39. Demeilliers, C., Jacquemin, E., Barbu, V. et al. Altered hepatobiliary gene Johnson syndrome. Hepatol, 1997;25(6):1539–42.
expressions in PFIC1: ATP8B1 gene defect is associated with CFTR down- 64. Ito, K., Suzuki, H., Hirohashi, T., Kume, K., Shimizu, T., and Sugiyama, Y.
regulation. Hepatol, 2006;43(5):1125–34. Functional analysis of a canalicular multispecific organic anion transporter
40. Knisely, A.S., Agostini, R.M., Zitelli, B.J., Kocoshis, S.A., and Boyle J.T. cloned from rat liver. J Biol Chem, 1998;273(3):1684–8.
Byler’s syndrome. Arch Dis Child, 1997;77(3):276–7. 65. Nies, A.T. and Keppler, D. The apical conjugate efflux pump ABCC2
41. Oshima, T., Ikeda, K., and Takasaka, T. Sensorineural hearing loss associated (MRP2). Pflugers Arch, 2007;453(5):643–59
with Byler disease. Tohoku J Exp Med, 1999;187(1):83–8. 66. Dubin, I.N. and Johnson F.B. Chronic idiopathic jaundice with unidentified
42. Winklhofer‐Roob, B.M., Shmerling, D.H., Solèr, R., and Briner, J. Progressive pigment in liver cells; a new clinicopathologic entity with a report of 12
idiopathic cholestasis presenting with profuse watery diarrhoea and recurrent cases. Medicine (Baltimore), 1954;33(3):155–97.
infections (Byler’s disease). Acta Paediatr, 1992;81(8):637–40. 67. Lee, J.H., Chen, H.L., Chen, H.L., Ni, Y.H., Hsu, H.Y., and Chang M.H.
43. Knisely A.S. Progressive familial intrahepatic cholestasis: a personal per- Neonatal Dubin–Johnson syndrome: long‐term follow‐up and MRP2 muta-
spective. Pediatr Dev Pathol, 2000;3(2),113–25. tions study. Pediatr Res, 2006;59(4 Pt 1):584–9.
44. Walkowiak, J., Jankowska, I., Pawlowska, J., Bull, L., Herzig, K.H., and 68. Donner, M.G. and Keppler, D. Up‐regulation of basolateral multidrug resist-
Socha, J. Normal pancreatic secretion in children with progressive familial ance protein 3 (Mrp3) in cholestatic rat liver. Hepatol, 2001;34(2):351–9.
intrahepatic cholestasis type 1. Scand J Gastroenterol, 2006;41(12):1480–3. 69. Chu, X.Y., Strauss, J.R., Mariano, M.A. et al. Characterization of mice lack-
45. Klomp, L.W., Vargas, J.C., van Mil, S.W. et al. Characterization of mutations in ing the multidrug resistance protein MRP2 (ABCC2). J Pharmakos Exp
ATP8B1 associated with hereditary cholestasis. Hepatol, 2004;40(1):27–38. Ther, 2006;317(2):579–89.
46. Strautnieks, S.S., Bull, L.N., Knisely et al. A gene encoding a liver‐specific 70. Kikuchi, S., Hata, M., Fukumoto, K. et al. Radixin deficiency causes conju-
ABC transporter is mutated in progressive familial intrahepatic cholestasis. gated hyperbilirubinemia with loss of Mrp2 from bile canalicular mem-
Nat Genet, 1998;20(3):233–8. branes. Nat Genet, 2002;31(3):320–5.
47. Stieger, B., Meier, Y., and Meier, P.J. The bile salt export pump. Pflugers 71. Wolpert, E., Pascasio, F.M., Wolkoff, A.W., and Arias I.M. Abnormal sulfo-
Arch, 2007;453(5):611–20. bromophthalein metabolism in Rotor’s syndrome and obligate heterozy-
48. Noe, J., Kullak‐Ublick, G.A., Jochum, W. et  al. Impaired expression and gotes. N Engl J Med, 1977;296(19):1099–101.
function of the bile salt export pump due to three novel ABCB11 mutations 72. Hrebícek, M., Jirásek, T., Hartmannová, H. et al. Rotor‐type hyperbilirubi-
in intrahepatic cholestasis. J Hepatol, 2005;43(3):536–43. naemia has no defect in the canalicular bilirubin export pump. Liver Int,
49. Lam, C.W., Cheung, K.M., Tsui, M.S., Yan, M.S., Lee, C.Y., and Tong, S.F. 2007;27(4):485–91.
A patient with novel ABCB11 gene mutations with phenotypic transition 73. van de Steeg, E., Stránecký, V., Hartmannová, H. et al. Complete OATP1B1
between BRIC2 and PFIC2. J Hepatol, 2006;44(1):240–2. and OATP1B3 deficiency causes human Rotor syndrome by interrupting con-
50. Strautnieks, S.S., Byrne, J.A., Pawlikowska, L. et  al. Severe bile salt export jugated bilirubin reuptake into the liver. J Clin Invest, 2012;122(2):519–28.
pump deficiency: 82 different ABCB11 mutations in 109 families. Gastroenterol, 74. Deleuze, J.F., Jacquemin, E., Dubuisson, C. et al. Defect of multidrug‐resist-
2008;134(4):1203–14. ance 3 gene expression in a subtype of progressive familial intrahepatic chol-
51. Chen, H.L., Chang, P.S., Hsu, H.C. et  al. FIC1 and BSEP defects in estasis. Hepatol, 1996;23(4):904–8.
Taiwanese patients with chronic intrahepatic cholestasis with low gamma‐ 75. Oude Elferink, R.P. and Paulusma, C.C. Function and pathophysiological impor-
glutamyltranspeptidase levels. J Pediatr, 2002;140(1):119–24. tance of ABCB4 (MDR3 P‐glycoprotein). Pflugers Arch, 2007;453(5):601–10.
52. Kagawa, T., Watanabe, N., Mochizuki, K. et  al. Phenotypic differences in 76. Lamireau, T., Bouchard, G., Yousef, I.M. et  al. Dietary lecithin protects
PFIC2 and BRIC2 correlate with protein stability of mutant Bsep and against cholestatic liver disease in cholic acid‐fed Abcb4‐deficient mice.
impaired taurocholate secretion in MDCK II cells. Am J Physiol Gastrointest Pediatr Res, 2007;61(2):185–90.
Liver Physiol, 2008;294(1):G58–67. 77. Baghdasaryan, A., Fickert, P., Fuchsbichler, A. et al. Role of hepatic phos-
53. Lam, P., Pearson, C.L., Soroka, C.J., Xu, S., Mennone, A., and Boyer, J.L. pholipids in development of liver injury in Mdr2 (Abcb4) knockout mice.
Levels of plasma membrane expression in progressive and benign mutations Liver Int, 2008;28(7):948–58.
29:  Molecular Cholestasis 361

  78. Smit, J.J., Schinkel, A.H., Oude Elferink, R.P. et al. Homozygous disrup- 102. Thoeni, C.E., Vogel, G.F., Tancevski, I. et al. Microvillus inclusion disease:
tion of the murine mdr2 P‐glycoprotein gene leads to a complete absence of loss of Myosin vb disrupts intracellular traffic and cell polarity. Traffic,
phospholipid from bile and to liver disease. Cell, 1993;75(3):451–62. 2014;15(1):22–42.
  79. Ziol, M., Barbu, V., Rosmorduc, O. et al. ABCB4 heterozygous gene muta- 103. Girard, M., Lacaille, F., Verkarre, V. et  al. MYO5B and bile salt export
tions associated with fibrosing cholestatic liver disease in adults. pump contribute to cholestatic liver disorder in microvillous inclusion dis-
Gastroenterol, 2008;135(1):131–41. ease. Hepatol, 2014;60(1):301–10.
  80. Gotthardt, D., Runz, H., Keitel, V. et al. A mutation in the canalicular phos- 104. Grammatikopoulos, T., Sambrotta, M., Strautnieks, S. et al. Mutations in
pholipid transporter gene, ABCB4, is associated with cholestasis, ductope- DCDC2 (doublecortin domain containing protein 2) in neonatal sclerosing
nia, and cirrhosis in adults. Hepatol, 2008;48(4):1157–66. cholangitis. J Hepatol, 2016;65(6):1179–1187.
  81. Jacquemin, E. Role of multidrug resistance 3 deficiency in pediatric and 105. Girard, M., Bizet, A.A., Lachaux, A. et al. DCDC2 mutations cause neona-
adult liver disease: one gene for three diseases. Semin Liver Dis, tal sclerosing cholangitis. Hum Mutat, 2016;37(10):1025–9.
2001;21(4):551–62 106. Carrion‐Castillo, A., Franke, B., and Fisher S.E. Molecular genetics of dys-
  82. de Vree, J.M., Jacquemin, E., Sturm, E. et al. Mutations in the MDR3 gene lexia: an overview. Dyslexia, 2013;19(4):214–40.
cause progressive familial intrahepatic cholestasis. Proc Natl Acad Sci 107. Schueler, M., Braun, D.A., Chandrasekar, G. et  al. DCDC2 mutations
USA, 1998;95(1):282–7. cause a renal‐hepatic ciliopathy by disrupting Wnt signaling. Am J Hum
  83. Dixon, P.H., Weerasekera, N., Linton, K.J. et al. Heterozygous MDR3 mis- Genet, 2015;96(1):81–92.
sense mutation associated with intrahepatic cholestasis of pregnancy: evi- 107. Shueler, M., Braun, D.A., Chandrasekar, G. et al. DCDC2 mutations cause
dence for a defect in protein trafficking. Hum Mol Genet, a renal‐hepatic ciliopathy by disrupting Wnt signaling. Am J Hum Genet,
2000;9(8):1209–17. 2015;96(1):82–92.
  84. Pauli‐Magnus, C., Lang, T., Meier, Y. et al. Sequence analysis of bile salt 108. Alagille, D., Odievre, M., Gautier, M., and Demerges J.P. Hepatic ductular
export pump (ABCB11) and multidrug resistance p‐glycoprotein 3 hypoplasia associated with characteristic facies, vertebral malformations,
(ABCB4, MDR3) in patients with intrahepatic cholestasis of pregnancy. retarded physical, mental and sexual development, and cardiac murmur. J
Pharmacogenet, 2004;14(2):91–102. Pediatr 1975;86:63–71.
  85. Wasmuth, H.E., Glantz, A., Keppeler, H. et al. Intrahepatic cholestasis of 109. Emerick, K.M., Rand, E.B., Goldmuntz, E., Krantz, I.D., Spinner, N.B.,
pregnancy: the severe form is associated with common variants of the hepa- and Piccoli D.A. Features of Alagille syndrome in 92 patients: frequency
tobiliary phospholipid transporter ABCB4 gene. Gut, 2007;56(2):265–70. and relation to prognosis. Hepatol, 1999;29(3):822–9.
  87. Rosmorduc, O. and Poupon, R. Low phospholipid associated cholelithiasis: 110. Chiaretti, A., Zampino, G., Botto, L., and Polidori, G. Alagille syndrome
association with mutation in the MDR3/ABCB4 gene. Orphanet J Rare and hepatocarcinoma: a case report. Acta Paediatr, 1992;81(11):937.
Dis, 2007;2:29. 111. Békássy, A.N., Garwicz, S., Wiebe, T., Homestand, I., and Jensen O.A.
 88. Rosmorduc, O., Hermelin, B., Boelle, P.Y., Parc, R., Taboury, J., and Hepatocellular carcinoma associated with arteriohepatic dysplasia in a 4‐
Poupon, R. ABCB4 gene mutation‐associated cholelithiasis in adults. year‐old girl. Med Pediatr Oncol, 1992;20(1):78–83.
Gastroenterol, 2003;125(2):452–9. 112. Kamath, B.M., Bason, L., Piccoli, D.A., Krantz, I.D., and Spinner N.B.
  89. Kano, M., Shoda, J., Sumazaki, R., Oda, K., Nimura, Y., and Tanaka, N. Consequences of JAG1 mutations. J Med Genet, 2003;40(12):891–5.
Mutations identified in the human multidrug resistance P‐glycoprotein 3 113. Oda, T., Elkahloun, A.G., Pike, B.L. et al. Mutations in the human Jagged1
(ABCB4) gene in patients with primary hepatolithiasis. Hepatol Res, gene are responsible for Alagille syndrome. Nat Genet,
2004;29(3):160–6. 1997;16(3):235–42.
  90. Shoda, J., Oda, K., Suzuki, H. et  al. Etiologic significance of defects in 114. Warthen, D.M., Moore, E.C., Kamath, B.M. et al. Jagged1 (JAG1) muta-
cholesterol, phospholipid, and bile acid metabolism in the liver of patients tions in Alagille syndrome: increasing the mutation detection rate. Hum
with intrahepatic calculi. Hepatol, 2001;33(5):1194–205. Mutat, 2006;27(5):436–43.
  91. Lang, C., Meier, Y., Stieger, B. et al. Mutations and polymorphisms in the 115. Lanford, P.J., Lan, Y., Jiang, R. et al. Notch signalling pathway mediates hair
bile salt export pump and the multidrug resistance protein 3 associated with cell development in mammalian cochlea. Nat Genet, 1999;21(3):289–92.
drug‐induced liver injury. Pharmacogenet Genomics, 2007;17(1):47–60. 116. Davies, J.A. and Fisher C.E. Genes and proteins in renal development. Exp
  92. Bramow, S., Ott, P., Thomsen Nielsen, F., Bangert, K., Tygstrup, N., and Nephrol, 2002;10(2):102–13.
Dalhoff, K. Cholestasis and regulation of genes related to drug metabolism 117. Hu, Q.D., Cui, X.Y., Ng, Y.K., and Xiao, Z.C. Axoglial interaction via the
and biliary transport in rat liver following treatment with cyclosporine A notch receptor in oligodendrocyte differentiation. Ann Acad Med Singapore,
and sirolimus (rapamycin). Pharmakon Toxicol, 2001;89(3):133–9. 2004;33(5):581–8.
  93. Morton, D.H., Salen, G., Batta, A.K. et al. Abnormal hepatic sinusoidal bile 118. Bray, S.J. Notch signalling: a simple pathway becomes complex. Nat Rev
acid transport in an Amish kindred is not linked to FIC1 and is improved by Mol Cell Biol, 2006;7(9):678–89.
ursodiol. Gastroenterol, 2000;119(1):188–95. 119. Ehebauer, M., Hayward, P., and Arias A.M. Notch, a universal arbiter of
  94. Sambrotta, M., Strautnieks, S., Papouli, E. et al. Mutations in TJP2 cause cell fate decisions. Science, 2006;314(5804):1414–5.
progressive cholestatic liver disease. Nat Genet, 2014;46(4):326–8 120. McCright B., Gao, X., Shen, L. et al. Defects in development of the kidney,
  95. Zhou, S., Hertel, P.M., Finegold, M.J. et al. Hepatocellular carcinoma associ- heart and eye vasculature in mice homozygous for a hypermorphic Notch2
ated with tight‐junction protein 2 deficiency. Hepatol, 2015;62(6):1914–6. mutation. Development, 2001;128(4):491–502.
  96. Knowles, B.C., Roland, J.T., Krishnan, M. et  al. Myosin Vb uncoupling 121. McCright B., Lozier, J., and Gridley, T. A mouse model of Alagille syn-
from RAB8A and RAB11A elicits microvillus inclusion disease. J Clin drome: Notch2 as a genetic modifier of Jag1 haploinsufficiency.
Invest, 2014;124(7):2947–62. Development, 2002;129(4):1075–82.
  97. Lapierre, L.A., Kumar, R., Hales, C.M. et al. Myosin vb is associated with 122. McDaniell R., Warthen, D.M., Sanchez‐Lara, P.A. et al. NOTCH2 muta-
plasma membrane recycling systems. Mol Biol Cell, 2001;12(6):1843–57. tions cause Alagille syndrome, a heterogeneous disorder of the notch sign-
  98. Wakabayashi, Y., Dutt, P., Lippincott‐Schwartz, J. et al. Rab11a and myosin aling pathway. Am J Hum Genet, 2006;79(1):169–73.
Vb are required for bile canalicular formation in WIF‐B9 cells. Proc Natl 123. Lozier, J., McCright B., and Gridley, T. Notch signaling regulates bile duct
Acad Sci USA, 2005;102(42):15087–92. morphogenesis in mice. PLoS ONE, 2008;3(3):e1851.
  99. Wakabayashi, Y., Lippincott‐Schwartz, J., and Arias I.M. Intracellular traf- 124. Gissen, P., Johnson, C.A., Morgan, N.V. et al. Mutations in VPS33B., encoding
ficking of bile salt export pump (ABCB11) in polarized hepatic cells: con- a regulator of SNARE‐dependent membrane fusion, cause arthrogryposis‐renal
stitutive cycling between the canalicular membrane and rab11‐positive dysfunction‐cholestasis (ARC) syndrome. Nat Genet, 2004;36(4):400–4.
endosomes. Mol Biol Cell, 2004;15(7):3485–96. 125. Cullinane, A.R., Straatman‐Iwanowska, A., Zaucker, A. et al. Mutations in
100. Muller, T., Hess, M.W., Schiefermeier, N. et al. MYO5B mutations cause VIPAR cause an arthrogryposis, renal dysfunction and cholestasis syn-
microvillus inclusion disease and disrupt epithelial cell polarity. Nat Genet, drome phenotype with defects in epithelial polarization. Nat Genet,
2008;40(10):1163–5. 2010;42(4):303–12.
101. Ruemmele, F.M., Muller, T., Schiefermeier, N. et al. Loss‐of‐function of 126. Banushi, B., Forneris, F., Straatman‐Iwanowska, A. et  al. Regulation of
MYO5B is the main cause of microvillus inclusion disease: 15 novel muta- post‐Golgi LH3 trafficking is essential for collagen homeostasis. Nat
tions and a CaCo‐2 RNAi cell model. Hum Mutat, 2010;31(5):544–51. Commun, 2016;7:12111.
362 THE LIVER:  REFERENCES

127. Lo, B., Li, L., Gissen, P. et al. Requirement of VPS33B., a member of the 151. Fellman, V., Lemmelä, S., Sajantila, A., Pihko, H., and Järvelä, I. Screening
Sec1/Munc18 protein family, in megakaryocyte and platelet alpha‐granule of BCS1L mutations in severe neonatal disorders suspicious for mitochon-
biogenesis. Blood, 2005;106(13):4159–66. drial cause. J Hum Genet, 2008;53(6):554–8.
128. Bem, D., Smith, H., Banushi, B. et al. VPS33B regulates protein sorting 152. Baker, R.A., Priestley, J.R.C., Wilstermann, A.M., Reese, K.J., and Mark
into and maturation of α‐granule progenitor organelles in mouse megakar- P.R. Clinical spectrum of BCS1L Monopathies and their underlying struc-
yocytes. Blood, 2015;126(2):133–43. tural relationships. Am J Med Genet A, 2018.
129. Hershkovitz D., Mandel, H., Ishida‐Yamamoto, A. et al. Defective lamellar 153. Fernandez‐Vizarra, E., Bugiani, M., Goffrini, P. et al. Impaired complex III
granule secretion in arthrogryposis, renal dysfunction, and cholestasis syn- assembly associated with BCS1L gene mutations in isolated mitochondrial
drome caused by a mutation in VPS33B. Arch Dermatol, 2008;144(3): encephalopathy. Hum Mol Genet, 2007;16(10):1241–52.
334–40. 154. Hinson, J.T., Fantin, V.R., Schönberger, J. et al. Missense mutations in the
130. Rogerson, C. and Gissen, P. VPS33B and VIPAR are essential for epider- BCS1L gene as a cause of the Björnstad syndrome. N Engl J Med, 2007;
mal lamellar body biogenesis and function. Biochim Biophys Acta Mol 356(8):809–19.
Basis Dis, 2018;1864(5):1609–21. 155. De Meirleir, L., Seneca, S., Damis, E. et al. Clinical and diagnostic charac-
131. Horslen, S.P., Quarrell, O.W., and Tanner M.S. Liver histology in the teristics of complex III deficiency due to mutations in the BCS1L gene. Am
arthrogryposis multiplex congenita, renal dysfunction, and cholestasis J Med Genet A, 2003;121A(2):126–31.
(ARC) syndrome: report of three new cases and review. J Med Genet, 156. Selvaag, E. Pili torti and sensorineural hearing loss. A follow‐up of
1994;31(1):62–4. Bjørnstad’s original patients and a review of the literature. Eur J Dermatol,
132. Bull, L.N., Mahmoodi, V., Baker, A.J. et  al. VPS33B mutation with ich- 2000;10(2):91–7.
thyosis, cholestasis, and renal dysfunction but without arthrogryposis: 157. Purhonen J,, Rajendran, J., Mörgelin, M. et al. Ketogenic diet attenuates
incomplete ARC syndrome phenotype. J Pediatr, 2006;148(2):269–71. hepatopathy in mouse model of respiratory chain complex III deficiency
133. Hanley, J., Dhar, D.K., Mazzacuva, F. et al. Vps33b is crucial for structural caused by a Bcs1l mutation. Sci Rep,. 2017;7(1):957.
and functional hepatocyte polarity. J Hepatol, 2017;66(5):1001–11. 158. Sase, M., Kobayashi, K., Imamura, Y. et al. Level of translatable messenger
134. Wang C., Cheng, Y., Zhang, X. et  al. Vacuolar protein sorting 33B is a RNA coding for argininosuccinate synthetase in the liver of the patients
tumor suppressor in hepatocarcinogenesis. Hepatol, 2018;68(6):2239–53. with quantitative‐type citrullinemia. Hum Genet, 1985;69(2):130–4.
135. Gissen, P., Tee, L., Johnson, C.A. et al. Clinical and molecular genetic fea- 159. Saheki, T., Nakano, K., Kobayashi, K. et al. Analysis of the enzyme abnor-
tures of ARC syndrome. Hum Genet, 2006;120(3):396–409. mality in eight cases of neonatal and infantile citrullinaemia in Japan.
136. Akbar, M.A., Mandraju, R., Tracy, C., Hu, W., Pasare, C., and Krämer, H. J Inherit Metab Dis, 1985;8(3):155–6.
ARC syndrome‐linked Vps33B protein is required for inflammatory endo- 160. Kobayashi, K., Sinasac, D.S., Iijima, M. et al. The gene mutated in adult‐
somal maturation and signal termination. Immunity, 2016;45(2):267–79. onset type II citrullinaemia encodes a putative mitochondrial carrier pro-
137. Deal, J.E., Barratt, T.M., and Dillon, M.J. Fanconi syndrome, ichthyosis, tein. Nat Genet, 1999;22(2):159–63.
dysmorphism, jaundice and diarrhoea‐‐a new syndrome. Pediatr Nephrol, 161. Ohura, T., Kobayashi, K., Tazawa, Y. et al. Clinical pictures of 75 patients
1990;4(4):308–13. with neonatal intrahepatic cholestasis caused by citrin deficiency (NICCD).
138. Hadj‐Rabia, S., Baala, L., Vabres, P. et  al. Claudin‐1 gene mutations in J Inherit Metab Dis, 2007;30(2):139–44.
neonatal sclerosing cholangitis associated with ichthyosis: a tight junction 162. Tabata, A., Sheng, J.S., Ushikai, M. et al. Identification of 13 novel muta-
disease. Gastroenterol, 2004;127(5):1386–90. tions including a retrotransposal insertion in SLC25A13 gene and fre-
139. Feldmeyer, L., Huber, M., Fellmann, F., Beckmann, J.S., Frenk, E., and quency of 30 mutations found in patients with citrin deficiency. J Hum
Hohl, D. Confirmation of the origin of NISCH syndrome. Hum Mutat, Genet, 2008;53(6):534–45.
2006;27(5):408–10. 163. Kobayashi, K., Iijima, M., Ushikai, M., Ikeda, S., and Saheki, T. Citrin
140. Szepetowski, S., Lacoste, C., Mallet, S., Roquelaure, B., Badens, C., and deficiency. J Jpn Pediatr Soc, 2006;110:1047–59.
Fabre, A. NISCH syndrome, a rare cause of neonatal cholestasis: a case 164. Soeda, J., Yazaki, M., Nakata, T. et al. Primary liver carcinoma exhibiting
report. Arch Pediatr, 2017;24(12):1228–34. dual hepatocellular‐biliary epithelial differentiations associated with citrin
141. Baala, L., Hadj‐Rabia, S., Hamel‐Teillac, D. et al. Homozygosity mapping deficiency: a case report. J Clin Gastroenterol, 2008;42(7):855–60.
of a locus for a novel syndromic ichthyosis to chromosome 3q27‐q28. 165. Sinasac, D.S., Moriyama, M., Jalil, M.A. et  al. Slc25a13‐knockout mice
J Invest Dermatol, 2002;119(1):70–6. harbor metabolic deficits but fail to display hallmarks of adult‐onset type II
142. Mazzon, E., Puzzolo, D., Caputi, A.P., and Cuzzocrea, S. Role of IL‐10 in citrullinemia. Mol Cell Biol, 2004;24(2):527–36.
hepatocyte tight junction alteration in mouse model of experimental colitis. 166. Weber, A.M., Tuchweber, B., Yousef, I. et al. Severe familial cholestasis in
Mol Med, 2002;8(7):353–66. North American Indian children: a clinical model of microfilament dys-
143. Maly, I.P. and Landmann, L. Bile duct ligation in the rat causes upregula- function? Gastroenterol, 1981;81(4):653–62.
tion of ZO‐2 and decreased colocalization of claudins with ZO‐1 and occlu- 167. Chagnon, P., Michaud, J., Mitchell, G. et al. A missense mutation (R565W)
din. Histochem Cell Biol, 2008;129(3):289–99. in cirhin (FLJ14728) in North American Indian childhood cirrhosis. Am
144. Fickert, P., Fuchsbichler, A., Marschall, H.U. et al. Lithocholic acid feeding J Hum Genet, 2002;71(6):1443–9.
induces segmental bile duct obstruction and destructive cholangitis in mice. 168. Freed, E.F. and Baserga, S.J. The C‐terminus of Utp4, mutated in childhood
Am J Pathol, 2006;168(2):410–22. cirrhosis, is essential for ribosome biogenesis. Nucleic Acids Res,
145. Anderson, J.M. Leaky junctions and cholestasis: a tight correlation. 2010;38(14):4798–806.
Gastroenterol, 1996;110(5):1662–5. 169. Freed, E.F., Prieto, J.L., McCann, K.L., McStay, B., and Baserga, S.J.
146. Mottino, A.D., Hoffman, T., Crocenzi, F.A., Sánchez Pozzi, E.J., Roma, NOL11, implicated in the pathogenesis of North American Indian childhood
M.G., and Vore, M. Disruption of function and localization of tight junc- cirrhosis, is required for pre‐rRNA transcription and processing. PLoS
tional structures and Mrp2 in sustained estradiol‐17beta‐D‐glucuronide‐ Genet, 2012;8(8):e1002892.
induced cholestasis. Am J Physiol Gastrointest Liver Physiol, 2007;293(1): 170. Aagenaes, O., Sigstad, H., and Bjorn‐Hansen, R. Lymphoedema in
G391–402. hereditary recurrent cholestasis from birth. Arch Dis Child,1970;
­
147. Kojima, T., Yamamoto, T., Murata, M., Chiba, H., Kokai, Y., and Sawada, 45(243):690–5.
N. Regulation of the blood‐biliary barrier: interaction between gap and 171. Drivdal, M., Trydal, T., Hagve, T.A., Bergstad, I., and Aagenaes, O.
tight junctions in hepatocytes. Med Electron Microsc, 2003;36(3):157–64. Prognosis, with evaluation of general biochemistry, of liver disease in lym-
148. Furuse, M., Hata, M., Furuse, K. et  al. Claudin‐based tight junctions phoedema cholestasis syndrome 1 (LCS1/Aagenaes syndrome). Scand
are crucial for the mammalian epidermal barrier: a lesson from claudin‐1‐ J Gastroenterol, 2006;41(4):465–71.
deficient mice. J Cell Biol, 2002;156(6):1099–111. 172. Iversen, K., Drivdal, L.M., Billaud Feragen, K.J., and Geirdal, A.Ø. Quality
149. Smeitink, J., van den Heuvel, L., and DiMauro S. The genetics and pathol- of life in adults with lymphedema cholestasis syndrome 1. Health Qual Life
ogy of oxidative phosphorylation. Nat Rev Genet, 2001;2(5):342–52. Outcomes, 2018;16(1):146.
150. Visapää, I., Fellman, V., Vesa, J. et al. GRACILE syndrome, a lethal meta- 173. Bull, L.N., Roche, E., Song, E.J. et al. Mapping of the locus for cholestasis‐
bolic disorder with iron overload, is caused by a point mutation in BCS1L. lymphedema syndrome (Aagenaes syndrome) to a 6.6‐cM interval on
Am J Hum Genet, 2002;71(4):863–76. ­chromosome 15q. Am J Hum Genet, 2000;67(4):994–9.
29:  Molecular Cholestasis 363

174. Viveiros, A., Reiterer, M., Schaefer, B. et al. CCBE1 mutation causing scle- 191. Gershoni‐Baruch, R., Gottfried, E., Pery, M., Sahin, A., and Etzioni, A.
rosing cholangitis: Expanding the spectrum of lymphedema‐cholestasis Immotile cilia syndrome including polysplenia, situs inversus, and extrahe-
syndrome. Hepatol, 2017;66(1):286–288. patic biliary atresia. Am J Med Genet, 1989;33(3):390–3.
175. Shah, S., Conlin, L.K., Gomez, L. et al. CCBE1 mutation in two siblings, 192. Mazziotti, M.V., Willis, L.K., Heuckeroth, R.O. et al. Anomalous development
one manifesting lymphedema‐cholestasis syndrome, and the other, fetal of the hepatobiliary system in the Inv mouse. Hepatol, 1999;30(2):372–8.
hydrops. PLoS One,. 2013;8(9):e75770. 193. Shimadera, S., Iwai, N., Deguchi, E., Kimura, O., Fumino, S., and
176. Alders, M., Hogan, B.M., Gjini, E. et al. Mutations in CCBE1 cause gener- Yokoyama, T. The inv mouse as an experimental model of biliary atresia. J
alized lymph vessel dysplasia in humans. Nat Genet, 2009;41(12):1272–4. Pediatr Surg, 2007;42(9):1555–60.
177. Hogan, B.M., Bos, F.L., Bussmann, J. et al. Ccbe1 is required for embryonic 194. Otto, E.A., Schermer, B., Obara, T., O’Toole J.F., Hiller, K.S., Mueller,
lymphangiogenesis and venous sprouting. Nat Genet, 2009;41(4):396–8. A.M. et al. Mutations in INVS encoding inversin cause nephronophthisis
178. Clayton, P.T. Diagnosis of inherited disorders of liver metabolism. J Inherit type 2, linking renal cystic disease to the function of primary cilia and left‐
Metab Dis, 2003;26(2–3):135–46. right axis determination. Nat Genet, 2003;34(4):413–20.
179. Balistreri, W.F. Neonatal cholestasis. J Pediatr, 1985;106(2):171–84. 195. Olbrich, H., Fliegauf, M., Hoefele, J. et  al. Mutations in a novel gene,
180. Dick, M.C. and Mowat, A.P. Hepatitis syndrome in infancy – an epidemio- NPHP3, cause adolescent nephronophthisis, tapeto‐retinal degeneration
logical survey with 10 year follow up. Arch Dis Child,1985;60(6):512–6. and hepatic fibrosis. Nat Genet, 2003;34(4):455–9.
181. Ozkan, T.B., Mistik, R., Dikici, B., and Nazlioglu H.O. Antiviral therapy in 196. Johnson, C.A., Gissen, P., and Sergi, C. Molecular pathology and genetics
neonatal cholestatic cytomegalovirus hepatitis. BMC Gastroenterol, 2007;7:9. of congenital hepatorenal fibrocystic syndromes. J Med Genet, 2003;40(5):
182. Sokol, R.J., Shepherd, R.W., Superina, R., Bezerra, J.A., Robuck, P., and 311–9.
Hoofnagle J.H. Screening and outcomes in biliary atresia: summary of a 197. Adams, M., Smith, U.M., Logan, C.V., and Johnson C.A. Recent advances
National Institutes of Health workshop. Hepatol, 2007;46(2):566–81. in the molecular pathology, cell biology and genetics of ciliopathies. J Med
183. Balistreri, W.F., Grand, R., Hoofnagle, J.H., Suchy, F.J., Ryckman, F.C., Genet, 2008;45(5):257–67.
Perlmutter, D.H., et al. Biliary atresia: current concepts and research direc- 198. Ibañez‐Tallon, I., Heintz, N., and Omran, H. To beat or not to beat: roles of
tions. Summary of a symposium. Hepatol, 1996;23(6):1682–92. cilia in development and disease. Hum Mol Genet, 2003;12(1):R27–35.
184. Sokol, R.J., Mack, C., Narkewicz, M.R., and Karrer F.M. Pathogenesis and 199. Lina, F. and Satlinb, L.M. Polycystic kidney disease: the cilium as a com-
outcome of biliary atresia: current concepts. J Pediatr Gastroenterol Nutr, mon pathway in cystogenesis. Curr Opin Pediatr, 2004;16(2):171–6.
2003;37(1):4–21. 200. Phillips, M.J., Azuma, T., Meredith, S.L. et al. Abnormalities in villin gene
185. Sokol, R.J. and Mack, C. Etiopathogenesis of biliary atresia. Semin Liver expression and canalicular microvillus structure in progressive cholestatic
Dis, 2001;21(4):517–24. liver disease of childhood. Lancet, 2003;362(9390):1112–9.
186. Ohi, R. Surgery for biliary atresia. Liver, 2001;21(3):175–82. 201. Nicastro, E. and D’Antiga, L. Next generation sequencing in pediatric
187. Davenport, M., Savage, M., Mowat, A.P., and Howard E.R. Biliary atresia hepatology and liver transplantation. Liver Transpl, 2018;24(2):282–93.
splenic malformation syndrome: an etiologic and prognostic subgroup. 202. Chen, H.L., Li, H.Y., Wu, J.F. et al. Panel‐based next‐generation sequenc-
Surgery, 1993;113(6):662–8. ing for the diagnosis of cholestatic genetic liver diseases: clinical utility and
188. Karrer, F.M. and Lilly J.R. Correction of biliary atresia and jejunal atresia challenges. J Pediatr, 2018;171:171–7.
in an infant. J Pediatr Surg, 1993;168(4):469–76. 203. Togawa, T., Sugiura, T., Ito, K. et al. Molecular genetic dissection and neona-
190. Roach, J.P. and Bruny J.L. Advances in the understanding and treatment of tal/infantile intrahepatic cholestasis using targeted next‐generation sequenc-
biliary atresia. Curr Opin Pediatr, 2008;20(3):315–9. ing. J Pediatr, 2016;171:171–7.
Pathophysiologic Basis for
30 Alternative Therapies for
Cholestasis
Claudia D. Fuchs, Emina Halilbasic, and Michael Trauner
Hans Popper Laboratory of Molecular Hepatology, Division of Gastroenterology and Hepatology,
Department of Internal Medicine III, Medical University of Vienna, Austria

INTRODUCTION transporters in a post‐transcriptional manner) thereby counter-


acting cholestasis [2]. In addition, UDCA stimulates biliary
Cholestasis is characterized by impaired bile formation with bicarbonate (HCO3−) secretion supporting the “biliary HCO3−
insufficient amounts of bile reaching the duodenum resulting in umbrella” [3], an important protective mechanism for hepato-
intrahepatic and systemic accumulation of bile acids (BAs) and cytes and cholangiocytes against highly toxic BA monomers
other potentially toxic cholephiles [1]. Important causes include in  bile by maintaining alkaline pH along the apical surface.
disturbances of hepatocellular and/or cholangiocellular bile In  addition, UDCA may also promote biliary phospholipid
secretion and destruction/obstruction of smaller and/or larger secretion thereby facilitating the formation of mixed micelles
bile ducts by mechanical processes (e.g. stones, tumors) or and limiting the amounts of free, non‐micellar bound BAs. As
immune‐mediated fibrosing cholangiopathies such as primary chaperone UDCA also reduces endoplasmatic reticulum stress,
biliary cholangitis (PBC) and primary sclerosing cholangitis exerts anti‐apoptotic effects, and its activation of the glucocorti-
(PSC). Ideally, causal therapy of cholestasis resolves the coid receptor (GR/NR3C1) may be responsible for some direct
­underlying cholestatic insult and promotes recovery which is anti‐inflammatory effects, it further contributes to the cytopro-
currently possible only in some cases (e.g. removal of stone, tective action of UDCA in cholestatic liver disease [2].
surgical removal or stenting of tumor/stricture, delivery in intra- As such, UDCA has found application in the treatment of a
hepatic cholestasis of pregnancy, discontinuation of causative broad range of cholestatic diseases but it has been approved
drug in drug‐induced liver injury [DILI]). Even when the under- only as first‐line treatment for PBC, where UDCA improves
lying cause has been resolved, recovery can be slow (e.g. persis- survival [4]. In other cholestatic disorders, such as PSC, UDCA
tent hepatocellular secretory failure after DILI) requiring often improves serum liver tests, without proven survival bene-
supportive therapy. Since the etiologies of many cholestatic dis- fit. Increasing its dose (28–30 mg kg−1/day) was even harmful in
orders such as PBC and PSC are not fully understood and causal PSC [2] although the mechanisms are still incompletely under-
therapies are still lacking, stimulation of bile secretion and stood. However, amelioration in liver enzymes upon UDCA
adaptive mechanisms may attenuate liver injury irrespective of treatment (in moderate dose) and their exacerbation after UDCA
the cause of cholestasis. Of note, targeting the immune mecha- withdrawal may point toward potential beneficial effects. The
nisms underlying or at least contributing to the pathogenesis of field has made impressive progress since introduction of UDCA
PBC and PSC has so far been rather disappointing. [2]. The identification of specific BA transport systems and
The first‐line therapy for many cholestatic liver diseases is dedicated BA receptors [5] now allows a more specific targeting
ursodeoxycholic acid (UDCA), a hydrophilic BA that reduces of impaired BA homeostasis (including adaptive/alternative
the toxicity of the endogenous BA pool and has several transport, metabolic pathways, and regulatory networks) irre-
­additional beneficial therapeutic mechanisms [2]. As a potent spective of the etiology of cholestasis. This chapter will focus
intracellular signaling molecule UDCA stimulates hepatobiliary on the pathophysiologic basis for alternative therapies for chol-
secretion (mainly by promoting vesicular targeting of estasis beyond UDCA, the first available drug for cholestasis.

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
30:  Pathophysiologic Basis for Alternative Therapies for Cholestasis 365

NUCLEAR RECEPTORS BA activation of FXR in enterocytes and reaches the liver,


where it binds to fibroblast growth factor receptor 4 (FGFR4)/β‐
Nuclear receptors (NRs) are the largest family of transcription Klotho complex and activates c‐Jun N‐terminal kinase‐­
factors comprising 48 members in humans and 49 members in dependent pathway which ultimately leads to CYP7A1
mice, which share a common structural organization and are repression [15] (Figure  30.1). Of note, SHP is required for
activated upon ligand binding [6]. They coordinate several key FGF15/19 to efficiently repress BA synthesis and mice lacking
hepatic functions including the regulation of BA synthesis and SHP are refractory to the inhibitory effects of either FXR
hepatobiliary excretory function, glucose and lipid metabolism, ­agonists or FGF15/19 on Cyp7a1 expression [7]. FXR further
inflammatory processes, fibrosis as well as liver regeneration, fine‐tunes BA levels through induction of other target genes,
and tumorigenesis [7]. After ligand (e.g. BA) binding, NRs including transcription factor V‐Maf avian musculoaponeu-
change their conformation, facilitating the recruitment of rotic  fibrosarcoma oncogene homolog G (MAFG), a global
­cofactors and dissociation of corepressors, implementing DNA transcriptional repressor of BA synthesis, lysine‐specific his-
binding and stimulating transcription of genes typically involved tone demethylase (LSD1), a key repressive epigenetic compo-
in metabolism and/or transport of the ligand (e.g. BAs) thus nent in a SHP complex, and β‐Klotho, the obligate co‐receptor
constituting the molecular basis for feedback regulation of for FGF15/19 [17].
metabolism [6]. Due to their central role in hepatic (patho)phys- Apart from regulation of BA synthesis, the intrahepatic
iology, NRs have become attractive targets for anticholestatic BA  concentration is controlled by FXR‐mediated inhibition
drugs serving as ligands for these receptors. The most relevant of  the basolateral BA uptake transporter sodium/taurocholate
NRs involved in BA signaling and hepatobiliary homeostasis co‐transporting polypeptide (NTCP/SLC10A1) and upregula-
include the farnesoid X receptor (FXR/NR1H4), pregnane tion of the canalicular BA transporter bile salt export pump
X  receptor (PXR/NR1I2), constitutive androstane receptor (BSEP/ABCB11) in hepatocytes (Figure 30.1), promoting BA
(CAR/NR1I3), and vitamin D receptor (VDR/NR1I1). elimination via induction of bile flow. Beside this induction of
Furthermore, the glucocorticoid receptor (GR/NR3C1) and the BA dependent bile flow, FXR also increases BA‐independent
fatty acid activated peroxisome proliferator‐activated receptors bile flow promoting biliary HCO3− transport by facilitating
(PPARα, PPARγ, and PPARδ) as regulators of inflammation, complex formation of carboanhydrase 14 and anion exchanger 2
fibrosis, and energy metabolism may also impact on biliary (AE2) [18]. FXR also induces alternative basolateral BA trans-
homeostasis and thereby counteract cholestatic liver injury. port in hepatocytes via organic solute transporter alpha and beta
Notably the current standard therapy of cholestasis, UDCA, (OSTα/β; SLC51A/SLC51B) and detoxification via enzymes
does not activate FXR (but may even counteract its activation by such as Cyp3a11/CYP3A4, Sult2a1/SULT2a1 (sulfotransferase
reducing the relative concentration of BAs with more potent family 2A member 1), and Ugt2b4/UGT2b4 [7] further prevent-
FXR ligand function [8]), and has low affinity to GR and VDR ing the accumulation of BAs within target cells (Figure 30.1).
mediating its anti‐inflammatory effects and may indirectly acti- In the liver, FXR is also present in cells other than hepato-
vate PXR after being metabolized into lithocholic acid (LCA) [2]. cytes. In cholangiocytes FXR regulates ductular bile formation
by alkalization and fluidization via induction of vasoactive
intestinal polypeptide receptor 1 (VPAC‐1) [7]. Moreover, FXR
Nuclear bile acid receptor (FXR)
activation in cholangiocytes induces expression of the antimi-
FXR serves as the main NR for BAs [8–10] regulating their crobial peptide cathelicidin, indicating an important role of
­synthesis and hepatic uptake as well as elimination from the BAs/FXR in maintaining sterility of biliary tree and protection
liver. It forms heterodimers with the retinoid X receptor (RXR, against bile duct inflammation [19].
NR2B), binding to the inverted repeat (IR)‐1 within the pro- Activation of FXR has antifibrotic effects [18], the underly-
moter sequence of target genes. FXR is predominantly present ing mechanisms and FXR expression in hepatic stellate cells
in liver, gastrointestinal tract, kidney, and adrenal glands acting (HSC) are discussed controversially. FXR has been reported to
as a central regulator of BA homeostasis, lipid, glucose [11], have a direct antifibrotic effect via activation of SHP in HSC
and amino acid metabolism [12] as well as inflammation [7] and [7], but other studies, however, have revealed only very low or
liver regeneration [13]. even no FXR and SHP expression in human HSCs and murine
FXR has a key role in controlling BA synthesis by regulating periductal myofibroblasts [20], arguing for a more indirect anti-
the expression of rate limiting enzymes of both classical fibrotic effect through control of cholestasis and inflammation.
(CYP7A1) and alternative pathways (CYP8B1 and CYP27A1). Present also in liver sinusoidal endothelial cells [21], FXR
This is mediated by induced expression of two key regulators, ameliorates portal hypertension not by targeting fibrosis, but by
an orphan nuclear receptor, small heterodimer partner (SHP, directly interfering with endothelial function. FXR activation
NR0B2) [11] and an intestinal hormone, fibroblast growth fac- reduces intrahepatic vascular resistance by upregulation of
tor 15/19 (human FGF19, mouse Fgf15) [14, 15]. BA‐activated eNOS, cystathionase, and dimethylaminohydrolase 1, promoting
FXR induces the hepatic expression of SHP, an atypical nuclear the generation of vasodilators such as nitric oxide and hydrogen
receptor lacking a DNA binding domain that functions as a sulfide, as well as a reduction of intrahepatic vasoconstrictors
potent transcriptional repressor, which in turn, interacts with endothelin‐1 and p moesin [18].
transcription factors such as hepatocyte nuclear factor (HNF)‐4α/ Mice lacking FXR display increased hepatic inflammation
NR2A1 and liver receptor homolog (LRH)‐1/NR5A2 to inhibit at  baseline and are more susceptible to lipopolysaccharide
their function and suppress CYP7A1 expression [16]. The ileal (LPS)‐induced inflammation [18]. FXR may have direct anti‐
hormone FGF15/19 is secreted in the portal circulation upon inflammatory effects by interference with nuclear factor
366 THE LIVER:  NUCLEAR RECEPTORS

MRP3

MRP4

OSTβ
OSTα
Hepatic stellate cell
FXR PXR PPARα
PXR PXR CAR
GR FXR
Inflammation
e.g. NFκB
Alternative BA export
FXR secretion of hydroxylated &
FGF βKlotho FGF BSEP BAs sulfated BAs
15/19
FGFR4 15/19 PPARγ PXR VDR

FXR
GR Fibrosis
MDR2/3 PL
PPARγ: MCP1, migration, proliferation
JNK PPARα HCO3 - AE2 PXR: TGF1β, transdifferentiation, proliferation
BAs NTCP
VDR: SMAD pathway

FXR FXR PXR CAR PPAR

FXR conjugated α/δ Immune cells


MRP2
Bilirubin BA detoxification
SHP PPAR PXR CYPs, SULTs, UGTs
BA synthesis α/δ
CYP7A1 Macrophage CD4+ T cell
Hepatocyte
PPARα PPARγ PPARδ VDR
PPARγ
FXR VDR GR
Reactive cholangiocyte Inflammation
Phenotype
Cholangioprotective PPARα: phosphorylation of p38 in CD4+ T cells
VCAM1 VPAC1, AE2, Cathelecithin PPARγ: cytocine secretion from macrophages
PPARδ: M2 activation of macropages
VDR: production of Tregs
Cholangiocyte

Figure 30.1  Nuclear receptors as therapeutic targets of alternative therapies in cholestasis. Nuclear receptor activation in hepatocytes maintains
the balance between BA synthesis, detoxification, uptake, and excretion by regulation of expression of hepatobiliary transporters and metabolic
pathways. FXR (in a SHP dependent manner) represses hepatic BA uptake (NTCP) and BA synthesis (CYP7A1). Moreover intestine‐derived
FGF15/19 to the FGFR4/bKlotho dimer counteracts CYP7A1 expression. Of note, human hepatocytes (in addition to enterocytes) also express
FGF19. FXR promotes BA secretion via induction of canalicular transporters (BSEP, MRP2, and MDR2/3) and induces BA elimination via alterna-
tive basolateral BA transporter OSTα/β. CAR and PXR regulate the adaption to increased intracellular BA concentrations via upregulation of MRP3
and 4 (alternative BA export) as well as induction of detoxification/hydroxylation enzymes. PPARα stimulates phospholipid secretion (via MDR2/3)
and is also involved in regulation of detoxification pathways. Activation of AE2 expression by GR stimulates biliary bicarbonate secretion, thus
reducing bile toxicity. Apart from regulating BA homeostasis, NRs have additional anti‐inflammatory and antifibrotic effects (right‐hand panel).
Their activation may result in induction of defensive mechanisms in bile duct epithelial cells. Activation of PPARγ in cholangiocytes reduces vas-
cular cell adhesion molecule (VCAM‐1) expression, thereby counteracting reactive cholangiocyte phenotype. Activation of FXR, VDR, and GR in
cholangiocytes exert cholangioprotective effects via upregulation of VPAC1, AE2, and Cathelicidin, respectively. Antifibrotic effects of NRs are
related to their activation in hepatic stellate cells. PPARγ, PXR, and VDR reduce expression of profibrogenic genes such as MCP1, TGFβ1, and
SMAD, respectively. Furthermore, they reduce migration, proliferation, as well as transdifferentiation of HSC into myofibroblasts. Anti‐inflamma-
tory effects of NRs are related to their activation in immune cells such as macrophages and CD4+ T cells. Green arrows indicate stimulatory effects
and red lines indicate suppressive effects. AE, anion exchanger; BAs, bile acids; Bili‐glu, bilirubin glucuronide; BSEP, bile salt export pump; CAR,
constitutive androstane receptor; CYP7A1, cholesterol‐7α‐hydroxylase; CYPs, cytochrome P450 enzymes; FGF, fibroblast growth factor; FXR,
farnesoid X receptor; GR, glucocorticoid receptor; HSC, hepatic stellate cell; MDR3, multidrug resistance protein 3; MRP2, multidrug resistance‐
associated protein 2; MRP3, multidrug resistance‐associated protein 3; MRP4, multidrug resistance‐associated protein 4; NTCP, sodium taurocho-
late cotransporting polypeptide; OSTα/β, organic solute transporter α and β; PC, phosphatidylcholine; PXR, pregnane X receptor; PPARα,
peroxisome proliferator‐activated receptor α; PPARγ, peroxisome proliferator‐activated receptor γ; SHP, small heterodimer partner; SULTs, sulfa-
tation enzymes; UGTs, glucuronidation enzymes; VDR, vitamin D receptor.

kappa‐b (NFκB) [18]. Recently, FXR has also been implicated inflammasomes [22] and on the other hand specific BA species
in activation of hepatic natural killer T cells and hepatic accu- (LCA) have been shown to inhibit inflammasome activation [23].
mulation of myeloid‐derived suppressor cells, counteracting Beside its direct anti‐inflammatory properties and modulation
immune‐mediated liver injury in rodents [18]. Anti‐­ of immune cell function, FXR represents a central molecular
inflammatory effects of FXR were not only observed in liver, switch controlling intestinal microbiota by protecting the gut
but also in intestine since pharmacological FXR activation from bacterial overgrowth and disruption of the epithelial bar-
improved intestinal inflammation and permeability in colitis rier [24]. Conversely, FXR dysfunction drives bacterial translo-
models [7]. Reciprocal interactions between BAs and immune cation, facilitated by increased intestinal permeability and
cells may exist, since the BA receptors FXR and TGR5/GBAR‐1 intestinal inflammation [18]. Pharmacological activation of
(Takeda G protein‐coupled receptor/G protein‐coupled bile acid FXR improved gene expression of tight junction proteins such
receptor 1) are expressed in macrophages [22]. BAs have been as claudin‐1 and occludin [25] as well as expression of induci-
identified on one hand as damage‐associated molecular patterns ble isoform nitric oxide synthase (iNOS) and angiopoietin
(DAMPs), triggering inflammatory response by binding to 1 (Ang1), genes involved in antibacterial defense by producing
30:  Pathophysiologic Basis for Alternative Therapies for Cholestasis 367

antimicrobial peptides [26]. Activation (or intestinal overex- of OCA in other cholestatic liver disorders such as PSC is
pression [24]) of FXR decreased bacterial overgrowth and undergoing clinical evaluation [34].
improved liver injury in the bile duct ligation (BDL) [24] as well Non‐steroidal FXR agonists (such as LJN452/Tropifexor and
as in the α‐naphthyl‐isothiocyanate (ANIT) model [27]. In line, GS‐9674) have different pharmacokinetic properties favoring
obstructive jaundice with virtually complete absence of BAs intestinal (instead of hepatic) FXR activation, and also do not
from the intestinal lumen promotes bacterial translocation in undergo enterohepatic circulation. Differential targeting of FXR
patients [28]. in hepatocytes, cholangiocytes, and/or enterocytes could criti-
Loss of FXR results in severe disruption of BA homeostasis cally determine the pharmacological effects of steroidal versus
as reflected by increased BA synthesis as well as increased BA non‐steroidal FXR ligands. Whether non‐steroidal FXR ligands
uptake and reduced hepatobiliary BA secretion in FXR knock- lack possible BA‐structure‐related side‐effects, such as pruritus,
out (KO) mice [29]. Although metabolic adaption to cholestatic needs to be determined [35]. An open question remains whether
conditions does not entirely rely only on FXR and – at least in activation of intestinal FXR resulting mainly in inhibition of
mice – can be partially compensated by CAR/PXR‐dependent BA synthesis via FGF15/19 signaling is sufficient to counteract
overexpression of detoxifying enzymes and alternative efflux cholestatic liver disease or whether more systemic anti‐­
systems [18, 30], a loss‐of‐function mutation in the FXR gene inflammatory effects are necessary.
results in severe neonatal progressive familial intrahepatic chol- Another important G‐protein coupled BA receptor, TGR5, is
estasis 5 (PFIC5) [31]. While impaired function of FXR leads to expressed in cholangiocytes, gallbladder epithelium, as well as
profound dysregulation of BA homeostasis, pharmacological endothelial cells and Kupffer cells (but not in hepatocytes),
activation of FXR has shown multiple beneficial effects in pre- where it regulates biliary HCO3− secretion and has anti‐inflam-
clinical models of cholestasis. Collectively, FXR: (i) suppresses matory effects [18] (also see Chapter  24 for more details). In
hepatic BA synthesis, reduces intestinal and hepatic BA uptake, contrast to FXR [18], TGR5 stimulates production of intestinal
while stimulating canalicular and basolateral BA efflux, thus glucagon like peptide 1 (GLP‐1) [7]. GLP‐1 signaling may pro-
resulting in a net reduction of the toxic hepatic BA overload; (ii) tect cholangiocytes from apoptosis and thus modulate their
alters bile composition by increasing phospholipid and HCO3− adaptive response to cholestasis in mice [18]. However, TGR5
secretion finally resulting in less toxic bile; (iii) mediates direct expression is reduced in Mdr2 KO mouse model of sclerosing
anti‐inflammatory effects in hepatocytes and non‐parenchymal cholangitis [36] and a pure TGR5 agonist did not improve liver
liver cells, together with immunomodulatory effects in the adap- disease in this model [18]. Additional concerns include a
tive immune system; (iv) impacts on the gut–liver axis by induc- ­potential role of TGR5 in development of pruritus [18], poten-
tion of FGF15/19, a suppressor of BA synthesis; (v) reduces tial pro‐proliferative and anti‐apoptotic effects in isolated
bacterial overgrowth and intestinal permeability; (vi) amelio- human cholangiocytes and cholangiocellular carcinoma cells
rates the complications of advanced liver disease such as portal [18] as well as TGR5 overexpression in human cholangiocarci-
hypertension and possibly carcinogenesis. These features make noma [37] (for more details see Chapter 24).
FXR ligands attractive compounds for alternative therapies in
cholestasis and immune‐mediated disorders such as PBC and
PSC including associated/underlying inflammatory bowel
Fibroblast growth factor 19 (FGF19)
­disease (IBD). Activation of FXR stimulates production of endogenous
Various steroidal (BA derived) and non‐steroidal (non‐BA FGF15/19 in the ileum, which is transported to the liver where
derived) FXR activators have been developed and have shown a it binds to FGFR4 and its co‐receptor β‐Klotho in hepatocytes,
range of beneficial effects in animal models of cholestasis. contributing to suppression of hepatic BA synthesis via c‐Jun
More specifically, the non‐steroidal FXR agonist GW4064 and N‐terminal kinase (JNK) and SHP, in addition to other effects
the steroidal FXR agonist obeticholic acid (OCA/INT747  –  a on lipid and glucose metabolism [15]. Interestingly, FGF19 may
6‐ethyl derivative of the primary BA chenodeoxycholic acid also exert anti‐inflammatory effects by counteracting NFκB
[18]) showed positive effects in chemically induced cholestatic signaling [38]. Selective overexpression of intestinal FXR
liver injury (α‐naphthyl isothiocyanate (ANIT) and estradiol improved liver injury in the Mdr2 KO mouse model of scleros-
induced cholestasis), in bile duct ligated animals as well as in ing cholangitis via induction of intestinal FGF15 expression and
the Mdr2 KO mouse model of sclerosing cholangitis [18]. subsequent reduction of BA pool size. In line, FGF19 applica-
Based on these encouraging preclinical data, FXR ligands tion to bile duct ligated mice reduced total BA pool size [39],
have already entered clinical development. Add‐on treatment improved jaundice and serum levels of liver transaminases as
with OCA in PBC patients, who biochemically did not respond well as necrosis. However, stimulation of endogenous FGF15/19
sufficiently to UDCA or OCA monotherapy, improved liver could also have side‐effects such as pro‐proliferative or pro‐car-
enzymes and markers of inflammation [7, 32, 33]. These studies cinogenic effects. Endogenous FGF19 and its receptor FGFR4
led to conditional approval of OCA as a second‐line therapy for have been implicated in development of hepatocellular carci-
the treatment of PBC patients with incomplete biochemical noma (HCC) and inhibitors of FGFR4‐related signaling are
response or intolerance to standard therapy with UDCA, which even developed as anticancer drugs [18]. As such, FGF19 is
is already reflected by current guidelines [4]. Long‐term clinical amplified in human HCC and in mice ectopic overexpression of
effects of OCA treatment in PBC patients are currently under FGF19 accelerates tumor development [18]. Importantly,
investigation. Since OCA functions as systemic FXR ligand FGF19 mimetics such as M70/NGM282 have amino acid modi-
undergoing enterohepatic circulation the risk of overdosing in fications at the N‐terminus allowing the dissection of metabolic
cirrhotic PBC patients needs to be taken into account. The role from pro‐proliferative actions thus maintaining their BA
368 THE LIVER:  NUCLEAR RECEPTORS

metabolic regulatory properties while lacking their pro‐­ sequesters NFκB) [45]. Fenofibrate (a more selective PPARα
proliferative effects. In line, adeno‐associated virus mediated ligand) downregulates the IL6 receptor subunits gp80 and gp130
expression of M70 in Mdr2 KO mouse model of sclerosing in the liver [46], thereby reducing phosphorylation of STAT3
cholangitis decreased hepatic inflammation and biliary fibrosis and c‐Jun [47].
[40]. Notably, prolonged M70 treatment did not accelerate HCC In addition to PPARα, PPARγ and PPARδ activation reduces
formation in the Mdr2 KO mouse and even had tumor preven- production of inflammatory mediators and cytokines in various
tive effects, while FGF19 application promoted hepatic carcino- immune cells including monocytes/macrophages, dendritic
genesis [40]. In PBC patients with insufficient response to cells, lymphocytes, as well as cholangiocytes [7].
UDCA NGM282 improved cholestatic liver enzymes [40]. Activation of all PPAR isoforms also has antifibrotic effects.
Diarrhea, headache, and nausea were observed as side‐effects, Although HSC do not show pronounced PPARα expression, its
suggesting that FGF19 may also impact on intestinal motility. activation led to reduced HSC activation and repression of
Of note, NGM282 has been shown to improve markers of fibrotic markers [48]. PPARγ is expressed only in quiescent
hepatic fibrosis without reducing cholestatic liver enzymes such HSC and has antifibrotic effects by suppressing their activation,
as alkaline phosphatase in patients with PSC [40] suggesting proliferation, and migration [49]. Accordingly, loss of PPARγ is
that FGF19 could also mediate potential direct anti‐inflammatory a typical feature of HSC activation [49].Treatment with thiazo-
and antifibrotic actions. To understand the full (patho)physio- lidinedione (a PPARγ agonist) improved bile duct proliferation
logical and therapeutic implications of FGF19 it has to be kept and hepatic fibrosis in bile duct ligated rats [7]. The plant extract
in mind that  –  in contrast to rodents  –  human FGF19 is also curcumin, the yellow pigment of the spice turmeric, also acti-
produced by hepatocytes, bile ducts, and gallbladder epithelium vated PPARγ in cholangiocytes, decreased infiltration of inflam-
[41]. Interestingly, serum FGF19 correlates positively with matory cells, and improved portal inflammation and fibrosis in
severity of the cholestasis [18]. The extent to which patients Mdr2 KO mouse model [7]. In contrast to PPARα and PPARγ,
with elevated endogenous levels of FGF19 may benefit PPARδ activation had pro‐ as well as antifibrotic effects.
from  exogenous FGF19 mimetic application remains to be Application of KD3010, but not GW501516, (two different
­elucidated. For example, the FGF19 mimetic could also com- PPARδ agonists) was hepatoprotective and antifibrotic in mice
pete with endogenous FGF19 at the FGFR4 binding site [18] subjected to BDL and CCl4. In vitro that KD3010, in contrast to
potentially counteracting development of HCC, while maintain- GW501516, increased the expression of several CYP enzymes
ing the beneficial metabolic effects on BA homeostasis. resulting in reduced reactive oxygen species production, possi-
bly counteracting hepatic fibrosis [50].
Peroxisome proliferator‐activated receptors Patients with obstructive jaundice undergoing percutaneous
transhepatic biliary drainage that received bezafibrate treatment
(PPARs)
showed increased biliary PL secretion. However, the same study
PPARα (NR1C1), PPARγ (NR1C3), and PPARδ (NR1C2), are showed that MDR3 expression is already increased in PBC
structural homologues which are activated by endogenous fatty patients at baseline and was not further upregulated by bezafi-
acids and their derivatives [7]. PPARδ is ubiquitously expressed, brate treatment [51], indicating that bezafibrate may induce
while PPARα is predominantly found in organs responsible for beneficial clinical effects via PL‐independent mechanisms.
fatty acid catabolism such as liver, heart, kidney, brown adipose Fibrates such as bezafibrate and fenofibrate have shown ben-
tissue, small and large intestine. PPARγ is highly expressed in eficial effects in several smaller, mostly uncontrolled studies in
adipose tissue and immune cells [42]. The main function of patients with PBC (not responding to UDCA) and  –  to lesser
PPARs is the regulation of lipid and energy homeostasis. extent – in PSC [52]. Recently, a larger randomized controlled
Beyond the realm of metabolism, PPARs also exert direct anti‐ trial demonstrated pronounced improvement in liver enzymes,
inflammatory effects. liver stiffness, and pruritus by bezafibrate in PBC patients with
Apart from being a key regulator of hepatic lipid metabolism, insufficient response to UDCA [53] and fibrates may be consid-
PPARα is also involved in BA metabolism and BAs can indi- ered a second‐line treatment in PBC [54]. Future long‐term
rectly stimulate human PPARα via FXR [43]. PPAR activation studies will have to explore the safety and efficacy of fibrates in
via bezafibrate represses BA synthesis (through reduction of PBC and other cholestatic disorders.
HNF4α binding to the CYP7A1 promoter) [7] as well as hepa- A selective and potent PPARδ agonist MBX‐8025, Seladelpar,
tocellular BA uptake (reduced Ntcp expression by direct mecha- reduced alkaline phosphatase in a phase II proof‐of‐concept
nisms) [44]. Bezafibrates (PPARα ligand with additional study in PBC [55] and further clinical trials are ongoing.
broader pan PPAR activity) also induce expression of enzymes Another promising PPARα/δ dual agonist, elafibranor, with
involved in BA detoxification (Sult2a1, Ugt2b4, and Ugt1a3) beneficial effects in non‐alcoholic steatohepatitis (NASH) and
[7]. In vitro, the PPARα agonist ciprofibrate also induced the diabetes [56], is currently also being tested in PBC.
BA uptake transporter ASBT (apical sodium dependent bile salt
transporter) in human cholangiocytes and Caco2 cells [2].
PPARα stimulates expression of multidrug resistant 3 (MDR3)
Glucocorticoid receptor (GR)
[44] thereby potentially contributing to elevated biliary GR/NR3C1 is ubiquitously expressed. Apart from its well‐
­phospholipid (PL) secretion and counteracting BA toxicity by known anti‐inflammatory and immunosuppressive function, GR
promoting formation of mixed micelles. regulates several metabolic pathways including not only carbo-
In addition, PPARα exerts anti‐inflammatory effects by hydrate and protein, but also BA homeostasis [2]. Glucocorticoids
­transrepressing NFκB signaling via induction of IκBα (which (via GR) induce BA uptake systems such as NTCP, ASBT, and
30:  Pathophysiologic Basis for Alternative Therapies for Cholestasis 369

OSTα/β [2]. Ligand‐activated GR also upregulates AE2 expres- CAR ligands such as phenobarbital and PXR ligands such as
sion, which should result in increasing cholangiocellular bicar- rifampicin have been used to treat pruritus for decades [43]. These
bonate secretion and restoring the bicarbonate umbrella [57]. antipruritogenic effects could be explained by their impact on BA
This observation is of particular interest since AE2 expression is detoxification and alternative BA elimination outlined above. Of
reduced in liver and inflammatory cells of PBC patients [2], note, rifampicin is recommended as second‐line treatment for
possibly responsible for making cholangiocytes more suscepti- pruritus by the current guidelines [4]. Moreover, PXR and CAR
ble to an autoimmune first hit. Besides glucocorticoids, UDCA activation may not only alleviate pruritus, but also have antichole-
also activates GR. The combination of UDCA and another GR static effects. As such, rifampicin (PXR ligand) improved liver
ligand dexamethasone (but not UDCA or dexamethasone alone) function in PBC patients [63] and p­ henobarbital (CAR ligand)
increased expression and function of AE2 via interaction with reduced serum BA concentration [43]. Interestingly, traditional
HNF1 and GR on the AE2 alternate promoter [57]. UDCA‐­ Chinese medicines which were used to treat neonatal jaundice by
activated GR is also anti‐inflammatory through suppressing enhancing bilirubin clearance have been identified as CAR
NFκB‐dependent transcription by preventing GR – p65 interac- ligands (e.g. Yin Chin/Artemisia capillaris) [64].
tion [2] representing another example for the dichotomous Stimulation of an adaptive response to cholestasis may
effects of NRs on both metabolism and inflammation. ­particularly be useful when the underlying cause cannot be
Glucocorticoids signal not only via GR, but also other NRs improved. Moreover, persistent secretory failure after DILI and
such as CAR, and glucocorticoids increase expression of PXR and prolonged mechanical cholestasis may be overcome by PXR
RXRα [43]. Budesonide also activates PXR in mice and humans, ligands such as rifampicin [2]. Stimulation of detoxification
while dexamethasone directly activates PXR in mice but not in (e.g. via CAR and PXR agonists) may complement other thera-
humans [58]. Clinically prednisolone and budesonide were tested pies (e.g. FXR agonists) targeting mainly hepatobiliary flow
in combination with UDCA in PBC patients. While the combina- and excretion of BAs and other cholephiles. Therefore, more
tion of prednisolone with UDCA was not superior to UDCA mon- specific and less toxic CAR/PXR agonists could be a valuable
otherapy [2], budesonide UDCA combination significantly addition to the therapeutic armamentarium in cholestasis.
improved serum liver tests compared to UDCA alone [2].
Vitamin D and A receptors
Xenobiotic sensors PXR and CAR Vitamin D receptor (VDR/NR1I1) is not only involved in the
Pregnane X receptor (PXR/NR1I2) and constitutive androstane regulation of calcium homeostasis, but also in cell proliferation
receptor (CAR/NR1I3) are key regulators of enzymes involved and differentiation. Expression of VDR is high in intestine and
in BA hydroxylation/detoxification [2] and alternative BA low in liver where it is mainly found in components of the
export. Both receptors are low affinity, broad specificity xenosen- immune system including Kupffer cells, monocytes, natural
sors, activated by a range of compounds which are structurally killer cells, T and B lymphocytes, and dendritic cells, endothe-
unrelated such as antibiotics including rifampicin, clotrimazole, lial cells, cholangiocytes, and HSCs [43]. Vitamin D represses
the antidepressant St. John’s wort, and synthetic steroids such as Th1 and Th17 response and increases differentiation of
5b‐pregnane‐3,20‐dione, pregnenolone 16a‐carbonitrile (PCN) ­regulatory T cells and NK cells, implying a critical role in the
and dexamethasone [2, 59]. In addition to xenobiotics, also pathogenesis of autoimmune diseases [65]. In line, VDR poly-
potentially toxic endogenous cholephiles such as secondary BAs morphisms are associated with several immune‐mediated liver
(e.g. LCA) and bilirubin activate these receptors [60]. diseases such as PBC and autoimmune hepatitis [7]. VDR,
PXR and CAR mitigate potentially harmful effects of toxic ­acting as a potential BA sensor along the gut–liver axis, also
BAs by coordinated activation of detoxification pathways such functions as a receptor for the secondary BA LCA, which is
as hydroxylation (Cyp3a11/CYP3A4, and CYP2B10), sulfation rather hydrophobic/cytotoxic. Activation of VDR by LCA
(SULT2a1/Sult2a1), conjugation (BACS/Bacs, BAT/Bat), and ­stimulates Cyp3a11 (and its human homologue of CYP3A4)
subsequent elimination via basolateral export pumps multidrug expression, an enzyme that detoxifies LCA in liver and intestine
resistance‐associated protein 3 (MRP3) and MRP4 (Figure 30.1) [7]. Additionally, SULT2A1/Sult2a1 (an enzyme catalyzing
[43]. In addition, they regulate bilirubin metabolism in the liver ­sulfation of BAs) as well as basolateral export pump MRP3 was
inducing its glucuronidation (UGT1A1/Ugt1a1) and canalicular upregulated by vitamin D application in enterocytes [7]. The
excretion (MRP2) (Figure 30.1) [43]. PXR was shown to inhibit role of the VDR pathway in BA detoxification and elimination
the interaction between PGC1α and HNF4α, thereby repressing may have also emerged to shield first‐pass tissues from the
CYP7A1 gene transcription and BA synthesis [7]. Interestingly, ­cytotoxic effects of BA overload.
PXR was identified as FXR target gene [61], pointing toward a Notably, VDR activation in cholangiocytes also increased
potential NR crosstalk in the protection from BA toxicity. expression of the antimicrobial peptide cathelicidin, which
Besides beneficial effects on BA detoxification and elimina- could contribute to the anti‐inflammatory properties of VDR in
tion, PXR may also exert direct anti‐inflammatory and antifi- immune‐mediated bile duct injury [7]. Cathelicidin is able to
brotic effects. In mice, activation of PXR inhibits inflammation neutralize the deleterious effect of LPS which accumulates in
and fibrosis, while PXR KO mice are more susceptible to the biliary tree in fibrosing cholangiopathies [19]. Furthermore,
inflammatory stimuli [7]. In human HSCs, activation of PXR activation of VDR ameliorated fibrosis, possibly by epigenetic
inhibits expression of the major profibrogenic cytokine TGF‐1β, modifications of the SMAD pathway in HSCs [2]. However,
preventing their transdifferentiation to fibrogenic myofibro- administration of vitamin D to Mdr2 KO mice did not improve
blasts and slows down proliferation [62]. hepatic fibrosis despite ameliorating serum liver enzymes [2].
370 THE LIVER:  TRANSPORTER MODULATION

Low vitamin D levels have been linked to progression of fibro- BA concentrations [18, 78]. Surgical interruption of the entero-
sis in various liver diseases including cholestatic liver injury [7]. hepatic BA circulation has been a longstanding therapeutic strat-
Low serum vitamin D level is also an indicator of advanced dis- egy in PFIC [79]. In addition, increasing fecal BA concentrations
ease and a predictor of UDCA non‐response in PBC patients and exposure of gut epithelial cells to BAs can stimulate poten-
[66]. Whether (high dose) vitamin D supplementation may also tial beneficial colonic BA signaling such as TGR5‐­regulated
improve hepatic immune‐tolerance, thereby reducing biliary GLP‐1 secretion from enteroendocrine L‐cells [80]. GLP‐1 has
injury and subsequent fibrosis in PBC remains to be evaluated. cholangioprotective effects protecting these cells from apoptosis
Vitamin A is the umbrella term for diverse retinoids and their and attenuating the reactive cholangiocyte phenotype [18].
precursor forms. Retinoids are important regulators of cell pro- Application of highly potent and selective ASBT inhibitors
liferation and differentiation that bind to two families of NRs, (Lopixipat and A4250) to Mdr2 KO mice increased fecal BA
retinoic acid receptors (RARα, β, and γ/NR1B1, B2, and B3) loss which cannot be compensated by increased endogenous BA
and retinoid X receptors (RXRα, β, and γ/NR2B1, B2, and B3) synthesis. Thus, despite increased Cyp7a1 and repressed Fgf15
[65]. A biologic active form of vitamin A  –  all‐trans retinoic expression by ASBT inhibition, hepatic as well as biliary BA
acid (ATRA) and its stereoisomer, 9‐cis‐retinoic acid – activate concentration and bile flow were reduced, leading to a beneficial
RARs, whereas RXRs are mainly activated by 9‐cis‐retinoic BA/PL ratio, resulting in an improvement of liver and bile duct
acid [67, 68]. RARs and RXRs are known to be key regulators injury in this model of biliary toxicity [18, 78]. Furthermore,
in activation of HSC, transdifferentiating them from vitamin ASBT inhibitor treatment also reduced recruitment of Kupffer
A‐storing cells to myofibroblasts, which are proliferative, con- cells and neutrophils to the liver and promoted expansion of anti‐
tractile, inflammatory, and chemotactic and characterized by inflammatory monocytes [18]. Similarly, application of the BA
enhanced matrix production [69]. In line with loss of vitamin A, sequestrant colesevelam completely reversed liver and bile duct
activation of HSC is accompanied by downregulation of RAR injury [78], further indicating that interruption of enterohepatic
and RXR expression [70]. On the other hand, activation of RAR circulation of BAs may have therapeutic potential in attenuating
and RXR may attenuate HSC activation and proliferation [71], cholestatic liver disease beyond their role in treatment of pruri-
and has been shown to suppress TGFβ and IL6 mRNA expres- tus. Although both therapeutic options result in reduced hepatic
sion, thereby reducing fibrosis [72]. ATRA also interferes with BA uptake and thereby lowering hepatobiliary BA toxicity, their
the innate immune system and inhibits proinflammatory mechanism of action may differ profoundly since enhanced con-
cytokine expression of macrophages [73]. version of primary into secondary BAs (seen under colesevelam,
ATRA also repressed CYP7a1 promoter activity [74], which but not under ASBT inhibitor treatment), may change intestinal
was complemented by UDCA [75], indicating potential syner- BA signaling. GLP‐1 expression in enteroendocrine L‐cells was
gistic effects. In line, combination of ATRA with UDCA attenu- increased in colesevelam treated Mdr2 KO mice, pointing toward
ated biliary fibrosis, bile duct proliferation, as well as hepatic elevated TGR5 activity [78]. Enteroendocrine L‐cell FXR activ-
inflammation in BDL rats and Mdr2 KO mice [72, 76]. However, ity and TGR5 mediated GLP‐1 secretion appear to be inversely
in a clinical pilot study in PSC combination of ATRA and related [18]. Resin‐bound BAs may activate colonic TGR5 but
UDCA did not achieve the primary endpoint (reduction of alka- not intracellular FXR [80], while non‐bound BAs signal intracel-
line phosphatase) despite reduction of the bile acid intermediate lularly via FXR [18]. This may have critical implications on
C4 and inflammation [77]. GLP‐1 secretion (stimulated by TGR5, but suppressed by FXR).
In addition to the differences in GLP‐1 signaling between ASBT
inhibitors and BA sequestrants hepatic and biliary BA composi-
TRANSPORTER MODULATION tion also varies. Interestingly, colesevelam (but not ASBT inhibi-
tor) resulted in a “hydrophilization” (thereby detoxification) of
Since the enterohepatic circulation of BAs depends on active BA the biliary BA composition [18, 78]. Potential mechanistic dif-
transport systems in liver and intestine, inhibition or stimulation ferences between these two treatment strategies may deserve
of their transport function not only alters the BA load to the liver ­further clinical evaluation.
but also BA signaling. Key transporter targets include the apical Several ASBT inhibitors have recently been investigated in
sodium dependent bile salt transporter (ASBT/SLC10A2) for early phase clinical trials for treatment of pruritus in PBC and a
intestinal reabsorption, the sodium taurocholate cotransporting range of pediatric cholestatic syndromes. While pharmacologi-
polypeptide (NTCP/SLC10A1) for hepatocellular reuptake of cal target engagement such as reduced serum BA levels, reduced
conjugated BAs and organic anion transporter (OATPs/SLCO/ FGF‐19 and increased levels of C4 (as marker of BA synthesis)
SCL21) mainly for unconjugated BAs, and perhaps under certain was mostly observed, the clinical effects on pruritus were rather
circumstances also bile salt export pump (BSEP/ABCB11) as variable (often not superior to placebo) and confounded by GI
rate limiting canalicular BA transport system. side‐effects such as diarrhea [81, 82].

ASBT inhibitors and BA sequestrants NTCP blockers


Interruption of enterohepatic circulation of BAs  –  either Reduction of NTCP function may also reduce intrahepatic
­pharmacologically with ASBT inhibitors or with BA seques- BA  levels, subsequently changing hepatobiliary BA excretion
trants – may be advantageous not only for treating pruritus but and, thereby, also hepatobiliary toxicity. BAs can induce
also the underlying cholestatic liver disease since depleting BAs ­inflammation‐based liver injury by stimulating the expression
from the enterohepatic circulation reduces hepatic and biliary and secretion of inflammatory cytokines in cultured mouse
30:  Pathophysiologic Basis for Alternative Therapies for Cholestasis 371

hepatocytes [83], thereby initiating hepatic neutrophil recruit- together with activation of MDR3 and bicarbonate secretion
ment which contributes to progression of cholestatic liver dis- could theoretically help to protect bile ducts from harmful BA
ease [84]. Pharmacological inhibition of NTCP may reduce concentration during the immediate post‐transplant period [18],
intracellular accumulation of BAs, resulting in amelioration of although this highly controversial hypothesis still needs to be
cholestasis and subsequent hepatocellular driven inflammatory tested. Interestingly, drugs such as cyclosporine A used for
response. Genetic absence of NTCP in mice and humans results immune suppression inhibit BSEP function [89].
in increased levels of unconjugated BAs in blood, but is well
tolerated and no pruritus, fat malabsorption, or liver dysfunction
were observed [18]. Myrcludex B, is a NTCP blocker, a small
Chaperones
peptide inhibitor designed to prevent hepatitis B virus uptake Impaired bile secretory function in cholestasis does not only
which, like BAs, need NTCP to enter hepatocytes [18]. In DDC result from changes in gene expression/transcription but may
(3,5‐diethoxycarbonyl‐1,4‐dihydrocollidine) fed mice as model also be due to post‐transcriptional changes with disturbed cell
for sclerosing cholangitis NTCP inhibition improved serum bio- polarity and impaired targeting and function of transporters to
chemistry as well as hepatic inflammation and fibrosis. This the canalicular (and – to a lesser degree – the basolateral) mem-
protective effect may be based on increased biliary PL/BA ratio, brane [5]. Therefore, stimulation of transporter targeting and/or
making the bile less toxic. Consistent with this hypothesis, no function may also have important therapeutic implications.
beneficial effects of NTCP inhibition were seen in the Mdr2 KO Chaperones are mainly used as therapies for hereditary trans-
mouse model lacking biliary PL secretion to begin with [85]. port defects, but could also be of interest in counteracting
Based on these preclinical findings, combination therapy with acquired cholestasis. They may directly interact with the variant
Myrcludex B (to reduce hepatocellular BA uptake) and PPARα of the ABC transporter proteins such as ABCB11/BSEP or
agonists (upregulation of Mdr2/MDR3) might be of particular ABCC2/MRP2, thereby decreasing its degradation and increas-
interest to counteract biliary toxicity (at least in conditions with ing its folding rate (correctors), or impact indirectly on protein
residual MDR2/3 functionality). Besides Myrcludex B, several synthesis via interaction with endogenous ER protein chaper-
drugs have been identified to prevent hepatitis B virus (HBV) ones (potentiators) [90]. Interestingly, tauroUDCA has been
infection via inhibition of NTCP, however, only some such as shown to enhance the secretory capacity of cholestatic hepatocytes
rosiglitazone, zafirlukast, and sulfasalazine inhibited simultane- by stimulation of exocytosis and insertion of canalicular trans-
ously NTCP‐mediated BA uptake, making them attractive port proteins into the apical membrane via PKC dependent
­therapeutic approaches for cholestasis [86, 87]. mechanisms (Figure 30.2) [91]. In MDCKII cells, 4‐phenylbutyrate
(4‐PBA) was shown to increase cell surface expression and
Potential indirect benefits of BSEP blockage transport capacity not only from BSEP variants but also from
the wild‐type form [91]. Treatment of rats with 4‐PBA enhanced
in mice
expression of wild‐type ABCB11 at the canalicular membrane
Although in humans BSEP deficiency results in the very severe [93]. In patients with ABCB11 mutations (PFIC‐2) 4‐PBA treat-
liver disease such as progressive familial intrahepatic cholesta- ment increased canalicular expression of ABCB11 [94]. Further
sis (PFIC) type 2, mice lacking BSEP are surprisingly protected promising observations are made with ivacaftor, originally
from development of severe cholestatic liver injury [18]. known as a cystic fibrosis transmembrane conductance regula-
Moreover, these mice are also resistant to acquired cholestasis tor (CFTR) potentiator [95] which has been shown to increase
induced either by bile duct ligation or DDC feeding [18]. phosphatidyl concentration in patients carrying certain muta-
Already at baseline, BSEP KO mice exhibit increased ­expression tions in the MDR3 gene [95]. Chaperones targeting ABC
of enzymes Cyp2b10 and Cyp3a11, involved in BA hydroxyla- transporters such as ABCB11/BSEP and ABCB4/MDR2/3
­
tion, resulting in a very hydrophilic, less toxic BA pool compo- could therefore also be of particular interest to improve acquired
sition, mainly consisting of tetrahydroxylated BAs [18]. This cholestasis.
metabolic preconditioning with a  –  potential non‐toxic  –
­hydrophilic BA pool, may protect these animals from develop-
ment of cholestatic liver injury. Thus, tetrahydroxylated BAs or
Tight junction sealers
activation of enzymes involved in formation of these BAs may Hepatocyte tight junctions are crucial for maintaining the struc-
have therapeutic potential to treat cholestatic liver disease. This ture and function of bile canaliculi and osmotic gradients gener-
hypothesis is in line with the fact that liver injury in bile duct‐ ated by active transport ultimately driving bile flow, while bile
ligated mice can be attenuated with agonists for CAR and PXR duct epithelium tight junctions are required to maintain epithe-
(activating enzymes involved in BA hydroxylation) [43]. lial cell polarity and prevent leakage of potentially toxic bile
Under rare circumstances, direct transient blockage of BSEP into the periportal space [96]. Loss of tight junction integrity
export function could even directly protect bile ducts from may result in regurgitation of bile and bile duct injury [96].
BA‐induced injury. This might be relevant in the context of liver Apart from several experimental models, disruption of tight
transplantation, where BSEP expression recovers faster than junctions have also been observed in PBC, PSC, and biliary
MDR3, resulting in less favorable biliary BA/PL ratio espe- atresia [97]. Notably, restoration of abnormal ZO1 staining was
cially when MDR3 variants are present [18]. Altered bile observed in PBC patients treated by UDCA [97]. Sulfasalazine,
­composition after liver transplantation has been associated with an anti‐inflammatory agent which is extensively used in long‐
the development of non‐anastomotic biliary strictures [88]. term therapy for chronic IBD increases expression of (intesti-
In  such exceptional conditions transient inhibition of BSEP nal) tight junctions in vivo [98] and protects against BA‐induced
372 THE LIVER:  TARGETING GUT–LIVER AXIS FOR CHOLESTATIC LIVER DISEASE

TUDCA
NTCP
Hepatocyte
TUDCA
& cAMP εPKC

Wild type

Subapical endosomal
recycling compartement
Golgi

Potentiator
Corrector
WT TLC Mutant
Transcription Translation
Folding

mRNA mRNA canaliculus


Misfolding
Mutant
Nucleus
Corrector Folded
mutant
ER

Figure 30.2  Therapeutic targeting of transporter trafficking and function at the post‐transcriptional level. After post‐translational modification in
the Golgi apparatus, wild‐type (WT, blue boxes) and mutated forms (green boxes) of transporter are shuttled to a subapical recycling compartment.
TauroUDCA, entering the hepatocytes via NTCP, and cAMP enhances the secretory capacity of cholestatic hepatocytes by stimulation of
­exocytosis and insertion of transport proteins into the apical membrane (green arrow) via εPKC‐dependent mechanisms. Conversely, transporter
internalization (back to the subapical recycling compartment) is accelerated by taurolithocholic acid (TLC, red arrows). Correctors can partially
rescue misfolding by improving folding at the ER and delaying the turnover to the plasma membrane with a presently poorly understood m ­ echanism.
Although rescued mutants retain in part their function, they are conformationally unstable and might be eliminated by ubiquitination. Potentiators
(e.g. Ivacaftor) may correct this phenotype.

apoptosis [99]. Therefore, this drug might be an attractive thera- protonated (in particular glycine) conjugated BAs [2]. Taurine
peutic for cholestatic liver diseases, especially for PSC patients conjugated and bi‐norUDCA, (lacking two methyl groups) do
also suffering from IBD and this is also currently being studied not undergo cholehepatic shunt and exert less (Tauro-norUDCA)
in ongoing clinical trials. or no (bi‐norUDCA) effect on biliary bicarbonate secretion
compared to unconjugated norUDCA, emphasizing the unique
therapeutic properties of norUDCA in the Mdr2 KO mouse
model of sclerosing cholangitis [2]. Anti‐inflammatory effects
norUDCA – A CHOLEHEPATIC DRUG of norUDCA were seen on isolated cholangiocytes and mac-
rophages, where it directly interferes with NFκB signaling
24‐norursodeoxycholic acid (norUDCA) is a side‐chain short- [102]. In several mouse models (Mdr2 KO, BDL, NEMO KO,
ened derivative of UDCA lacking a methyl group and confer- Schistosomiasis) norUDCA treatment reduced hepatic infiltra-
ring resistance to taurine and glycine conjugation, facilitating its tion of inflammatory cells [102], indicating that norUDCA may
penetration into cholangiocytes [100]. This results in cholehe- even have direct anti‐inflammatory effects. Moreover, norUDCA
patic shunting between hepatocytes and cholangiocytes (bypass- also exerts antifibrotic and antiproliferative effects [103],
ing the classical enterohepatic circulation), and stimulates including downregulation of cyclins and inhibition of mTOR
biliary bicarbonate secretion, resulting in alkalization of bile signaling with induction of autophagy [18]. Notably a recent
(Figure 30.3). NorUDCA is actively secreted in its anionic form phase 2 study demonstrated improvement of cholestatic liver
(norUDCA−) into the canaliculus. NorUDC is converted to the enzymes independent from the previous exposure and response
protonated form (norUDCA) by accepting a hydrogen cation to UDCA [104], encouraging an ongoing phase 3 study.
(generated by ionization of carbonic acid into hydrogen and
bicarbonate) (Figure  30.3). Protonated norUDCA is then
absorbed passively into cholangiocytes (where it gets deproto-
nated again) subsequently entering periductular capillary plexus TARGETING GUT–LIVER AXIS FOR
and returns back to the sinusoid to be resecreted into the canali- CHOLESTATIC LIVER DISEASE
culus in its anionic form (norUDCA−). Acceptance of the proton
of carbonic acid by norUDCA− results in increased alkalization In healthy conditions an enormous amount of gut‐derived mol-
of the bile by generation of bicarbonate [101]. Alkalization of ecules reach the liver via the portal blood without triggering any
bile by bicarbonate enrichment may be a key protective mecha- inflammatory response. Under pathophysiological conditions,
nism (“bicarbonate umbrella”) against BA toxicity to cholangi- when the intestinal barrier function is defective or in case of
ocytes [3], preventing uncontrolled membrane permeation of imbalanced gut bacterial homeostasis, an altered composition of
30:  Pathophysiologic Basis for Alternative Therapies for Cholestasis 373

BA detoxification Basolateral export

norUDC– norUDC– CYPs MRP3


BAs
SULTs, UGTs
MRP4

Blood
Hepatocyte
norUDC–

H2O + CO2
Kidney
H2CO3
Anti-proliferative norUDC–
Anti-inflammatory H+ + HCO3–
Anti-fibrotic
norUDCA +
HCO3–
norUDC–

AE2 HCO3–
Cholehepatic shunting
Bile Cholangiocyte
HCO3–

Figure 30.3  Therapeutic mechanisms of norUDCA. The anionic form of norUDCA (norUDC−) is secreted into the canaliculus by unknown
mechanisms. norUDC− accepts a hydrogen molecule deriving from ionization of carbonic acid (H2CO3). The acceptance of the proton of H2CO3 by
norUDC− leaves behind a HCO3− in the ductular lumen. NorUDCA is passively absorbed into cholangiocytes and transported to the periductular
capillary plexus, returning to the sinusoids and to being resecreted into the canaliculus, thereby completing a cholehepatic shunt, allowing “ductal
targeting” of anti‐inflammatory, antifibrotic, and antiproliferative effects to injured bile ducts. Through the process of cholehepatic shunting
norUDCA also increases biliary bicarbonate (HCO3−) concentration, thereby stabilizing the bicarbonate umbrella and facilitates BA detoxification
and elimination via basolateral efflux pumps facilitating their subsequent renal excretion. BAs, bile acid.

gut‐derived products or even aberrant gut‐primed lymphocytes range of liver diseases. Several studies using high‐throughput
can reach the liver, where it may elicit or exacerbate hepatic sequencing technology reported reduced diversity and signifi-
inflammation. PBC and PSC are cholestatic autoimmune liver cant shifts in the overall composition of the gut microbiota in
diseases in which the end results are caused by immune‐­ PSC patients and appear to be distinct from IBD in the absence
mediated bile duct injury. Both conditions are frequently associ- of PSC [61]. The role of dysbiosis in pathogenesis of cholangio-
ated with gut inflammation, PSC with IBD and PBC with pathies has also been emphasized by recent reports of a changed
coeliac disease [105], and dysbiosis [106]. This clinical obser- microbiome in PBC [106]. Several animal models demonstrated
vation has stimulated several intriguing pathogenic concepts in that enteric dysbiosis and/or administration of bacterial antigens
which gut commensals, pathogens, and intestinal antigens have can result in hepatobiliary inflammation with features resem-
been implicated in causing bile duct injury, in particular in PSC. bling PSC [61]. Interestingly, a recent GWAS uncovered several
One of the first theories connecting intestinal inflammation with new PSC risk loci related to immune regulation including fuco-
liver injury is passive leakage of proinflammatory microbial syltransferase‐2, known to influence microbiota and affecting
components to the portal circulation and the possibility of an susceptibility to microbial infection [61]. The importance of the
antigenic trigger of microbial origin. A study in rats with small presence of gut microbiota was reflected in aggravation of
bowel bacterial overgrowth described significant hepatic inflam- cholestatic liver phenotype and cholangiocyte senescence in
mation leading to fibrosis due to toxin translocation [107]. On germ free Mdr2 KO mice (as an established model of PSC) [61].
the other side, obstructive cholestasis with absence of intralumi- However, the key question remains whether the alterations in
nal BAs leads to bacterial overgrowth promoting bacterial trans- gut microbiota cause liver and bile duct disease or only appear
location [61]. This observation could be explained by the to be secondary to cholestasis or when the disease has already
well‐known direct antibacterial effects of BAs. Furthermore, it progressed. Since reduced diversity dysbiosis is observed in
has been recently shown that BAs exert also indirect effects of multiple inflammatory and metabolic diseases it is tempting to
microbiota via activation of intestinal FXR mediating the assume that these aspects of observation are secondary to dis-
expression of anti‐inflammatory genes such as Ang1, iNOS, and ease and disease‐related inflammation. Future evidence for the
IL18 [24]. The fact that BAs are able to shape microbiota com- potential cause–effect relationship between the microbiome and
munities by growth inhibition of certain bile sensitive bacteria liver disease may come from interventional studies targeting gut
[108] as well as by anti‐inflammatory effects via FXR activation microbiota.
open opportunities for BA‐based therapies in liver diseases Several absorbable and non‐absorbable antibiotics modulat-
associated with intestinal inflammation. ing microbiota have been tested in PSC [109] and revealed bio-
In addition to intestinal inflammation‐related liver injury chemical improvement, with vancomycin being one of the most
changes in microbiota (dysbiosis) have been described in a wide promising agents [110]. In addition probiotics have been
374 THE LIVER:  CONCLUSIONS AND OUTLOOK

explored but so far not recalled convincing clinical results [111].rationale (indicated by association of PBC and genetic variants
Finally, there has been an increasing interest in fecal microbiota in IL12A and IL12RB2 loci [117]) and good efficacy in other
transplantation which is currently being tested as a treatment diseases such as psoriasis in Crohn’s disease [118].
option for PSC [112]. In small studies including PBC patients with an incomplete
Given the key role of BA in gut microbiome homeostasis an response to UDCA, selective B cell depletion with rituximab
effect of UDCA on the microbiota profile could be suspected. In had limited biochemical efficacy although a significant decrease
an interventional trial in PBC [106] UDCA treatment partially of autoantibody production was observed [119]. In addition,
relieved microbiome dysbiosis after six months [106]. since fatigue in PBC has been thought to be driven by muscle
Interestingly, it was also found that patients with PBC and inad- bioenergetics, abnormality related to AMA was reduced by
equate response to UDCA had increased abundance of disease‐ rituximab treatment; rituximab was also tested for fatigue in
associated genera compared with those with adequate response. PBC, but no effect was observed [120]. Moreover, other
In line it has also been shown that UDCA attenuated disease and biologicals such as abatacept (chimeric CTLA4 protein),
­
normalized microbiota changes in mouse models of colitis [113]. FFP104 (anti‐CD40 antibody), NI0801 (anti‐Cxcl10 antibody),
Homing of gut primed T cells has been implicated in the E6011 (anti‐Cxcl1 antibody), and Etrasimod (S1P receptor
pathogenesis of PSC [61]. Effector T cells activated during colitismodulator) are tested in PBC.
express receptors such as α4β7 integrin for MADCAM and Cenicriviroc (CVC) is a novel antagonist for both C‐C
CCR9 for CCL25, facilitating homing to the gut and from the gut chemokine receptors 2 and 5 (CCR2 and 5) that are expressed in
to the liver [61]. Vedolizumab, a gut‐specific α4β7 integrin‐neu- many inflammatory cells including neutrophils, T cells, and
tralizing monoclonal antibody and CCX282‐B, a small molecule monocytes [121]. Peribiliary‐recruited inflammatory cells may
inhibiting CCR9, have been developed for treatment of IBD. contribute to the pathogenesis of cholangiopathies. CVC treat-
Activation of VAP‐1 stimulates hepatic expression of ment improved cholestatic liver injury in the Mdr2 KO mouse
MADCAM‐1 and CCL25, subsequently initiating the recruit- model of sclerosing cholangitis by reducing macrophage
ment of mucosal T cells to the liver [61]. Recently, serum levels recruitment to the liver [122]. Combining treatments that reduce
of VAP‐1 have been shown to correlate with severity of disease in BA pool size with the one that blocks inflammation may have
PSC patients [61]. Therefore, the drugs that may regulate migra- superior effects in the treatment of cholestasis. Combining CVC
tion of activated gut mucosal lymphocytes to the liver appear to and ATRA in cholestatic models (BDL rats and Mdr2 KO
be an attractive option for treating PSC. Several small studies that
mouse) potentiated the effect of monotherapies, indicating a
tested anti‐TNFα agents on reducing biliary inflammation have multitargeted therapy as an important paradigm for treating
been disappointing [114]. A retrospective study in IBD patients cholestatic liver injury [123]. CVC is currently being tested in
with PSC, who have been treated with vedolizumab, did not have patients with PSC [124].
any beneficial effect on liver enzymes [114]. Given its role in Among the emerging antifibrotic inhibitors αVβ6 and lysyl
mediating α4β7/MADCAM‐1 interactions VAP‐1 inhibition may oxidase homolog 2 (LOXL2) are attractive targets. αVβ6 is
also represent a therapeutic target for counteracting cholestatic expressed during epithelial repair in cholestatic hepatocytes and
liver disease [61]. VAP‐1 antibody BTT1023 is currently being cholangiocytes and has a critical role in activation of TGFb1.
clinically tested in PSC patients [115]. Protective effects of αVβ6 inhibition (STX100) were observed
in the Mdr2 KO mouse model of sclerosing cholangitis as well
as in bile duct ligated animals. The humanized monoclonal
anti‐αVβ6 antibody STX100 is under clinical investigation for
TARGETING LIVER INFLAMMATION idiopathic pulmonary fibrosis and chronic allograft nephropathy
AND FIBROSIS [125]. LOXL2, promoting collagen and elastin cross‐linking
[126], as well as cholangiocellular tight junction permeability
Apart from their detergent and proapoptotic properties, BAs [127], has been shown to be critical for development of fibrosis.
cause liver injury in cholestasis by stimulating cytokine‐­ However, despite promising preclinical findings the clinical
mediated inflammatory response and fibrosis [83]. In this results in simtuzumab (LOXL2 inhibitor) treated PSC patients
­process increased hepatocellular BA levels initiated the event by were negative [128].
releasing proinflammatory cytokines (e.g. by activating Egr‐1)
subsequently attracting inflammatory cells which then lead to
tissue injury. Conclusively, reduction of BA load to the liver
and/or inflammatory response by different treatment strategies CONCLUSIONS AND OUTLOOK
would diminish cholestatic liver injury.
Since PSC and PBC are considered to be immune‐mediated Our progress in understanding the molecular pathophysiology
bile duct diseases, several classic immunosuppressive and newer of bile formation and cholestasis has led to the development
immunomodulatory approaches have been tested over time with of  novel therapeutic options as alternative and second‐line
rather disappointing results in regard to efficacy and/or safety ­therapies to UDCA. While UDCA acts mainly at a post‐­
profile [116]. Newer approaches have targeted IL12 [117, 118] transcriptional level, some of these new alternative therapies
and IL23 [118] which have both been implicated in the develop- target key regulatory transcription factors such as FXR, GR, and
ment of PBC [117]. However, ustekinumab, a monoclonal PPARs. Other candidates include key regulators of adaptive BA
­antibody targeting the shared subunit IL12p40, showed rather metabolic and transport pathways such as the xenobiotic sensors
disappointing effects in PBC [118], despite a good scientific CAR and PXR, the FXR downstream target FGF19, inhibitors
30:  Pathophysiologic Basis for Alternative Therapies for Cholestasis 375

of BA uptake systems within the enterohepatic circulation (e.g. 15. Inagaki, T., Choi, M., Moschetta, A. et  al. Fibroblast growth factor 15
ASBT), or novel BA derivatives (e.g. norUDCA) undergoing functions as an enterohepatic signal to regulate bile acid homeostasis.
­
Cell Metab, 2005;2:217–25.
cholehepatic shunting even bypassing the enterohepatic circula- 16. Lu, T.T., Makishima, M., Repa, J.J. et al. Molecular basis for feedback regu-
tion. Direct immunological approaches with immunosuppres- lation of bile acid synthesis by nuclear receptors. Mol Cell, 2000;6:507–15.
sive or biological agents have so far been rather disappointing in 17. Byun, S., Kim, D.H., Ryerson, D. et al. Postprandial FGF19‐induced phos-
treating cholestatic disorders such as PBC and PSC, but novel phorylation by Src is critical for FXR function in bile acid homeostasis.
Nat Commun, 2018;9:2590.
approaches target the gut–liver axis, proinflammatory mono-
18. Trauner, M., Fuchs, C.D., Halilbasic, E., and Paumgartner, G. New therapeu-
cytes, and integrin‐signaling in fibrosis. Moreover, several of tic concepts in bile acid transport and signaling for management of cholesta-
the novel alternative therapeutic approaches targeting BA trans- sis. Hepatol, 2017;65:1393–404.
port and metabolism exert not only anticholestatic but also pro- 19. D’Aldebert, E., Biyeyeme Bi Mve, M.J., Mergey, M. et al. Bile salts control
found anti‐inflammatory as well as immune‐modulatory actions the antimicrobial peptide cathelicidin through nuclear receptors in the human
biliary epithelium. Gastroenterol, 2009;136:1435–43.
which could be key in treating immune‐mediated cholangiopa-
20. Fickert, P., Fuchsbichler, A., Moustafa, T. et al. Farnesoid X receptor criti-
thies such as PBC/PSC and inflammatory/immune responses cally determines the fibrotic response in mice but is expressed to a low extent
secondary to cholestasis irrespective of their underlying etiol- in human hepatic stellate cells and periductal myofibroblasts. Am J Pathol,
ogy. Finally, the gut–liver axis offers additional opportunities 2009;175:2392–405.
for therapeutic modulation of the microbiome (e.g. in PSC and 21. Schwabl, P., Hambruch, E., Seeland, B.A. et al. The FXR agonist PX20606
ameliorates portal hypertension by targeting vascular remodelling and
IBD). A major challenge will be to clinically test and apply ­sinusoidal dysfunction. J Hepatol, 2017;66:724–33.
these expanding therapeutic opportunities and their combina- 22. Hao, H., Cao, L., Jiang, C. et  al. Farnesoid X receptor regulation of the
tions in an individualized personalized medicine approach. NLRP3 inflammasome underlies cholestasis‐associated sepsis. Cell Metab,
2017;25:856–67.
23. Guo, C., Xie, S., Chi, Z. et al. bile acids control inflammation and metabolic
disorder through inhibition of NLRP3 inflammasome. Immunity,
ACKNOWLEDGMENTS 2016;45:802–16.
24. Inagaki, T., Moschetta, A., Lee, Y.K. et  al. Regulation of antibacterial
defense in the small intestine by the nuclear bile acid receptor. Proc Natl
This work was supported by grants F3517‐B20 and I 2755 from
Acad Sci USA, 2006;103:3920–5.
the Austrian Science Foundation (to MT). 25. Verbeke, L., Farre, R., Verbinnen, B. et al. The FXR agonist obeticholic acid
prevents gut barrier dysfunction and bacterial translocation in cholestatic
rats. Am J Pathol, 2015;185:409–19.
26. Ding, L., Yang, L., Wang, Z., and Huang, W. Bile acid nuclear receptor FXR
REFERENCES and digestive system diseases. Acta Pharm Sin B, 2015;5:135–44.
27. Liu, Y., Binz, J., Numerick, M.J. et al. Hepatoprotection by the farnesoid X
1. Erlinger, S. What is cholestasis in 1985? J Hepatol, 1985;1:687–93. receptor agonist GW4064 in rat models of intra‐ and extrahepatic cholesta-
2. Beuers, U., Trauner, M., Jansen, P., and Poupon, R. New paradigms in the sis. J Clin Invest, 2003;112:1678–87.
treatment of hepatic cholestasis: from UDCA to FXR PXR and beyond. 28. Kuzu, M.A., Kale, I.T., Col, C., Tekeli, A., Tanik, A., and Koksoy, C.
J Hepatol, 2015;62:S25–37. Obstructive jaundice promotes bacterial translocation in humans.
3. Beuers, U., Hohenester, S., de Buy Wenniger, L.J., Kremer, A.E., Jansen, Hepatogastroenterology, 1999;46:2159–64.
P.L., and Elferink, R.P. The biliary HCO3(‐) umbrella: a unifying hypothesis 29. Sinal, C.J., Tohkin, M., Miyata, M., Ward, J.M., Lambert, G., and Gonzalez,
on pathogenetic and therapeutic aspects of fibrosing cholangiopathies. F.J., Targeted disruption of the nuclear receptor FXR/BAR impairs bile acid
Hepatol, 2010;52:1489–96. and lipid homeostasis. Cell, 2000;102:731–44.
4. European Association for the Study of the Liver. EASL clinical practice 30. Wagner, M., Fickert, P., Zollner, G. et  al. Role of farnesoid X receptor in
guidelines: the diagnosis and management of patients with primary biliary determining hepatic ABC transporter expression and liver injury in bile duct‐
cholangitis. J Hepatol, 2017;67:145–72. ligated mice. Gastroenterol, 2003;125:825–38.
5. Trauner, M. and Boyer, J.L. Bile salt transporters: molecular characteriza- 31. Gomez‐Ospina, N., Potter, C.J., Xiao, R. et al. Mutations in the nuclear bile
tion, function, and regulation. Physiol Rev, 2003;83:633–71. acid receptor FXR cause progressive familial intrahepatic cholestasis. Nat
6. Evans, R.M. and Mangelsdorf, D.J. Nuclear receptors, RXR and the big Commun, 2016;7:10713.
bang. Cell, 2014;157:255–66. 32. Nevens, F., Andreone, P., Mazzella, G. et  al. A placebo‐controlled trial
7. Halilbasic, E., Baghdasaryan, A., and Trauner, M. Nuclear receptors as drug of  obeticholic acid in primary biliary cholangitis. N Engl J Med,
targets in cholestatic liver diseases. Clin Liver Dis 2013;17:161–189. 2016;375:631–43.
8. Parks, D.J., Blanchard, S.G., Bledsoe, R.K. et al. Bile acids: natural ligands 33. Hirschfield, G.M., Mason, A., Luketic, V. et al. Efficacy of obeticholic acid
for an orphan nuclear receptor. Science, 1999;284:1365–8. in patients with primary biliary cholangitis and inadequate response to
9. Makishima, M., Okamoto, A.Y., Repa, J.J. et al. Identification of a nuclear ­ursodeoxycholic acid. Gastroenterol, 2015;148:751–61.
receptor for bile acids. Science, 1999;284:1362–5. 34. Kris, V., Kowdley, C.L.B., Levy, C. et al. The AESOP trial: a randomized,
10. Wang, H., Chen, J., Hollister, K., Sowers, L.C., and Forman, B.M. double‐blind, placebo‐controlled, phase 2 study of obeticholic acid in
Endogenous bile acids are ligands for the nuclear receptor FXR/BAR. Mol patients with primary sclerosing cholangitis. Hepatol, 2017;66:1254A.
Cell, 1999;3:543–53. 35. Papazyan, R., Liu, X., Liu, J. et al. FXR activation by obeticholic acid or
11. Thomas, C., Pellicciari, R., Pruzanski, M., Auwerx, J., and Schoonjans, K. nonsteroidal agonists induces a human‐like lipoprotein cholesterol change in
Targeting bile‐acid signalling for metabolic diseases. Nat Rev Drug Discov, mice with humanized chimeric liver. J Lipid Res, 2018;59:982–93.
2008;7:678–93. 36. Reich, M., Spomer, L., Hoehne, J. et al. Expression of the bile acid receptor
12. Massafra, V., Milona, A., Vos, H.R. et  al. Farnesoid X receptor activation TGR5 in livers of PSC patients and Mdr2−/− (Abcb4−/−) mice. J Hepatol,
promotes hepatic amino acid catabolism and ammonium clearance in mice. 2017;66:S555.
Gastroenterol, 2017;152:1462–76. 37. Reich, M., Klindt, C., Deutschmann, K., Spomer, L., Haussinger, D., and
13. Huang, W., Ma, K., Zhang, J. et al. Nuclear receptor‐dependent bile acid sign- Keitel, V. Role of the G protein‐coupled bile acid receptor TGR5 in liver
aling is required for normal liver regeneration. Science, 2006;312:233–6. damage. Dig Dis, 2017;35:235–40.
14. Holt, J.A., Luo, G., Billin, A.N. et al. Definition of a novel growth factor‐ 38. Drafahl, K.A., McAndrew, C.W., Meyer, A.N., Haas, M., and Donoghue,
dependent signal cascade for the suppression of bile acid biosynthesis. D.J. The receptor tyrosine kinase FGFR4 negatively regulates NF‐kappaB
Genes Dev, 2003;17:1581–91. signaling. PLoS One, 2010;5:e14412.
376 THE LIVER:  REFERENCES

39. Modica, S., Petruzzelli, M., Bellafante, E. et  al. Selective activation of 61. Hov, J.R. and Karlsen, T.H. The microbiome in primary sclerosing cholangitis:
nuclear bile acid receptor FXR in the intestine protects mice against choles- current evidence and potential concepts. Semin Liver Dis, 2017;37:314–31.
tasis. Gastroenterol, 2012;142:355–65. 62. Haughton, E.L., Tucker, S.J., Marek, C.J. et al. Pregnane X receptor activa-
40. Hirschfield, G.M., Chazouilleres, O., Drenth, J.P. et al. Effect of NGM282, a tors inhibit human hepatic stellate cell transdifferentiation in vitro.
FGF19 analogue, in primary sclerosing cholangitis: a multicentre, rand- Gastroenterology, 2006;131:194–209.
omized, double‐blind, placebo‐controlled phase 2 trial. J Hepatol, 63. Bachs, L., Pares, A., Elena, M., Piera, C., and Rodes, J. Comparison of
2019;70(3):483–93. rifampicin with phenobarbitone for treatment of pruritus in biliary cirrhosis.
41. Zweers, S.J., Booij, K.A., Komuta, M. et  al. The human gallbladder Lancet, 1989;1:574–6.
secretes fibroblast growth factor 19 into bile: towards defining the role of 64. Huang, W., Zhang, J., and Moore, D.D. A traditional herbal medicine
fibroblast growth factor 19 in the enterobiliary tract. Hepatology, enhances bilirubin clearance by activating the nuclear receptor CAR. J Clin
2012;55:575–83. Invest, 2004;113:137–43.
42. Varga, T., Czimmerer, Z., and Nagy, L. PPARs are a unique set of fatty acid 65. Yang, C.Y., Leung, P.S., Adamopoulos, I.E., and Gershwin, M.E. The impli-
regulated transcription factors controlling both lipid metabolism and inflam- cation of vitamin D and autoimmunity: a comprehensive review. Clin Rev
mation. Biochim Biophys Acta, 2011;1812:1007–22. Allergy Immunol, 2013;45:217–26.
43. Zollner, G. and Trauner, M. Nuclear receptors as therapeutic targets in chole- 66. Guo, G.Y., Shi, Y.Q., Wang, L. et  al. Serum vitamin D level is associated
static liver diseases. Br J Pharmacol, 2009;156:7–27. with disease severity and response to ursodeoxycholic acid in primary biliary
44. Honda, A., Ikegami, T., Nakamuta, M. et  al. Anticholestatic effects of cholangitis. Aliment Pharmacol Ther, 2015;42:221–30.
­bezafibrate in patients with primary biliary cholangitis treated with ursode- 67. Levin, A.A., Sturzenbecker, L.J., Kazmer, S. et al. 9‐cis retinoic acid stereoi-
oxycholic acid. Hepatology, 2013;57:1931–41. somer binds and activates the nuclear receptor RXR alpha. Nature,
45. Delerive, P., Gervois, P., Fruchart, J.C., and Staels, B. Induction of 1992;355:359–61.
IkappaBalpha expression as a mechanism contributing to the anti‐inflamma- 68. Idres, N., Marill, J., Flexor, M.A., and Chabot, G.G. Activation of retinoic
tory activities of peroxisome proliferator‐activated receptor‐alpha activators. acid receptor‐dependent transcription by all‐trans‐retinoic acid metabolites
J Biol Chem, 2000;275:36703–7. and isomers. J Biol Chem, 2002;277:31491–8.
46. Gervois, P., Kleemann, R., Pilon, A. et  al. Global suppression of IL‐6‐ 69. Tsuchida, T. and Friedman, S.L. Mechanisms of hepatic stellate cell activa-
induced acute phase response gene expression after chronic in vivo treatment tion. Nat Rev Gastroenterol Hepatol, 2017;14:397–411.
with the peroxisome proliferator‐activated receptor‐alpha activator fenofi- 70. Ohata, M., Lin, M., Satre, M., and Tsukamoto, H. Diminished retinoic acid
brate. J Biol Chem, 2004;279:16154–60. signaling in hepatic stellate cells in cholestatic liver fibrosis. Am J Physiol,
47. Ricote, M. and Glass, C.K. PPARs and molecular mechanisms of transre- 1997;272:G589–96.
pression. Biochim Biophys Acta, 2007;1771:926–35. 71. Sharvit, E., Abramovitch, S., Reif, S., and Bruck, R. Amplified inhibition of
48. Ip, E., Farrell, G.C., Robertson, G., Hall, P., Kirsch, R., and Leclercq, I. stellate cell activation pathways by PPAR‐gamma, RAR and RXR agonists.
Central role of PPARalpha‐dependent hepatic lipid turnover in dietary stea- PLoS One, 2013;8:e76541.
tohepatitis in mice. Hepatology, 2003;38:123–32. 72. He, H., Mennone, A., Boyer, J.L., and Cai, S.Y. Combination of retinoic acid
49. Marra, F., Efsen, E., Romanelli, R.G. et al. Ligands of peroxisome prolifera- and ursodeoxycholic acid attenuates liver injury in bile duct‐ligated rats and
tor‐activated receptor gamma modulate profibrogenic and proinflammatory human hepatic cells. Hepatol, 2011;53:548–57.
actions in hepatic stellate cells. Gastroenterol, 2000;119:466–78. 73. Wang, X., Allen, C., and Ballow, M. Retinoic acid enhances the production
50. Iwaisako, K., Haimerl, M., Paik, Y.H. et al. Protection from liver fibrosis by of IL‐10 while reducing the synthesis of IL‐12 and TNF‐alpha from LPS‐
a peroxisome proliferator‐activated receptor delta agonist. Proc Natl Acad stimulated monocytes/macrophages. J Clin Immunol, 2007;27:193–200.
Sci USA, 2012;109:E1369–76. 74. Crestani, M., Sadeghpour, A., Stroup, D., Galli, G., and Chiang, J.Y. The
51. Nakamuta, M., Fujino, T., Yada, R. et al. Therapeutic effect of bezafibrate opposing effects of retinoic acid and phorbol esters converge to a common
against biliary damage: a study of phospholipid secretion via the PPARalpha‐ response element in the promoter of the rat cholesterol 7 alpha‐hydroxylase
MDR3 pathway. Int J Clin Pharmacol Ther, 2010;48:22–8. gene (CYP7A). Biochem Biophys Res Commun, 1996;225:585–92.
52. Ghonem, N.S., Assis, D.N., and Boyer, J.L. Fibrates and cholestasis. 75. Le Vee, M., Jouan, E., Stieger, B., and Fardel, O. Differential regulation of
Hepatology, 2015;62:635–43. drug transporter expression by all‐trans retinoic acid in hepatoma HepaRG
53. Corpechot, C., Chazouilleres, O., Rousseau, A. et al. A placebo‐controlled cells and human hepatocytes. Eur J Pharm Sci, 2013;48:767–74.
trial of bezafibrate in primary biliary cholangitis. N Engl J Med, 76. Cai, S.Y., Mennone, A., Soroka, C.J., and Boyer, J.L. All‐trans‐retinoic acid
2018;378:2171–81. improves cholestasis in alpha‐naphthylisothiocyanate‐treated rats and
54. Lindor, K.D., Bowlus, C.L., Boyer, J., Levy, C., and Mayo, M. Primary bil- Mdr2−/− mice. J Pharmacol Exp Ther, 2014;349:94–8.
iary cholangitis: 2018 practice guidance from the American Association for 77. Assis, D.N., Abdelghany, O., Cai, S.Y. et al. Combination therapy of all‐trans
the Study of Liver Diseases. Hepatology, 2019;69(1):394–419. retinoic acid with ursodeoxycholic acid in patients with primary sclerosing
55. Jones, D., Boudes, P.F., Swain, M.G. et al. Seladelpar (MBX‐8025), a selec- cholangitis: a human pilot study. J Clin Gastroenterol, 2017;51:e11–16.
tive PPAR‐delta agonist, in patients with primary biliary cholangitis with an 78. Fuchs, C.D., Paumgartner, G., Mlitz, V. et al. Colesevelam attenuates chole-
inadequate response to ursodeoxycholic acid: a double‐blind, randomised, static liver and bile duct injury in Mdr2(−/−) mice by modulating composition,
placebo‐controlled, phase 2, proof‐of‐concept study. Lancet Gastroenterol signalling and excretion of faecal bile acids. Gut, 2018;67:1683–91.
Hepatol, 2017;2:716–26. 79. Whitington, P.F., Freese, D.K., Alonso, E.M., Schwarzenberg, S.J., and
56. Sookoian, S. and Pirola, C.J. Elafibranor for the treatment of NAFLD: one Sharp, H.L. Clinical and biochemical findings in progressive familial intra-
pill, two molecular targets and multiple effects in a complex phenotype. Ann hepatic cholestasis. J Pediatr Gastroenterol Nutr, 1994;18:134–41.
Hepatol, 2016;15:604–9. 80. Harach, T., Pols, T.W., Nomura, M., Maida, A., Watanabe, M., Auwerx, J.,
57. Arenas, F., Hervias, I., Uriz, M., Joplin, R., Prieto, J., and Medina, J.F. and Schoonjans, K. TGR5 potentiates GLP‐1 secretion in response to ani-
Combination of ursodeoxycholic acid and glucocorticoids upregulates onic exchange resins. Sci Rep, 2012;2:430.
the  AE2 alternate promoter in human liver cells. J Clin Invest, 81. Al‐Dury, S. and Marschall, H.U. Ileal bile acid transporter inhibition for the
2008;118:695–709. treatment of chronic constipation, cholestatic pruritus, and NASH. Front
58. Zimmermann, C., van Waterschoot, R.A., Harmsen, S., Maier, A., Gutmann, Pharmacol, 2018;9:931.
H., and Schinkel, A.H. PXR‐mediated induction of human CYP3A4 82. Shneider, B.L., Spino, C., Kamath, B.M. et  al. Placebo‐controlled rand-
and  mouse Cyp3a11 by the glucocorticoid budesonide. Eur J Pharm Sci, omized trial of an intestinal bile salt transport inhibitor for pruritus in
2009;36:565–71. Alagille syndrome. Hepatol Commun, 2018;2:1184–98.
59. Moore, L.B., Parks, D.J., Jones, S.A. et al. Orphan nuclear receptors consti- 83. Cai, S.Y., Ouyang, X., Chen, Y. et  al. Bile acids initiate cholestatic liver
tutive androstane receptor and pregnane X receptor share xenobiotic and injury by triggering a hepatocyte‐specific inflammatory response. JCI
steroid ligands. J Biol Chem, 2000;275:15122–7. Insight, 2017;2:e90780.
60. Huang, W., Zhang, J., Chua, S.S., Qatanani, M., Han, Y., Granata, R., and 84. Saito, J.M. and Maher, J.J. Bile duct ligation in rats induces biliary expres-
Moore, D.D. Induction of bilirubin clearance by the constitutive androstane sion of cytokine‐induced neutrophil chemoattractant. Gastroenterology,
receptor (CAR). Proc Natl Acad Sci USA, 2003;100:4156–61. 2000;118:1157–68.
30:  Pathophysiologic Basis for Alternative Therapies for Cholestasis 377

85. Slijepcevic, D., Roscam Abbing, R.L.P. et al. Na(+) ‐taurocholate cotrans- 107. Lichtman, S.N., Sartor, R.B., Keku, J., and Schwab, J.H. Hepatic inflam-
porting polypeptide inhibition has hepatoprotective effects in cholestasis in mation in rats with experimental small intestinal bacterial overgrowth.
mice. Hepatol, 2018. doi: 10.1002/hep.29888. Gastroenterology, 1990;98:414–23.
86. Donkers, J.M., Zehnder, B., van Westen, G.J.P. et al. Reduced hepatitis B 108. Watanabe, M., Fukiya, S., and Yokota, A. Comprehensive evaluation of the
and D viral entry using clinically applied drugs as novel inhibitors of the bactericidal activities of free bile acids in the large intestine of humans and
bile acid transporter NTCP. Sci Rep, 2017;7:15307. rodents. J Lipid Res, 2017;58:1143–52.
87. Kaneko, M., Futamura, Y., Tsukuda, S. et al. Chemical array system, a plat- 109. Elfaki, D.A. and Lindor, K.D. Antibiotics for the treatment of primary scle-
form to identify novel hepatitis B virus entry inhibitors targeting sodium rosing cholangitis. Am J Ther, 2011;18:261–5.
taurocholate cotransporting polypeptide. Sci Rep, 2018;8:2769. 110. Tabibian, J.H., Weeding, E., Jorgensen, R.A. et  al. Randomised clinical
88. Buis, C.I., Geuken, E., Visser, D.S. et  al. Altered bile composition after trial: vancomycin or metronidazole in patients with primary sclerosing
liver transplantation is associated with the development of nonanastomotic cholangitis – a pilot study. Aliment Pharmacol Ther, 2013;37:604–12.
biliary strictures. J Hepatol, 2009;50:69–79. 111. Shen, J., Zuo, Z.X., and Mao, A.P. Effect of probiotics on inducing remis-
89. Byrne, J.A., Strautnieks, S.S., Mieli‐Vergani, G., Higgins, C.F., Linton, sion and maintaining therapy in ulcerative colitis, Crohn’s disease, and
K.J., and Thompson, R.J. The human bile salt export pump: characteriza- pouchitis: meta‐analysis of randomized controlled trials. Inflamm Bowel
tion of substrate specificity and identification of inhibitors. Dis, 2014;20:21–35.
Gastroenterology, 2002;123:1649–58. 112. Ali, A.H., Carey, E.J., and Lindor, K.D. The microbiome and primary scle-
90. Martin, G.M., Chen, P.C., Devaraneni, P., and Shyng, S.L., Pharmacological rosing cholangitis. Semin Liver Dis, 2016;36:340–8.
rescue of trafficking‐impaired ATP‐sensitive potassium channels. Front 113. Van den Bossche, L., Hindryckx, P., Devisscher, L. et al. Ursodeoxycholic
Physiol, 2013;4:386. acid and its taurine‐ or glycine‐conjugated species reduce colitogenic dys-
91. Beuers, U., Bilzer, M., Chittattu, A. et al. Tauroursodeoxycholic acid inserts biosis and equally suppress experimental colitis in mice. Appl Environ
the apical conjugate export pump, Mrp2, into canalicular membranes and Microbiol, 2017;83.
stimulates organic anion secretion by protein kinase C‐dependent mecha- 114. Tse, C.S., Loftus, E.V., Jr., Raffals, L.E., Gossard, A.A., and Lightner, A.L.
nisms in cholestatic rat liver. Hepatology, 2001;33:1206–116. Effects of vedolizumab, adalimumab and infliximab on biliary inflamma-
92. Hayashi, H. and Sugiyama, Y. 4‐phenylbutyrate enhances the cell surface tion in individuals with primary sclerosing cholangitis and inflammatory
expression and the transport capacity of wild‐type and mutated bile salt bowel disease. Aliment Pharmacol Ther, 2018;48:190–5.
export pumps. Hepatology, 2007;45:1506–16. 115. Arndtz, K., Corrigan, M., Rowe, A. et al. Investigating the safety and activity
93. Hayashi, H. and Sugiyama, Y. Short‐chain ubiquitination is associated with of the use of BTT1023 (Timolumab), in the treatment of patients with pri-
the degradation rate of a cell‐surface‐resident bile salt export pump (BSEP/ mary sclerosing cholangitis (BUTEO): a single‐arm, two‐stage, open‐label,
ABCB11). Mol Pharmacol, 2009;75:143–50. multi‐centre, phase II clinical trial protocol. BMJ Open, 2017;7:e015081.
94. Gonzales, E., Grosse, B., Schuller, B. et al. Targeted pharmacotherapy in 116. Mousa, H.S., Carbone, M., Malinverno, F., Ronca, V., Gershwin, M.E., and
progressive familial intrahepatic cholestasis type 2: Evidence for improve- Invernizzi, P. Novel therapeutics for primary biliary cholangitis: toward a
ment of cholestasis with 4‐phenylbutyrate. Hepatology, 2015;62:558–66. disease‐stage‐based approach. Autoimmun Rev, 2016;15:870–6.
95. Vauthier, V., Housset, C., and Falguieres, T. Targeted pharmacotherapies 117. Hirschfield, G.M., Liu, X., Xu, C. et al. Primary biliary cirrhosis associated
for defective ABC transporters. Biochem Pharmacol, 2017;136:1–11. with HLA, IL12A, and IL12RB2 variants. N Engl J Med,
96. Rao, R.K. and Samak, G. Bile duct epithelial tight junctions and barrier 2009;360:2544–55.
function. Tissue Barriers, 2013;1:e25718. 118. Mousa, H.S., Lleo, A., Invernizzi, P., Bowlus, C.L., and Gershwin, M.E.
97. Sakisaka, S., Kawaguchi, T., Taniguchi, E. et al. Alterations in tight junc- Advances in pharmacotherapy for primary biliary cholangitis. Expert Opin
tions differ between primary biliary cholangitis and primary sclerosing Pharmacother, 2015;16:633–43.
cholangitis. Hepatology, 2001;33:1460–8. 119. Myers, R.P., Swain, M.G., Lee, S.S., Shaheen, A.A., and Burak, K.W. B‐
98. Xu, B., Li, Y.L., Xu, M. et  al. Geniposide ameliorates TNBS‐induced cell depletion with rituximab in patients with primary biliary cholangitis
experimental colitis in rats via reducing inflammatory cytokine release and refractory to ursodeoxycholic acid. Am J Gastroenterol, 2013;108:933–41.
restoring impaired intestinal barrier function. Acta Pharmacol Sin, 120. Khanna, A., Jopson, L., Howel, D. et al. Rituximab is ineffective for treat-
2017;38:688–98. ment of fatigue in primary biliary cholangitis: a phase 2 randomized con-
99. Rust, C., Bauchmuller, K., Bernt, C. et al. Sulfasalazine reduces bile acid trolled trial. Hepatology, 2018.
induced apoptosis in human hepatoma cells and perfused rat livers. Gut, 121. Mack, M., Cihak, J., Simonis, C. et al. Expression and characterization of
2006;55:719–27. the chemokine receptors CCR2 and CCR5 in mice. J Immunol,
100. Hofmann, A.F. and Hagey, L.R., Key discoveries in bile acid chemistry and 2001;166:4697–4704.
biology and their clinical applications: history of the last eight decades. J 122. Guicciardi, M.E., Trussoni, C.E., Krishnan, A. et al. Macrophages contrib-
Lipid Res, 2014;55:1553–95. ute to the pathogenesis of sclerosing cholangitis in mice. J Hepatol,
101. Yoon, Y.B., Hagey, L.R., Hofmann, A.F., Gurantz, D., Michelotti, E.L., and 2018;69:676–86.
Steinbach, J.H. Effect of side‐chain shortening on the physiologic proper- 123. Yu, D., Cai, S.Y., Mennone, A., Vig, P., and Boyer, J.L. Cenicriviroc, a
ties of bile acids: hepatic transport and effect on biliary secretion of 23‐nor‐ cytokine receptor antagonist, potentiates all‐trans retinoic acid in reducing
ursodeoxycholate in rodents. Gastroenterology, 1986;90:837–52. liver injury in cholestatic rodents. Liver Int, 2018;38:1128–38.
102. Halilbasic, E., Steinacher, D., and Trauner, M. Nor‐ursodeoxycholic acid as 124. Bennett, L.D., Fox, J.M., and Signoret, N. Mechanisms regulating
a novel therapeutic approach for cholestatic and metabolic liver diseases. chemokine receptor activity. Immunology, 2011;134:246–56.
Dig Dis, 2017;35:288–92. 125. Dyson, J.K., Hirschfield, G.M., Adams, D.H. et al. Novel therapeutic tar-
103. Fickert, P., Wagner, M., Marschall, H.U. et al. 24‐norUrsodeoxycholic acid gets in primary biliary cholangitis. Nat Rev Gastroenterol Hepatol,
is superior to ursodeoxycholic acid in the treatment of sclerosing cholangi- 2015;12:147–58.
tis in Mdr2 (Abcb4) knockout mice. Gastroenterology, 2006;130:465–81. 126. Barry‐Hamilton, V., Spangler, R., Marshall, D. et al. Allosteric inhibition of
104. Fickert, P., Hirschfield, G.M., Denk, G. et  al. norUrsodeoxycholic acid lysyl oxidase‐like‐2 impedes the development of a pathologic microenvi-
improves cholestasis in primary sclerosing cholangitis. J Hepatol, ronment. Nat Med, 2010;16:1009–17.
2017;67:549–58. 127. Pollheimer, M.J., Racedo, S., Mikels‐Vigdal, A. et al. Lysyl oxidase‐like
105. Trivedi, P.J. and Adams, D.H. Mucosal immunity in liver autoimmunity: a protein 2 (LOXL2) modulates barrier function in cholangiocytes in choles-
comprehensive review. J Autoimmun, 2013;46:97–111. tasis. J Hepatol, 2018;69:368–77.
106. Tang, R., Wei, Y., Li Y. et  al. Gut microbial profile is altered in primary 128. Muir, A.J., Levy, C., Janssen, H.L.A. et al. Simtuzumab for primary scle-
biliary cholangitis and partially restored after UDCA therapy. Gut, rosing cholangitis: phase 2 study results with insights on the natural history
2018;67:534–41. of the disease. Hepatology, 2019;69(2):684–98.
Adaptive Regulation
31 of Hepatocyte Transporters
in Cholestasis
James L. Boyer
Department of Internal Medicine and Liver Center, Yale University School of Medicine, New Haven, CT, USA

INTRODUCTION II: Conjugating enzyme reactions, for example, sulfation, glucu-


ronidation, and amidation, and Phase III: Hepatic efflux mecha-
Bile secretion is a unique and vital function of the liver. It deliv- nisms. Each of these steps is regulated by a series of transporters
ers bile acids to the intestine for the solubilization and absorption (Phase 0 and III) or enzyme reactions (Phase I and II).
of dietary lipids and serves as an excretory pathway for endoge- This chapter focuses on the molecular adaptations in these
nous metabolic products including cholesterol, bilirubin, and hepatocyte membrane transporters and enzymatic reactions that
porphyrins as well as foreign compounds and xenobiotics. Many serve to attenuate cholestatic liver injury and its systemic
liver disorders can impair this secretion, resulting in retention of effects. Cellular events that impair signal transduction path-
bile and the syndrome of cholestasis. In turn, cholestatic liver ways, cytoskeleton structures, tight junction and gap junctional
injury results in adaptive responses in the expression of mem- proteins, and the targeting of intracellular vesicles to the apical
brane transport proteins and hepatic enzymes that synthesize and canalicular domain that maintains the secretory polarity of the
excrete bile acids, bilirubin, and other solutes in an attempt to hepatocyte all contribute to the cholestatic phenotype [1] but are
reduce the hepatic accumulation of these products and thus not considered here.
attenuate liver tissue injury. Ultimately these adaptive responses Much information has been obtained from animal models of
are not sufficient and chronic cholestatic injury leads to the cholestasis and from discoveries of the genetic basis of several
development of biliary cirrhosis and progressive liver failure. hereditary and acquired human cholestatic disorders. Studies in
Developing therapeutic strategies that might enhance these human hepatocyte cell lines and cholestatic liver disorders, as
intrinsic adaptations is an unmet medical need. Here we review well as mouse knockout models have contributed significantly
current knowledge of the mechanism of these adaptive responses. to advances in this field.
The major membrane transport systems that result in the forma-
tion of bile are illustrated in Figure 31.1. Their gene designations,
membrane locations, and primary functions are briefly summa-
rized in Table 31.1. Hepatic enzymes that are involved in the regu- GENERAL OVERVIEW – ACQUIRED
lation of bile salt synthesis and metabolism are summarized in DEFECTS IN BILE TRANSPORT
Table  31.2. A more detailed set of references may be found in PROTEINS
Chapter  23 of the earlier 5th edition. My apologies to authors
whose articles were not able to be included in this 6th edition. The molecular characterization of hepatic transport systems led
to analyses of their response to cholestatic liver injury resulting
in the following paradigm: Determinants of bile formation
Steps in bile formation adapt to cholestatic liver injury in order to minimize hepatic
The overall process of hepatic bile secretion can be divided injury by: (i) diminishing the hepatic uptake of bile salts and
functionally into four different phases: Phase 0: Hepatic uptake other solutes, (ii) reducing bile acid synthesis, (iii) augmenting
mechanisms, Phase I: Hydroxylating enzyme reactions, Phase bile acid detoxification mechanisms, and (iv) upregulating

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
31:  Adaptive Regulation of Hepatocyte Transporters in Cholestasis 379

BSEP
BA FIC1 MRP4
NTCP PS
Na+ BA-S–
BA– MRP2
BA–, OA– OATPs OA– MDR1
OC+ GSH MRP3
OC+
GSH Bil-G–
HCO3– PL
BCRP OA–S
MDR3
OCT1 H+,Na+
OC+
OSTα-β
Chol
OATs ABCG5/8
OA– OC+
? MATE-1

Figure 31.1  Membrane transporters that determine the uptake and excretion of bile acids and other organic solutes in hepatocytes (see tables for
full terminology).

Table 31.1  Nomenclature, location, and function of the major hepatocyte plasma membrane bile acid and organic solute transporters involved
in bile secretion (Phase 0 and III)
Name Abbreviation (gene) Phase Location Function
Sodium‐taurocholate NTCP (SLC10A1) 0 Basolateral membrane of Primary carrier for conjugated bile salt
cotransporter polypeptide hepatocytes uptake from portal blood
Organic anion‐transporting OATPs (SLCO1B1,1B3 0 and III Basolateral membrane of Broad substrate carriers for sodium‐
polypeptides and 3A1) hepatocytes independent uptake of bile salts, organic
anions, and other amphipathic organic
solutes from portal blood. OATP3A1 is
induced in cholestasis
Organic solute transporter OSTα/β (SLC51A and III Basolateral membrane of Heteromeric solute carrier for facilitated
alpha/beta SLC51B) hepatocytes, cholangiocytes, transport of bile acids across basolateral
ileum, and proximal tubule membrane of ileum. Expression induced in
of kidney liver in cholestasis
Organic cation OCT‐1 (SLC22A1) 0 Basolateral membrane of Facilitates sodium‐independent hepatic
transporter‐1 hepatocytes uptake of small organic cations
Organic anion transporter OAT‐2 (SLC22A7) 0 Basolateral membrane of Facilitates sodium‐independent hepatic
2 hepatocytes uptake of drugs and prostaglandins
Multidrug‐resistance MDR1 (ABCB1) III Canalicular and ATP‐dependent excretion of various organic
protein-1 cholangiocyte apical cations, xenobiotics, and cytotoxins into
(P‐glycoprotein)a membrane bile; barrier function in cholangiocytes
Multidrug‐resistance MDR3 (ABCB4) III Canalicular membrane ATP‐dependent translocation of
protein-3 phosphatidylcholine from inner to outer
(phospholipid transporter)a leaflet of membrane bilayer (a floppase)
Bile salt export pumpa BSEP (ABCB11) III Canalicular membrane ATP‐dependent bile salt transport into bile;
stimulates bile salt‐dependent bile flow
Multidrug‐resistance‐ MRP2 (ABCC2) III Canalicular membrane Mediates ATP‐dependent multispecific
associated protein 2 organic anion transport (e.g. bilirubin
(canalicular multispecific diglucuronide) into bile; contributes to bile
organic anion transporter)a salt‐independent bile flow by GSH transport
Multidrug‐resistance‐ MRP3 (ABCC3) III Basolateral membrane of Expression induced in cholestasis.
associated protein 3a hepatocytes and Transports bilirubin and bile acid
cholangiocytes glucuronide conjugates
Multidrug‐resistance‐ MRP4 (ABCC4) III Basolateral membrane of Expression induced in
associated protein 4a hepatocyte; apical membrane cholestasis – transports sulfated bile acid
of proximal tubule of kidney conjugates and cyclic nucleotides
Multidrug‐resistance‐ MRP6 (ABCC6) III Basolateral membrane of ATP‐dependent transport of organic anions
associated protein‐6a hepatocyte and small peptides. Mutations of MRP6
gene result in pseudoxanthoma elasticum
Breast cancer resistance BRCP (ABCG2) III Canalicular membrane, ATP‐dependent multispecific drug
proteina proximal tubule of kidney transporter, particularly sulfate conjugates;
protoporphyrins are endogenous substrates.
Substrates overlap with MRP2
Sterolin‐1 and 2a ABCG5/G8 III Canalicular membrane and Heteromeric ATP dependent transporter for
apical membrane of intestine cholesterol and plant sterols (stilbestrol)
Multidrug and toxin MATE‐1 (SLC47A1) III Canalicular membrane and Organic cation/H+ exchanger extrudes
extrusion protein 1 brush border of kidney cationic xenobiotics
a
 These transporters are members of the ATP‐binding cassette superfamily.
380 THE LIVER: GENERAL OVERVIEW – ACQUIRED DEFECTS IN BILE TRANSPORT PROTEINS

Table 31.2  Hepatic enzymes involved in the regulation of bile salt synthesis and bile salt and bilirubin metabolism (Phase I and II) and their
adaptive responses to cholestasis
Abbreviation
Name (gene) Function Adaptive response in cholestasisa
Cytochrome P450 7A1 (CYP7A1) Cholesterol 7α‐hydroxylase, the rate‐limiting step in Downregulation limits the synthesis
bile acid synthesis from cholesterol of bile acids
Cytochrome P450 8B1 (CYP8B1) Sterol 12α‐hydroxylase, the pathway to cholic acid More favored pathway resulting in
synthesis. Controls ratio of cholic acid to increased % of bile acids as cholic
chenodeoxycholic acid acid
Cytochrome P450 27A1 (CYP27A1) Sterol 27‐hydroxylase, the mitochondrial “acidic” Unchanged
alternative pathway of bile acid side‐chain oxidation.
Favors chenodeoxycholic acid formation
Cytochrome P450 3A4 (CYP3A4) Mediates Phase I bile acid hydroxylation Unchanged or stimulated; may
facilitate ability of bile acids to
undergo Phase II conjugations
Sulfotransferase SULT2A1 Mediates Phase II conjugation with sulfate Unchanged or stimulated;
facilitates bile acid conjugation to
substrates for MRP2 and MRP4
Uridine glucuronyl (UGT1A1/ Mediates Phase II conjugation with glucuronic acid Unchanged or stimulated: mediates
transferase B4/B7) bile acid and bilirubin conjugation
to substrates for MRP2 and MRP3
Bile acid CoA synthetase BACS Forms bile acid CoA‐thioesters, substrates for BAAT —a
Bile acid CoA:amino BAAT/Bat Mediates taurine and glycine conjugation of bile —a; reduced in sepsis
acid N‐acyltransferase acids (amidation)
a
 The adaptive response in cholestasis is not known.

mechanisms that facilitate the export of bile salts and other Sulfated bile salt conjugates in particular are formed in the
toxic substances from hepatocytes [2]. Thus, transport proteins cholestatic human liver albeit to a lesser extent in mouse liver.
on the basolateral sinusoidal membrane which normally func- OSTα‐OSTβ, which mirrors the tissue expression of ASBT, also
tion to remove bile salts and other cholephiles selectively from plays a major role in bile acid efflux from the cholestatic hepat-
portal blood are usually downregulated during cholestatic liver ocyte in human liver [11]. Downregulation of Asbt in the renal
injury by both transcriptional and post‐transcriptional mecha- proximal tubule results in diminished absorption of bile acids
nisms [3, 4]. In contrast, some canalicular transport proteins, from the glomerular filtrate, thereby facilitating bile salts excre-
particularly the multidrug resistance protein (MDR) homologs, tion in the urine [17]. Mrp2 and Mrp4 (expressed on the apical
are either not severely impaired or may actually be upregulated; luminal membrane of the proximal tubule) may also facilitate
these include MDR1/Mdr1, BSEP/Bsep (bile salt export pump), the tubular excretion of bile acid and other divalent conjugates.
and MDR3/Mdr2. These later findings suggest that the choles- Less is known about the response of intestinal bile acid transport
tatic hepatocyte attempts to maintain canalicular export func- proteins in cholestasis, although ASBT is generally downregu-
tion. Perhaps more importantly, transporters such as MRP3/ lated on the luminal brush border of the ileum, reducing the
Mrp3, MRP4/Mrp4 (5–10) and OSTα‐OSTβ (the organic solute return of bile acids to the liver in the enterohepatic circulation
transporter) [11] that are expressed at the basolateral sinusoidal [18], and intestinal MRP2/Mrp2 is diminished in both humans
membrane at low levels in normal liver are substantially upregu- and rats with obstructive cholestasis [19]. Adaptations also
lated in cholestatic hepatocytes, facilitating the efflux of hydro- occur in hepatic enzymes that determine the synthesis and
phobic bile salts and other products back to the circulation metabolism of bile acids. In general, these changes result in a
where they may be cleared in part by the kidney [12]. Thus, smaller and less hydrophobic bile salt pool, although significant
cholestasis results in a partial reversal of bile secretory polarity. species differences exist in composition and response. For
Less is known concerning molecular responses in cholangio- example, in cholestatic mice, bile acid pools are enriched in the
cytes during cholestasis. However, bile duct proliferation is highly hydrophilic bile acid, muricholic acid. Enzymes that
characteristic of most cholestatic disorders. The apical sodium‐ regulate bile acid and bilirubin conjugation are also significantly
dependent bile salt transporter (ASBT), originally identified upregulated in cholestasis. The major enzymes and their func-
in  the ileum, is also expressed on the luminal membranes of tion and adaptive responses are summarized in Table 31.1.
cholangiocytes and the proximal tubules of the kidney and func- Thus, a complex pattern of adaptive responses in transporter
tions to remove bile salts from bile and glomerular filtrate, expression and metabolism occurs in hepatocytes, cholangio-
respectively [13]. Because the cholestatic hepatocyte continues cytes, kidney, and intestine, that attempt to mitigate tissue dam-
to excrete bile salts, albeit at a reduced rate, ASBT may remove age from the retention of bile salts and other toxic substrates.
bile salts from an obstructed biliary tree. MRP3, MRP4, Much of this adaptive regulatory response is mediated by
OSTα‐β, and OATP3A1 are also located on the blood side of the transcriptional events that are regulated by liver‐enriched tran-
cholangiocyte and are upregulated in the cholestatic liver [5, 11, scription factors (hepatocyte nuclear factors [HNFs]) and
14]. MRP3/Mrp3 has a high affinity for glucuronidated conju- nuclear receptors (NRs) [4, 2]. Members of the HNF family
gates and may be the means by which bilirubin glucuronide is tend to regulate constitutive expression of genes whereas NRs
effluxed back into the blood in cholestatic liver. MRP4/Mrp4 induce adaptive responses in gene expression and are activated
may provide the same role for sulfated conjugates [15, 16]. by specific ligands. In cholestasis, these ligands are bile acids,
31:  Adaptive Regulation of Hepatocyte Transporters in Cholestasis 381

bilirubin, oxysterols, and drugs or xenobiotics [21]. Chief nucleus into the cytoplasm, where it may be degraded, thus
among these are bile acids which regulate gene expression pri- influencing the expression of genes dependent on its NR part-
marily via the farnesoid X receptor (FXR). When bile acids ners [25]. Reductions in mRNA levels of FXR, and its target
accumulate in the cholestatic liver, the expression of genes that gene SHP (short heterodimeric partner), have also been reported
are regulated by FXR are usually enhanced. Other NRs that play in patients with cholestasis from PFIC1 and 2 (progressive
an important role in regulating gene expression are the pregnane familial intrahepatic cholestasis type 1 and type 2) and biliary
X receptor (PXR), the constitutive androstane receptor (CAR), atresia [26, 27]. Reductions in PXR and CAR have also been
the liver X receptor (LXR) [22], the vitamin D receptor (VDR), reported in late stages of biliary atresia [27]. PPAR agonists
and peroxisome proliferator‐activated receptor alpha (PPARα) (fibrates) stimulate MDR3, inhibit bile acid synthesis, have anti‐
[23]. Knockout animals for FXR, PXR, CAR, and LXR are each inflammatory effects [28], and improve liver function in patients
more susceptible to cholestatic injury than their respective wild‐ with primary biliary cholangitis (PBC) [29, 30].
type [4, 22–24]. Other ligand stimulated NRs involved in the Several other regulators of transcription for which no specific
adaptive response to cholestasis include the retinoic X receptor ligands are known, but that also influence the expression of bile
alpha (RXRα) and the glucocorticoid receptor (GR). RXRα transporters and enzymes, include the small heterodimer part-
plays a particularly important role in these responses since it is ner‐1 (SHP‐1), liver receptor homologue‐1 (LRH‐1), and
the obligate heterodimeric partner for the class II low‐affinity HNF‐4α. HNF‐4α is a primary determinant of HNF‐1α expres-
NRs that include FXR, PXR, LXR, CAR, and PPARα. sion, which determines the transcription of genes that encode for
Cholestasis also can affect the expression of NRs. Studies of cytochrome P4507A (CYP7A), NTCP, and OATPs (organic
inflammatory and fibrotic models of cholestasis indicate that anion transporting polypeptides) as examples  [31]. Table  31.3
RXRα expression is reduced and may be translocated out of the lists these NRs and their activating ligands and key transcription

Table 31.3  Nuclear receptors, important ligands, transcription factors, and target genes regulating the adaptive response in cholestasis
(? = from rodents only)a
Nuclear receptors Ligands Major target genes Anticipated function in cholestasis
FXR (farnesoid X Bile acids, GW4064, BSEP, OATP1B3, MRP2, MDR3, Inhibition of bile acid synthesis and ileal
receptor) – NR1H4 6α‐ethyl‐CDCA, OSTα/β, CYP3A4, UGT2B4 and bile acid uptake via SHP and FGF‐15/19.
fexaramines B7, SULT2A1, BACS and BAAT, Upregulation of Phase I and II bile acid
SHP, I‐BABP, FGF‐15/19, PXR hydroxylation and conjugation, induction
of bile acid canalicular and basolateral
transporters
PXR (pregnane X Xenobiotics, LCA, CYP3A4, OATP1B1?, SULT2A1, Induction of Phase I and II bile acid and
receptor) – NRII ursodeoxycholic acid, UGT1A1 MRP2, MRP3, MDR1, bilirubin conjugation reactions and bile
rifampicin GST acid and bilirubin alternative export pumps
CAR (constitutive androstane Xenobiotics, bilirubin, CYP2B, CYP3A4, OATP1B1, Induction of Phase I and II bile acid and
receptor) – NR3 phenobarbital, Yin Chin, MRP2, MRP4, UGT1A1, bilirubin conjugation reactions and bile
TCPOBOP SULT2A1, GSTs acid and bilirubin alternative export pumps
LXRα (liver X Oxysterols (metabolites CYP7A1, CYP8B, UGT1A3 Inhibits bile acid synthesis while
receptor) – NR1H3 of cholesterol) ABC5/8?, SHP, OSTα/β?, LRH‐1 stimulating Phase II and III steps
RARα (retinoic acid All‐trans‐retinoic acid NTCP?, MRP2?, ASBT, MRP3 ? Induction of RXR partners via
receptor) – NR1B1 metabolism to 9‐cis‐retinoic acid; loss in
cholestasis upregulates MRP3
VDR (vitamin D Vitamin D, lithocholic CYP3A4, SULTs?, MRP3? Induction of Phase I and II bile acid
receptor) – NRIII acid hydroxylation
GR (glucocorticoid receptor) Corticosteroids NTCP, ASBT, MRP2, BSEP, AE2 ? Induction of AE2 (together with
ursodeoxycholate)
RXRα (retinoic X 9‐cis‐Retinoic acid Obligate heterodimeric partner for Reduction of RXR in cholestasis has
receptor) – NR2B1 all class II NRs variable effects depending on the gene
Others
SHP‐1 small heterodimer None; Upregulated by Interacts with LRH‐1 to suppress Suppresses bile acid synthesis, and hepatic
partner) – NR0B2 FXR CYP7A1, CYP8B1, CYP27A1. and ileal bile acid uptake
Also inhibits NTCP?, and ASBT?,
OATP1B1
LRH‐1(FTF) – liver receptor ? Blocked by SHP CYP7A1, CYP8B1. MRP3, Inhibition of bile acid synthesis (CYP8B1)
homologue‐1; fetal ASBT? and ileal uptake (ASBT?) via SHP;
transcription factor – NR5A2 upregulates MRP3
HNF‐4α (hepatocyte nuclear None HNF‐1, CYP7A1, CYP8B1, Inhibition by SHP via FXR of bile acid
factor‐4) – NR2A1 CYP27A1 OATP1B1, NTCP? synthesis and bile acid uptake
PPARα (peroxisome Fatty acids, eicosanoids, BACS and BAAT SULT2A1, Induces Phase II bile acid and bilirubin
proliferator‐activated fibrates, statins UGT2B4, MDR2/3 conjugation reactions. Inhibits bile acid
receptor‐alpha) – NR1C1 CYP7A synthesis
a
 Transporter and enzyme gene abbreviations include: ABCG5 and G8 (sterolin 1 and 2); ASBT (apical sodium‐dependent bile acid transporter, SLC10A2); BAAT (bile acid
CoA:amino acid N‐acyltransferase); BACS (bile acid CoA‐synthase); BSEP (bile salt export pump, ABCB11); CYP7A (cytochrome P450 7A); CYP8B (cytochrome P450 8B);
CYP27A1 (cytochrome P450 27A1); FGF‐15/19 (fibroblast growth factor 15 or 19); GSTs (glutathione S‐transferases); I‐BABP (ileal bile acid binding protein); MDR1 or 2/3
(multidrug resistance protein 1 or 2/3, ABCB1or 4); MRP2, 3, and 4 (multidrug resistance associated protein 2, 3, or 4, ABCC2,3, or 4); NTCP (sodium taurocholate
cotransporting polypeptide, SLC10A1); OATP‐1B1, 1B3 (organic anion transporting polypeptide C or 8, SLC01B1 or 1B3); OSTα‐OSTβ (organic solute transport protein alpha
and beta); SHP‐1 (small heterodimer partner 1); SULTs (sulfotransferases); UGTs (uridine 5′‐glucuronosyl‐transferases).
382 THE LIVER: GENERAL OVERVIEW – ACQUIRED DEFECTS IN BILE TRANSPORT PROTEINS

factors and illustrates their major target genes and also summa- can be suppressed by bile acids via a small heterodimer part-
rizes the anticipated regulatory responses in cholestasis. Further ner‐dependent ­mechanism [40].
details of these interactions are provided below. However, because In contrast, the rat Ntcp promoter contains binding regions
of differences in gene expression between species, it is difficult to for the transactivators, HNF1, and retinoid receptors (RARα/
generalize. For additional information on adaptive responses in RXRα) [41, 42]. Bile acids downregulate rat Ntcp via FXR‐
animal models of cholestasis and the role of NRs, the reader is dependent Shp expression and subsequent inhibition of retinoid
also referred to several recent comprehensive reviews [4, 32]. activation of RARα/RXRα [43]. Several additional upstream
regions are involved in cytokine responses. Cholestasis pro-
duced by administration of endotoxin (LPS) results in loss of
Regulation of transporter genes HNF1 and the RXRα : RARα heterodimer [42], which decreases
in the cholestatic hepatocyte Ntcp expression. LPS also results in release of cytokines (tumor
necrosis factor alpha [TNF‐α] and IL‐1β) that may contribute to
Ion transporters: (Na+K+‐ATPase and Na+/H+ exchange)
the diminished Ntcp expression [35]. In mice, Ntcp repression
Several major ion transport proteins are located on the basolat- by common bile duct ligation (CBDL) and cholic acid or tauro-
eral membrane and regulate important homeostatic housekeep- cholate feeding is mediated by FXR and does not depend on
ing functions in the hepatocyte, including maintenance of the cytokines, whereas Ntcp repression by LPS is independent of
electrical potential, cell volume, and intracellular pH. These FXR. Although impairment of Ntcp in rodents and NTCP in
transporters include Na+/K+‐ATPase and the Na+/H+ exchanger‐ humans can explain the reduced Na+‐dependent uptake of con-
isoform 1 (NHE‐1). Although many cholestatic agents have jugated bile salts, sodium‐independent mechanisms for hepatic
been shown to inhibit the sodium pump in vitro, the molecular bile salt uptake persist due to continued expression of several
expression of the sodium pump is either not significantly other sinusoidal membrane organic anion transporters such as
affected or somewhat upregulated, as discerned from studies in Oatp2/Oatp1a4 and Oatp4/Oatp1b2 in mice and possibly
several different animal models of cholestasis [33–35]. This OATP1B3 in humans [44, 45].
adaptive response could result from an attempt to counteract
increased sodium entry and cell swelling that may result from Organic anion transporting polypeptides,
the detergent properties of retained bile salts.
OATPs/Oatps (SLCO/Slco)
Na+/H+ exchange is also upregulated at both transcriptional
and post‐transcriptional levels following common bile duct liga- The OATP/Oatps are members of a large super family of
tion in the rat, resulting in an increase in intracellular pH. This transporters (solute carrier organic anion transporter family
response may also contribute to sodium entry and cell swelling [SLCO]) with broad substrate specificity that are capable of
in the cholestatic liver [36]. translocating a wide range of organic anions, including
unconjugated and conjugated bile salts, bulky organic cations,
and even certain uncharged organic substrates [46]. These
Transporters on the basolateral membrane proteins appear to function as anion exchangers, exchanging
of the hepatocyte (phase 0) the extracellular anion with either intracellular bicarbonate or
glutathione [41,42], although the exact driving forces are still
Sodium taurocholate cotransporting polypeptide,
not known. OATPs in human liver consist of OATP1B1
NTCP/Ntcp (SLC10A1/Slc10a1)
(SLCO1B1, ­formally OATP‐C), OATP1B3 (SLCO1B3, for-
Sodium taurocholate cotransporting polypeptide (Ntcp) and its mally OATP8), OATP1A2 (SLCO1A2, formally OATP‐A),
human homolog NTCP is the major determinant of the selective OATP2B1 (SLCO2B1, formally OATP‐B), and OATP3A1.
hepatic uptake of conjugated bile salts from the portal circula- Human OATP1B1 is the most widely studied and the major
tion and is substantially downregulated in both PBC [27, 37], sodium‐independent uptake mechanism for bile acids. Its
biliary atresia [7], PFIC, and animal models of cholestasis [33, expression is reduced in alcoholic hepatitis [37] and PBC,
35, 38]. NTCP mRNA is also reduced in patients with alcoholic particularly in later stages of the disease [47]. Although tran-
hepatitis [37] and other inflammatory conditions. suggesting scriptional mechanisms are not known for most OATPs in
mediation by cytokines. cholestatic patients, in vitro studies indicate that  OATP1B1,
The transcriptional regulation of NTCP/Ntcp differs like NTCP, is regulated by both FXR/SHP‐dependent and ‐
between species [39]. In human tissue, the mechanism is com- independent mechanisms where HNF1α is also a primary
plex as bile acids induce SHP via FXR, which reduces HNF4α transcription factor [33]. In contrast to OATP1B1, studies in
binding to bile acid response elements (BAREs) in the NTCP primary sclerosing cholangitis suggested that OATP1A2,
promoter which, in turn, inhibits its transactivating effects on (formally OATP‐A), mRNA expression is actually increased
HNF1α [31]. HNF1α expression is highly dependent on acti- [44]. However, OATP1A2 is expressed in cholangiocytes in
vation by HNF4α, which is a main regulator of NTCP expres- human liver and not hepatocytes [48]. Bile acids also stimu-
sion [31]. Bile acids also have SHP‐independent effects on late expression of OATP1B3 via FXR, suggesting that
HNF4α binding. However, SHP has no direct effect on NTCP/ OATP1B3 might function in reverse and extrude organic ani-
Ntcp ­promoter activity, and bile acid‐induced signaling path- ons from the cholestatic human hepatocyte. Recently,
ways via c‐Jun N‐terminal kinase may be involved. Other OATP3A1 was found to be upregulated in the liver of patients
studies suggest that the human NTCP gene is also activated by with obstructive cholestasis and in primary human hepato-
the glucocorticoid receptor and PPARγ coactivator‐1α, and cytes via FGF19 activation of transcription factors SP1 and
31:  Adaptive Regulation of Hepatocyte Transporters in Cholestasis 383

nuclear factor‐kappa B where it functions as a bile acid efflux During cholestasis, elevated hepatic levels of bile acids
transporter [14]. a­ ctivate FXR, which induces SHP and inhibits CYP7A1 gene
Several cholestatic animal models have been used to evaluate transcription and bile acid synthesis [53]. In the intestine reduc-
the expression of Oatp1al (Slco1a1, Oatp1), Oatp1a4 (Slc1a4, tions in bile salt concentrations in the intestinal lumen decrease
Oatp2), Oatp1b2 (Slco1b2), and Oatp 4. CBDL, ethinylestradiol production of FGF15/19, an FXR regulated hormone in the
(EE), and LPS all result in downregulation of Oatp1al, although ileum, that interacts with the FGFR4 receptor in hepatocytes in
mRNA levels remain unchanged after EE treatment, suggesting conjunction with β‐klotho [54]. Reductions in FGF15/19 lead to
that the mechanism of EE cholestasis is mainly post‐transcrip- an upregulation of CYP7A1/Cyp7a1. A decline in FGF15/19
tional (see Chapter 23 in the 5th edition for specific references). also reduces gallbladder filling and contractions [55]. FGF19 is
Estrogen‐induced cholestasis results in downregulation of all upregulated in extrahepatic obstruction in human liver, inhibit-
basolateral Oatps. Cytokines, particularly IL‐1β and TNF‐α, ing bile acid synthesis [56]. In addition, proinflammatory
play a major role in the regulation of OATP/Oatps and also other cytokines (TNFα and IL‐1β) also play a role via c‐Jun N‐termi-
bile acid transporters in cholestasis [49]. Basolateral and canali- nal kinase (JNK)/cJun post‐transcriptional signal transduction
cular transporter systems are downregulated by both cytokines. pathways [57].
Decreased binding activities of nuclear receptor heterodimers Thus, multiple transcriptional and post‐transcriptional mech-
may also be explained in part by a reduction of the ubiquitous anisms exist to reduce the formation of bile acids during choles-
heterodimerization partner, RXR‐α. tasis. However, the master switch appears to be at the level
of  the intestine, where bile acid homeostasis is exquisitely
Organic cation transporter (OCT‐1, SLC22a1) ­regulated via FGF15/19.
In rat models of CBDL and α‐naphthyl isothiocyanate
OCT‐1 (organic cation transporter) is the major hepatic uptake (ANIT) administration, Cyp8b1 expression, but not Cyp7a1,
transporter for small organic cations. OCT‐1 expression in was significantly inhibited [58] whereas 17α‐ethinylestradiol‐
human cholestatic liver has not been assessed but animal models induced cholestasis decreased Cyp7a1 but not the acidic path-
of cholestasis (CBDL and LPS) indicate downregulation of way via Cyp27a. In patients with primary biliary cholangitis,
rOct‐1 mRNA and protein and impaired uptake of Oct‐1 sub- CYP7A1 mRNA decreased to 10–20% of control levels with no
strates [50]. The human OCT‐1 gene is transactivated by changes in CYP27A or CYP8B1 [9]. Thus, adaptive changes
HNF‐4α and, like other basolateral hepatic uptake transporters, that serve to reduce bile acid synthesis occur in both rodents and
is suppressed by bile acids via SHP [51]. Thus, hOCT‐1 is also patients with cholestasis.
likely to be downregulated in cholestasis affecting uptake of
substrates like metformin.
Bile acid hydroxylation (phase I)
Organic anion transporter 2 and 3 (OAT‐2 and ‐3, Studies in bile duct ligated rats indicate that most Cyp450‐­mediated
Slc22a6 and Slc22a8) reactions are diminished. In vitro studies show that unconjugated
These two members of the OAT family are expressed in liver hydrophobic bile acids are more potent inhibitors than conjugated
and kidney, and transport a variety of organic anions including bile acids [59]. However, during cholestasis hydroxylation reac-
bile acids. However, the effects of cholestasis on the expression tions mediated predominantly by CYP3A4 attempt to decrease
of these transporters in liver is not well studied. the hydrophobicity of the bile acid pool accounting for the signifi-
cant levels of (poly)hydroxylated bile acids in the urine of chole-
static animals and patients [60]. CYP3A4 is regulated by the NRs
Bile acid synthesis PXR, FXR [60], VDR, and CAR [61]. Bile acids also induce
Cyp3a11 after CBDL in mice [60]. However, CYP3A4 mRNA is
Bile acids are formed from cholesterol in the liver by a series
only mildly altered in patients with PBC [9].
of enzymatic pathways consisting of the classic (neutral) path-
way or the alternative (acidic) pathway [52]. The former is
mediated by CYP7A1 and is rate limiting in the overall pro-
Bile acid conjugation (phase II)
cess. This pathway also involves CYP8B1, which leads to the
production of CA and which determines the relative hydropho- Glucuronidation, sulfation, and amidation of bile acids (major
bicity of bile by determining the ratio of chenodeoxycholic Phase II reactions) also diminish their toxicity during cholesta-
acid (CDCA) to CA that normally is approximately equal. sis and facilitates their excretion back into blood via MRP3 and
CYP27A1 mediates the alternative pathway which leads MRP4 with subsequent elimination by the kidney. Human
­primarily to the formation of CDCA. All bile acid species are ­uridine glucuronyl transferase 1A3 (UGT1A3) forms acyl che-
conjugated (amidated) with either glycine or taurine by bile nodeoxycholic acid 24‐glucuronide. Cytosolic sulfotransferase
acid CoA‐synthase (BACS) and bile acid CoA:amino acid N‐ 2A (SULT2A) also plays a significant role in sulfation reac-
acyltransferase (BAAT). These bile acid conjugates are then tions, particularly in females. This enzyme may be upregulated
excreted into bile via the BSEP and can undergo both deami- by Car‐dependent mechanisms [61] in rodents. In contrast,
dation and 7α‐dehydroxylation by intestinal bacteria, produc- SULT2A is negatively regulated through CDCA‐mediated FXR
ing deoxycholic acid and lithocholic acid. Unconjugated bile activation in mice and humans [62].
acids are reamidated when they recycle back to the liver in the Amidation reactions for bile acids are determined by both
enterohepatic circulation [52]. BACS and BAAT and both are regulated by FXR [63]. Few
384 THE LIVER: GENERAL OVERVIEW – ACQUIRED DEFECTS IN BILE TRANSPORT PROTEINS

studies have examined these enzymes during cholestatic liver hepatic injury is reduced after bile duct obstruction in Ostα
injury. Mutations in BAAT have been described in hypercholane- knockout mice [72].
mia in childhood [64].
Canalicular membrane of the hepatocyte (MDR1/
Mdr1a,b; MDR3/Mdr2; MRP2/Mrp2; BSEP/Bsep;
Hepatic efflux mechanisms (phase III) BCRP/Bcrp; ABCG5/G8;Abcg5/8)
Basolateral membrane of the hepatocyte (the multidrug Despite downregulation of hepatic uptake mechanisms and
resistance associated proteins MRPS/Mrps (ABCC) upregulation of basolateral efflux transporters that function to
and the organic solute transporter, OSTα‐OSTβ retard the accumulation of bile acids and other toxic substrates,
Cholestatic liver injury results in reversal of the secretory polar- hepatic levels of bile acids and other constituents of bile con-
ity of the hepatocyte. This phenomenon is associated with tinue to accumulate in the cholestatic liver and are primary sub-
the  adaptive upregulation of transporters on the basolateral strates for the apical canalicular membrane efflux pumps. These
membrane that facilitates the extrusion of bile salts and other transporters are all members of the ABC superfamily and
organic solutes back into the hepatic sinusoids where they can the  inability of these rate‐limiting transporters to effectively
be eliminated in the urine. excrete these substrates into bile becomes the major determinant
Mrp1, ‐3, and ‐4 and Ostα‐β are normally expressed at low of the cholestatic phenotype.
levels in normal liver [65, 66] but are generally upregulated
during cholestasis [5, 6, 11]. Mrp1, ‐5, and ‐6 are of little Canalicular organic solute transporters
functional importance in the cholestatic liver while Mrp3 and
Multidrug resistance protein, MDR1/Mdr1a,b
‐4 are capable of preferentially transporting glucuronidated
and sulfated bile acids, respectively [15], and their induction (ABCB1/Abcb1)
during cholestasis may account for the appearance of these MDR1 encodes for the drug efflux pump also known as P‐gly-
bile acid conjugates in the urine [12]. Induction of Mrp3 after coprotein 170 (Pgp‐170). MDR1 transports a variety of drugs,
BDL in mice is due to TNF‐α dependent upregulation of xenobiotics and lipids although its importance in excretion of
Lrh‐1 with increased binding of Lrh‐1 to the Mrp3 promoter endogenous substrates remains unclear. Mdr1 is capable of
[67]. Hepatic injury is more severe in bile duct ligated TNF‐ transporting bile acids, albeit with a much lower affinity than
receptor knockout mice [67]. Bsep [73]. Disease causing MDR1 mutations have not been
Studies in HepG2 cells indicate that human MRP3 expression found. However, polymorphisms have significant effects on
is repressed by RXR α : RARα which occupies Sp1 activator sites drug absorption, excretion, and toxicity [74]. Most forms of
in the MRP3 promoter. Because RXRα : RARα expression is cholestasis result in significant upregulation of Pgp‐170
diminished by cholestatic liver injury, loss of RXRα : RARα may Mdr1a/b mRNA in both animal models and humans with the
lead to upregulation of MRP3/Mrp3 expression by derepressing level of expression correlating with the severity of the cholesta-
Sp1 activation in cholestatic disorders [68]. More recent studies sis [75]. MDR1 protein is increased in late‐stage PBC and in
in Mrp4 and Mrp3 knockout mice suggest that Mrp4, rather than biliary atresia [9, 27]. Molecular regulation of MDR1/ Mdr1 is
Mrp3, may be more important in protecting the hepatocytes from thought to be mediated by NF‐κB transcriptional mechanisms
bile acid toxicity [8, 69]. MRP4 protein, but not mRNA, levels are [76]. Both CAR and PXR activators can also stimulate MDR1
significantly increased in patients with stage III and IV primary expression [77].
biliary cholangitis [9], PFIC1 [7] and late‐stage biliary atresia
[27], suggesting post‐transcriptional regulation [9]. Mrp3 in mice
MDR3/Mdr2 (ABCB4/Abcb4)
is regulated by CAR, PXR, and VDR whereas Mrp4 is induced
primarily by CAR and PPARα [70]. The importance of this phospholipid export pump in the patho-
Both CAR and aryl hydrocarbon receptor (AHR) response genesis of cholestasis is dramatically demonstrated by muta-
elements are present in the MRP4 promoter so that bilirubin, a tions in the MDR3/Mdr2 gene that result in PFIC3 in children
CAR activator, as suggested by studies in Car knockout mice [78] and biliary cirrhosis in the mouse knockout model, Mdr2−/−
[70], may stimulate MRP4 expression. There is no evidence that [79]. MDR3/ Mdr2/is a phospholipid floppase. In its absence
other basolateral Mrps expressed in the liver, including Mrp5 phosphatidylcholine cannot be excreted into bile, and bile salts
and ‐6, play any protective role in the adaptive response in cannot form mixed micelles resulting in progressive injury to
cholestatic liver injury. the bile duct epithelium and, in the mouse, histological findings
OSTα‐β/Ostα‐β is a facilitated heteromeric transporter of reminiscent of primary sclerosing cholangitis [80]. Over time a
bile acids and other organic solutes where the direction of biliary‐type cirrhosis develops and, in some cases, hepatocellu-
transport is dependent on the electrochemical gradient lar carcinoma.
between the cell and blood. It is weakly expressed in the Children with PFIC3 also have defective phospholipid excre-
liver in rodents but is significantly upregulated by FXR reg- tion in bile, bile duct proliferation by histological examination,
ulated mechanisms at both mRNA and protein levels primar- and elevated γ‐glutamyl transferase, thus distinguishing them
ily in patients with stage III and IV PBC, biliary atresia, and from PFIC1 and 2, where bile duct proliferation is absent and
in bile duct ligated mice and rats [11, 27]. Ostα‐β is also γ‐glutamyl transferase is normal [81].
regulated by Lxr in the mouse which shares a DR‐1 binding Phospholipid excretion is partially deficient in heterozygotes
site with Fxr in the mouse promoter [71]. Paradoxically who normally do not have a cholestatic phenotype [78, 82].
31:  Adaptive Regulation of Hepatocyte Transporters in Cholestasis 385

However, mothers of PFIC3 patients are obligate heterozygotes expression. MRP2 protein is also markedly reduced in liver
(MDR3+/−) and are at risk for developing cholestasis during the biopsies from patients with inflammatory cholestasis [37].
third trimester of pregnancy when high levels of estrogens are The Mrp2 promoter is activated by RXRα : RARα, and
present [83, 84]. Polymorphisms and mutations in MDR3 can depressed by IL‐1β following bile duct ligation in rats [42].
predispose patients to cholestatic liver injury when exposed to Thus, Mrp2 expression in obstructive cholestasis is associated
other potential cholestatic agents, including drugs and environ- with cytokine‐dependent alterations in the RXRα : RARα NRs.
mental toxins [85]. A missense mutation in ABCB4 has been Response elements for CAR, PXR, FXR, and AHR also are pre-
described in ductopenic cholestatic liver disease in adults. sent in the MRP2/Mrp2 promoter so the molecular regulation of
MDR3‐deficient patients present a wide spectrum of disease this transporter in cholestasis remains complex [4].
including intrahepatic cholestasis of pregnancy, low phospho- Since bilirubin and glutathione are antioxidants, their hepatic
lipid associated cholelithiasis (LPAC), cholestatic cirrhosis, and retention in cholestasis may have cytoprotective effects.
death in childhood and adults [86].
MDR3/Mdr2 is upregulated in most forms of cholestasis,
The bile salt export pump (BSEP/Bsep, ABCB11/Abcb11)
including bile duct obstruction and ANIT [58] and in patients
with biliary atresia [27]. MDR3/Mdr2 is regulated in large part This gene encodes for the “sister of P‐glycoprotein”. the canali-
by Fxr/FXR‐mediated mechanisms [87], as well as PPARs in cular membrane bile salt export pump and the major, if not the
mice and in humans [88]. Trials with fibrates in patients with sole determinant of bile salt dependent bile flow. Mutations in
PBC are based in part on this finding. BSEP result in PFIC2 [97], benign recurrent intrahepatic chol-
estasis (BRIC), and intrahepatic cholestasis of pregnancy in
adults. More than 100 different mutations have been described
MRP2/Mrp2 (ABCC2/Abcc2)
in families with these disorders [97]. PFIC2 resembles the
MRP2 encodes for the conjugate drug export pump, also known Byler’s disease (PFIC1) phenotype, with the absence of bile
as multidrug resistance protein 2 or the canalicular multidrug duct proliferation and normal γ‐glutamyl transferase levels in
organic anion transporter (cMOAT) [65, 66]. This ABC trans- serum. Bsep knockout mice also demonstrate impaired bile salt
porter is the major export pump for bilirubin diglucuronides, transport into bile. Altogether the evidence is compelling that
glutathione conjugates, and divalent bile acids conjugated with BSEP/Bsep is the canalicular transporter that determines bile
sulfates and glucuronides. It is a major determinant of bile acid salt dependent bile formation.
independent bile flow and drug conjugate excretion. Expression Experimental observations suggest that BSEP/Bsep is varia-
varies considerably in human liver [89] and single nucleotide bly preserved during cholestatic liver injury. Bsep is only mod-
polymorphisms (SNPs) affect drug clearance. Mutations in the erately impaired in several rat models of cholestasis [96].
Mrp2 gene result in a stop codon and premature termination of Although a variety of cholestatic agents, including LPS, EE,
protein translocation in the mutant TR–/GY [90] and the Eisai cyclosporin A, and rifamycin, profoundly inhibit ATP depend-
hyperbilirubinemic rat (EHBR) [91]. Mutations in the MRP2 ent bile salt transport in vitro in isolated rat liver canalicular
gene in humans result in the Dubin–Johnson syndrome [92] a membrane vesicles [98], downregulation in vivo is less pro-
genetically determined cause of conjugated hyperbilirubinemia. nounced. After 3 days of CBDL in the rat Bsep mRNA and
Although not cholestatic by definition, since bile salt excretion ­protein expression are inhibited by ~30 and 50%, respectively,
is normal, this mutation results in impaired excretion of a vari- but after 7–14 days recover to ~60 and 80% of control
ety of amphipathic organic anions, including leukotrienes, ­values, respectively. Immunofluorescence studies indicate that
conjugated bilirubin, divalent bile acid conjugates, and
­ the  transporter remains at the canalicular membrane [96].
coproporphyrin isomer series 1, in addition to a variety of other Furthermore, bile salt excretion continues in the face of com-
compounds including bromosulfophthalein (BSP), indocyanine plete bile duct obstruction, albeit at a reduced rate [96] and is
green, and oral cholecystographic agents [65]. Antibiotics, such mediated by TNF‐α and Il‐1β [99]. LPS and EE administration
as ampicillin and ceftriaxone, and heavy metals are also excreted in vivo to rats also results in only partial inhibition of Bsep
by MRP2. Although it is not known if bile secretion is impaired expression [100]. Although reduced staining of BSEP protein is
in the Dubin–Johnson syndrome, bile salt independent bile flow seen in patients with inflammatory cholestasis [37], BSEP
is reduced in the rat model as a result of impaired glutathione expression is maintained in patients with primary biliary chol-
excretion [93]. Glutathione is a low‐affinity substrate for Mrp2 , angitis [47] and is reduced in early but not late stages of biliary
the major pathway for excretion of glutathione, oxidized GSSG atresia [7, 27], where it is maintained in its normal amount and
and glutathione conjugates [94]. Polymorphisms in MRP2 can location [101]. BSEP/Bsep is strongly regulated by FXR/Fxr in
result in diminished glutathione excretion and thus may contrib- human, rat, and mouse [102] and Bsep induction by bile acids is
ute to other forms of toxic and cholestatic liver injury [89, 95]. absent in the FXR knockout mouse [103]. Thus, differences in
Indeed, the mRNA and protein expression levels of Mrp2, are levels of BSEP/Bsep expression in cholestasis may be related in
markedly downregulated in animal models of cholestasis, part to differences in the levels of expression of this nuclear
including CBDL, EE, and particularly following administration receptor. For example, while initially reduced, FXR and BSEP
of LPS [38]. The latter explains the increase in serum conju- levels return to normal in late‐stage cases of biliary atresia [27].
gated bilirubin characteristic of sepsis‐induced jaundice LRH‐1 also can regulate BSEP expression and thus may play a
[38,  96]. Post‐transcriptional events can result in retrieval of supporting role to FXR in maintaining hepatic bile acid levels
Mrp2 from the canalicular membrane to a submembranous and in coordinating the expression of both CYP7A1 and BSEP
localization early in cholestasis prior to changes in mRNA to determine bile acid synthesis and excretion [104]. The nuclear
386 THE LIVER:  REFERENCES

factor erythroid 2‐related factor 2 (NRF2) also transactivates the Canalicular ion transporters (AE2, also SLC4A2)
human BSEP promoter by binding to a Maf recognition element
(MARE). NRF2 is activated by oxidative stress which can result
Anion Exchanger‐2 (AE2)
from the accumulation of “toxic” bile acids [105]. This transporter encodes for the canalicular Cl−/HCO3−
Taurocholate transport is competitively cis‐inhibited by ­various exchanger that regulates the excretion of bicarbonate, a partial
cholestatic drugs, including cyclosporin A, rifamycin, rifamycin determinant of canalicular bile salt independent bile formation
SV, and glibenclamide in Sf9 cells expressing rat Bsep [98]. [114]. Anion exchanger‐2 (AE2) is also expressed on the lumi-
Altogether these findings suggest that the cholestatic effects of nal membrane of the cholangiocyte and is a determinant of
these compounds may be determined in part by the extent to bicarbonate excretion from this epithelium, providing a protec-
which the canalicular export pumps continue to function as export tive “biliary bicarbonate umbrella” [115]. Thus, impairment of
pumps. Interestingly, the cholestatic metabolite estradiol‐17β‐ AE2 would be expected to reduce bile flow and thus might
glucuronide inhibited ATP‐dependent taurocholate transport only ­predispose to other cholestatic insults.
when Mrp2 was co‐expressed in Sf9 cells, suggesting that this Initial studies showed that AE2 gene expression and protein
compound results in trans‐inhibition of Bsep‐mediated bile salt immunoreactivity at the canaliculus and in bile ducts were
transport only after excretion into the bile canaliculi by Mrp2 reduced in patients with PBC but not in other cholestatic and
[98]. Thus, some drugs may produce cholestasis by inhibiting non‐cholestatic liver diseases [116]. Reduced expression of
BSEP only after excretion into bile. Polymorphisms in BSEP, AE2 mRNA has also been observed in salivary glands in patients
particularly V444A, also affect the level of expression of BSEP in with PBC and sicca syndrome, suggesting that there may be a
human liver [89] and may predispose to some forms of drug‐ generalized deficiency of this transporter in this disease.
induced cholestasis and cholestasis of pregnancy [85, 106]. Measurements of the Cl−/HCO3− anion exchanger (AE) in iso-
Mutations of BSEP are associated with cholestatic liver diseases lated cholangiocytes showed that the cAMP‐stimulated AE
of varying severity including PFIC2 and BRIC2. Genetic poly- activity is diminished in PBC compared to both healthy and dis-
morphisms are also linked to intrahepatic cholestasis of preg- eased controls. MicroRNA‐506 (miR‐506) is upregulated in
nancy (ICP) and drug‐induced liver injury [106]. cholangiocytes of PBC patients and AE2 may be a target of
miR‐506 [117]. Stimulation of the activity of AE2 may be one
Other canalicular solute and lipid transporters/flippases of several mechanisms by which ursodeoxycholic acid (UDCA)
(BCRP, ABCG2) and ABCG5/8, MATE‐1, and FIC1) may improve outcomes in PBC patients.

The breast cancer resistance protein, BRCP (ABCG2), is also


expressed on the canalicular membrane. BRCP shares broad sub-
strate specificity with MRP2 and excretes sulfated drug conjugates CONCLUSION
[107]. Bcrp does not appear to have a significant role in the adap-
tive response to cholestasis in the liver but may be more important This chapter has focused on the adaptive response that mem-
for solute export in the kidney and intestine. BRCP is downregu- brane transporters in hepatocytes play in inherited and acquired
lated in the duodenum of patients with obstructive cholestasis but forms of cholestasis and the role of NRs in this process.
not in bile duct ligated rats [108]. Polymorphisms in BRCP are Information about the mechanisms of these transcriptional
speculated to play a role in troglitazone sulfate excretion and could events is rapidly expanding, revealing the complexity and the
possibly result in cholestasis induced by this metabolite. interrelatedness of these responses in the form of extended net-
Sterolin 1 and 2 (ABCG5/8) are two ABC transporters that works. Adaptative mechanisms also occur in cholangiocytes,
form a heterodimer at the canalicular membrane and account in kidney, and intestine that contribute to the ability to modulate
large part for the excretion of cholesterol and plant sterols [109]. this disorder but are beyond the scope of this review. In the
Little is known about its role in cholestasis, although estrogen‐ future, progress will come from new therapeutic strategies, most
induced cholestasis results in diminished mRNA expression [110]. likely combinations of novel nuclear receptor ligands that stim-
Reduced expression of ABCG5/8 would be expected to contribute ulate these protective pathways.
to the hypercholesterolemia that develops during cholestasis.
The multidrug and toxic compound extrusion protein‐1
(MATE‐1) is also expressed in the liver on the apical canalicular
membrane. Members of the MATE family are organic cation ACKNOWLEDGMENTS
exporters that excrete metabolic or xenobiotic organic cations
from the body by way of a H+ or Na+ exchange mechanism Publications cited from this laboratory were supported by
[111]. However, little is known about its specific role in the liver USPHS DK 25636 and DK P30–34989.
and whether it is regulated during cholestasis.
The gene product of familial intrahepatic cholestasis‐1 (FIC1,
ATP8B1) is a P‐type ATPase which functions as a phosphatidyl-
serine flippase at the canalicular membrane [112]. Mutations in REFERENCES
FIC1 result in PFIC‐1, also called Byler’s disease [113]. PFIC‐1
1. Trauner, M. and Boyer, J.L. Bile salt transporters: molecular characterization,
is phenotypically similar to PFIC‐2. Whether FIC1’s function is function and regulation. Physiol Rev, 2003;83:633–71.
impaired in other forms of cholestasis and contributes further to 2. Boyer, J.L. New perspectives for the treatment of cholestasis: lessons from
cell injury is not clear. basic science applied clinically. J Hepatol, 2007;46:365–71.
31:  Adaptive Regulation of Hepatocyte Transporters in Cholestasis 387

 3. Stahl, S., Davies, M.R., Cook, D.I. and Graham, M.J. Nuclear hormone 26. Demeilliers, C., Jacquemin, E., Barbu, V. et al. Altered hepatobiliary gene
receptor‐dependent regulation of hepatic transporters and their role in the expressions in PFIC1: ATP8B1 gene defect is associated with CFTR down-
adaptive response in cholestasis. Xenobiotica, 2008;38:725–77. regulation. Hepatology, 2006;43:1125–34.
 4. Halibasic, E.C.T. and Trauner, M. Bile acid transporters and regulatory 27. Chen, H.L., Liu, Y.J., Chen, H.L. et al. Expression of hepatocyte transporters
nuclear receptors in the liver and beyond. J Hepatol, 2013;58:155–68. and nuclear receptors in children with early and late‐stage biliary atresia.
  5. Soroka, C.J., Lee, J.M., Azzaroli, F. and Boyer, J.L. Cellular localization and Pediatr Res, 2008.
up‐regulation of multidrug resistance‐associated protein 3 in hepatocytes 28. Ghonem, N.S., Assis, D.N., and Boyer, J.L. Fibrates and cholestasis.
and cholangiocytes during obstructive cholestasis in rat liver. Hepatology, Hepatology, 2015;62:635–43.
2001;33:783–91. 29. Levy, C. and Lindor, K.D. Editorial: itching to know: role of fibrates in PBC.
  6. Denk, G.U., Soroka, C.J., Takeyama, Y. et al. Multidrug resistance‐associ- Am J Gastroenterol, 2018;113:56–7.
ated protein 4 is upregulated in liver but downregulated in kidney in obstruc- 30. Levy, C., Peter, J.A., Nelson, D.R. et al. Pilot study: fenofibrate for patients
tive cholestasis in the rat. J Hepatol, 2004;40:585–91. with primary biliary cirrhosis and an incomplete response to ursodeoxy-
  7. Keitel, V., Burdelski, M., Waqrskulat, U. et al. Expression and localization of cholic acid. Aliment Pharmacol Ther, 2011;33:235–42.
hepatobiliary transport proteins in progressive familial intrahepatic cholesta- 31. Jung, D. and Kullak‐Ublick, G.‐A. Hepatocyte nuclear factor 1a: a key
sis. Hepatology, 2005;41:1160–72. ­mediator of the effect of bile acids on gene expression. Hepatology, 2003;37
  8. Mennone, A., Soroka, C.J., Cai, S.Y. et al. Mrp4‐/‐ mice have an impaired cyto- :622–31.
protective response in obstructive cholestasis. Hepatology, 2006;43:1013–21. 32. Staudinger, J.L., Woody, S., Sun, M., and Cui, W. Nuclear‐receptor‐mediated
  9. Zollner, G., Wagner, M., Fickert, P. et al. Expression of bile acid synthesis regulation of drug‐ and bile‐acid‐transporter proteins in gut and liver. Drug
and detoxification enzymes and the alternative bile acid efflux pump MRP4 Metab Rev,2013;45:48–59.
in patients with primary biliary cirrhosis. Liver Int, 2007;27:920–9. 33. Gartung, C., Ananthanarayanan, M., Rahman, M.A. et al. Down‐regulation
10. Takeyama, Y., Uehara, Y., Inomata, S. et al. Alternative transporter pathways of expression and function of the rat liver Na+/bile acid cotransporter in
in patients with untreated early‐stage and late‐stage primary biliary cirrhosis. extrahepatic cholestasis. Gastroenterology, 1996;110:199–209.
Liver Int, 2009;29:406–14. 34. Landmann, L. Cholestasis‐induced alterations of the trans‐ and paracellular
11. Boyer, J.L., Trauner, M., Mennone, A. et al. Upregulation of a basolateral pathways in rat hepatocytes. Histochemistry,1995;103:3–9.
FXR‐dependent bile acid efflux transporter OSTalpha‐OSTbeta in cholesta- 35. Green, R.M., Beier, D., and Gollan, J.L. Regulation of hepatocyte bile salt
sis in humans and rodents. Am J Physiol Gastrointest Liver Physiol, transporters by endotoxin and inflammatory cytokines in rodents.
2006;290:G1124–30. Gastroenterology, 1996;111:193–8.
12. Makino, I., Hashimoto, H., Shinozaki, K., Yoshino, K., and Nakagawa, S. 36. Elsing, C., Reichen, J., Marti, U., and Renner, E.L. Hepatocellular Na+/H+
Sulfated and nonsulfated bile acids in urine, serum, and bile of patients with exchange is activated at transcriptional and posttranscriptional levels in rat
hepatobiliary diseases. Gastroenterol, 1975;68:545–53. biliary cirrhosis. Gastroenterology, 1994;107:468–78.
13. Lee, J., Azzaroli, F., Wang, L. et al. Adaptive regulation of bile salt transport- 37. Zollner, G., Fickert, P., Zenz, R. et al. Hepatobiliary transporter expression
ers in kidney and liver in obstructive cholestasis in the rat. Gastroenterology, in percutaneous liver biopsies of patients with cholestatic liver diseases.
2001;121:1473–84. Hepatology, 2001;33:633–46.
14. Pan, Q., Zhang, X., Zhang, L. et al. Solute carrier organic anion transporter 38. Trauner, M., Arrese, M., Lee, H., Boyer, J.L., and Karpen, S. Endotoxin
family member 3A1 is a bile acid efflux transporter in cholestasis. downregulates rat hepatic Ntcp gene expression via decreased activity of
Gastroenterology, 2018; 155(5):1578–92. critical transcription factors. J Clin Invest, 1998;101:2092–100.
15. Kruh, G.D., Belinsky, M.G., Gallo, J.M., and Lee, K. Physiological and phar- 39. Jung, D., Hagenbuch, B., Fried, M., Meier, P.J., and Kullak‐Ublick, G.A.
macological functions of Mrp2, Mrp3 and Mrp4 as determined from recent Role of liver‐enriched transcription factors and nuclear receptors in regulat-
studies on gene‐disrupted mice. Cancer Metastasis Rev, 2007;26:5–14. ing the human, mouse, and rat NTCP gene. Am J Physiol Gastrointest Liver
16. Zamek‐Gliszczynski, M.J., Hoffmaster, K.A., Nezasa, K., Tallman, M.N. Physiol, 2004;286:G752–61.
and Brouwer, K.L. Integration of hepatic drug transporters and phase II 40. Eloranta, J.J., Jung, D., and Kullak‐Ublick, G.A. The human Na+‐taurocho-
metabolizing enzymes: mechanisms of hepatic excretion of sulfate, glucuro- late cotransporting polypeptide gene is activated by glucocorticoid receptor
nide, and glutathione metabolites. Eur J Pharm Sci, 2006;27:447–86. and peroxisome proliferator‐activated receptor‐gamma coactivator‐1alpha,
17. Lee, J.M., Azzaroli, F., Gigliozzi, A., Mennone, A., and Boyer, J.L. Adaptive and suppressed by bile acids via a small heterodimer partner‐dependent
regulation of the ileal sodium‐dependent bile salt transporter (ISBT), and the mechanism. Mol Endocrinol,2006;20:65–79.
multispecific organic anion transporter (MRP2), in kidney and cholangio- 41. Karpen, S.J., Sun, A., Kudish, B. et al. Multiple factors regulate the rat liver
cytes in chronic cholestasis  –  alternative pathways for bile salt excretion. basolateral sodium‐dependent bile acid cotransporter gene promoter. J Biol
Hepatology, 1999;30:417A. Chem, 1996;271:15211–21.
18. Hruz, P., Zimmermann, C., Gutmann, H. et  al. Adaptive regulation of the 42. Denson, L.A., Auld, K.L., Schiek, D.S., McClure, M.H., Mangelsdorf, D.J.,
ileal apical sodium dependent bile acid transporter (ASBT) in patients with and Karpen, S.J. Interleukin‐1β suppresses retinoid transactivation of two
obstructive cholestasis. Gut, 2006;55:395–402. hepatic transporter genes involved in bile formation. J Biol Chem, 2000;275:
19. Dietrich, C.G., Geier, A., Salein, N. et al. Consequences of bile duct obstruc- 8835–43.
tion on intestinal expression and function of multidrug resistance‐associated 43. Denson, L.A., Sturm, E., Echevarria, W. et al. The orphan nuclear receptor,
protein 2. Gastroenterology, 2004;126:1044–53. shp, mediates bile acid‐induced inhibition of the rat bile acid transporter,
20. Chiang, J.Y. Bile acid metabolism and signaling. Compr Physiol, 2013;3: Ntcp. Gastroenterology, 2001;121:140–7.
1191–1212. 44. Kullak‐Ublick, G.‐A., Beuers, U., Fahney, C., Hagenbuch, B., Meier, P.J.,
21. Boyer, J.L. Nuclear receptor ligands: rational and effective therapy for and Paumgartner, G. Identification and functional characterization of the
chronic cholestatic liver disease? Gastroenterology, 2005;129:735–40. promoter region of the human organic anion transporting polypeptide gene.
22. Guo, G.L., Lambert, G., Negishi, M. et al. Complementary roles of farnesoid Hepatology, 1997;26:991–7.
X receptor, pregnane X receptor, and constitutive androstane receptor in pro- 45. Jung, D., Podvinec, M., Meyer, U.A. et al. Human organic anion transport-
tection against bile acid toxicity. J Biol Chem, 2003;278:45062–71. ing polypeptide 8 promoter is transactivated by the farnesoid X receptor/bile
23. Trottier, J., Milkiewicz, P., Kaeding, J., Verreault, M., and Barbier, O. acid receptor. Gastroenterology, 2002;122:1954–66.
Coordinate regulation of hepatic bile acid oxidation and conjugation by 46. Hagenbuch, B. and Stieger, B. The SLCO (former SLC21) superfamily of
nuclear receptors. Mol Pharm, 2006;3:212–22. transporters. Mol Aspects Med, 2013;34:396–412.
24. Stedman, C.A., Liddle, C., Coulter, S.A. et al. Nuclear receptors constitutive 47. Zollner, G., Fickert, P., Silbert, D. et al. Adaptive changes in hepatobiliary
androstane receptor and pregnane X receptor ameliorate cholestatic liver transporter expression in primary biliary cirrhosis. J Hepatol, 2003;38:
injury. Proc Natl Acad Sci USA, 2005;102:2063–8. 717–27.
25. Zimmerman, T.L., Thevananther, S., Ghose, R., Burns, A.R., and Karpen, 48. Lee, W., Glaeser, H., Smith, L.H. et  al. Polymorphisms in human organic
S.J. Nuclear export of retinoid X receptor alpha in response to interleukin‐ anion‐transporting polypeptide 1A2 (OATP1A2): implications for altered
1beta‐mediated cell signaling: roles for JNK and SER260. J Biol Chem, drug disposition and central nervous system drug entry. J Biol Chem,
2006;281:15434–40. 2005;280:9610–7.
388 THE LIVER:  REFERENCES

49. Geier, A., Dietrich, C.G., Voigt, S. et  al. Effects of proinflammatory 74. Maeda, K. and Sugiyama, Y. Impact of genetic polymorphisms of transport-
cytokines on rat organic anion transporters during toxic liver injury and chol- ers on the pharmacokinetic, pharmacodynamic and toxicological properties
estasis. Hepatology, 2003;38:345–54. of anionic drugs. Drug Metab Pharmacokinet, 2008;23:223–35.
50. Denk, G.U., Soroka, C.J., Mennone, A., Koepsell, H., Beuers, U., and Boyer, 75. Schrenk, D., Gant, T.W., Preisegger K‐H., Silverman, J.A., Marino, P.A., and
J.L. Down‐regulation of the organic cation transporter 1 of rat liver in Thorgeirsson, S.S. Induction of multidrug resistance gene expression during
obstructive cholestasis. Hepatology, 2004;39:1382–9. cholestasis in rats and nonhuman primates. Hepatology, 1993;17:854–60.
51. Saborowski, M., Kullak‐Ublick, G.A., and Eloranta, J.J. The human organic 76. Bentires‐Alj, M., Barbu, V., Fillet, M. et al. NF‐kappaB transcription factor
cation transporter 1 gene is transactivated by hepatocyte nuclear factor‐ induces drug resistance through MDR1 expression in cancer cells. Oncogene,
4{alpha}. J Pharmacol Exp Ther, 2006;317(2):778–85. 2003;22:90–7.
52. Hofmann, A.F. and Hagey, L.R. Bile acids: chemistry, pathochemistry, biol- 77. Geick, A., Eichelbaum, M., and Burk, O. Nuclear receptor response elements
ogy, pathobiology, and therapeutics. Cell Mol Life Sci, 2008;65:2461–83. mediate induction of intestinal MDR1 by rifampin. J Biol Chem, 2001;276:
53. Goodwin, B., Jones, S.A., Price, R.R. et  al. A regulatory cascade of the 14581–7.
nuclear receptors FXR, SHP‐1, and LRH‐1 represses bile acid biosynthesis. 78. Deleuze, J.F., Jacquemin, E., Dubuisson, C. et al. Defect on multidrug‐resist-
Mol Cell, 2000;6:517–26. ance 3 gene expression in a subtype of progressive familial intrahepatic chol-
54. Lin, B.C., Wang, M., Blackmore, C., and Desnoyers, L.R. Liver‐specific estasis. Hepatology, 1996;23:904–8.
activities of FGF19 require Klotho beta. J Biol Chem, 2007;282:27277–84. 79. Smit, J.J., Schinkel, A.H., Oude Elferink, R.P. et al. Homozygous disruption
55. Inagaki, T., Choi, M., Moschetta, A. et al. Fibroblast growth factor 15 func- of the murine mdr2 P‐glycoprotein gene leads to a complete absence of
tions as an enterohepatic signal to regulate bile acid homeostasis. Cell phospholipid from bile and to liver disease. Cell, 1993;75:451–62.
Metab, 2005;2:217–25. 80. Fickert, P., Fuchsbichler, A., Wagner, M. et al. Regurgitation of bile acids
56. Schaap, F.G., van der Gaag, N.A., Gouma, D.J., and Jansen, P.L. High expres- from leaky bile ducts causes sclerosing cholangitis in Mdr2 (Abcb4) knock-
sion of the bile salt‐homeostatic hormone fibroblast growth factor 19 in the out mice. Gastroenterology, 2004;127:261–74.
liver of patients with extrahepatic cholestasis. Hepatology, 2009;49:1228–35. 81. Bull, L.N., Carlton VE.H., Stricker, N.L. et al. Genetic and morphological find-
57. Li, T., Jahan, A., and Chiang, J.Y. Bile acids and cytokines inhibit the human ings in progressive familial intrahepatic cholestasis (Byler disease [PFIC‐1] and
cholesterol 7 alpha‐hydroxylase gene via the JNK/c‐jun pathway in human Byler syndrome): evidence for heterogeneity. Hepatology, 1997;26:155–64.
liver cells. Hepatology, 2006;43:1202–10. 82. de Vree, J.M., Jacquemin, E., Sturm, E. et al. Mutations in the MDR3 gene
58. Liu, Y., Binz, J., Numerick, M.J. et al. Hepatoprotection by the farnesoid X cause progressive familial intrahepatic cholestasis. Proc Natl Acad Sci USA,
receptor agonist GW4064 in rat models of intra‐ and extrahepatic cholesta- 1998;95:282–7.
sis. J Clin Invest, 2003;112:1678–87. 83. Jacquemin, E., Cresteil, D., Manouvrier, S., Boute, O., and Hadchouel, M.
59. Chen, J. and Farrell, G.C. Bile acids produce a generalized reduction of the Heterozygous non‐sense mutation of the MDR3 gene in familial intrahepatic
catalytic activity of cytochromes P450 and other hepatic microsomal cholestasis of pregnancy. Lancet, 1999;353:210–11.
enzymes in vitro: Relevance to drug metabolism in experimental cholestasis. 84. Pauli‐Magnus, C., Lang, T., Meier, Y. et  al. Sequence analysis of bile salt
J Gastroenterol Hepatol, 1996;11:870–7. export pump (ABCB11) and multidrug resistance p‐glycoprotein 3 (ABCB4,
60. Marschall, H.U., Wagner, M., Bodin, K. et al. Fxr(‐/‐) mice adapt to biliary MDR3) in patients with intrahepatic cholestasis of pregnancy. Pharmacogenet,
obstruction by enhanced phase I detoxification and renal elimination of bile 2004;14:91–102.
acids. J Lipid Res, 2006;47:582–92. 85. Lang, C., Meier, Y., Stieger, B. et al. Mutations and polymorphisms in the
61. Saini, S.P., Sonoda, J., Xu, L. et al. A novel constitutive androstane recep- bile salt export pump and the multidrug resistance protein 3 associated with
tor‐mediated and CYP3A‐independent pathway of bile acid detoxification. drug‐induced liver injury. Pharmacogenet Genomics, 2007;17:47–60.
Mol Pharmacol, 2004;65:292–300. 86. Gotthardt, D., Runz, H., Keitel, V. et al. A mutation in the canalicular phos-
62. Miyata, M., Matsuda, Y., Tsuchiya, H. et al. Chenodeoxycholic acid‐mediated pholipid transporter gene, ABCB4, is associated with cholestasis, ductope-
activation of the farnesoid X receptor negatively regulates hydroxysteroid nia, and cirrhosis in adults. Hepatology, 2008;48(4):1157–66.
sulfotransferase. Drug Metab Pharmacokinet, 2006;21:315–23. 87. Huang, L., Zhao, A., Lew, J.L. et al. Farnesoid X receptor activates transcrip-
63. Pircher, P.C., Kitto, J.L., Petrowski, M.L. et al. Farnesoid X receptor regu- tion of the phospholipid pump MDR3. J Biol Chem, 2003;278:51085–90.
lates bile acid‐amino acid conjugation. J Biol Chem, 2003;278:27703–11. 88. Ghonem, N.S., Ananthanarayanan, M., Soroka, C.J., and Boyer, J.L.
64. Carlton, V.E., Harris, B.Z., Puffenberger, E.G. et al. Complex inheritance of Peroxisome proliferator‐activated receptor alpha activates human multidrug
familial hypercholanemia with associated mutations in TJP2 and BAAT. Nat resistance transporter 3/ATP‐binding cassette protein subfamily B4 tran-
Genet, 2003;34:91–6. scription and increases rat biliary phosphatidylcholine secretion. Hepatology,
65. Konig, J., Nies, A.T., Cui, Y., Leier, I., and Keppler, D. Conjugate export 2014;59:1030–42.
pumps of the multidrug resistance protein (MRP) family: localization, sub- 89. Meier, Y., Pauli‐Magnus, C., Zanger, U.M. et al. Interindividual variability of
strate specificity, and MRP2‐mediated drug resistance. Biochim Biophys canalicular ATP‐binding‐cassette (ABC)‐transporter expression in human
Acta, 1999;1461:377–94. liver. Hepatology, 2006;44:62–74.
66. Borst, P., Evers, R., Kool, M., and Wijnholds, J. The multidrug resistance 90. Buechler, M., Koenig, J., Brom, M. et al. cDNA cloning of the hepatocyte
protein family. Biochim Biophys Acta, 1999;1461:347–57. canalicular isoform of the multidrug resistance protein, cMrp, reveals a novel
67. Bohan, A., Chen, W.S., Denson, L.A., Held, M.A., and Boyer, J.L. Tumor conjugate export pump deficient in hyperbilirubinemic mutant rats. J Biol
necrosis factor alpha‐dependent up‐regulation of Lrh‐1 and Mrp3(Abcc3) Chem, 1996;271:15091–8.
reduces liver injury in obstructive cholestasis. J Biol Chem, 91. Ito, K., Suzuki, H., Hirohashi, T., Kazuhiko, K., Shimizu, T., and Sugiyama,
2003;278:36688–98. Y. Molecular cloning of canalicular multispecific organic anion transporter
68. Chen, W., Cai, S.Y., Xu, S., Denson, L.A., Soroka, C.J., and Boyer, J.L. defective in EHBR. Am J Physiol, 1997;272:G16–22.
Nuclear receptors RXRalpha:RARalpha are repressors for human MRP3 92. Kartenbeck, J., Leuschner, U., Mayer, R., and Keppler, D. Absence of the
expression. Am J Physiol Gastrointest Liver Physiol, 2007;292:G1221–27. canalicular isoform of the MRP gene‐encoded conjugate export pump from
69. Belinsky, M.G., Dawson, P.A., Shchaveleva, I. et al. Analysis of the in vivo the hepatocytes in Dubin–Johnson syndrome. Hepatology, 1996;23:1061–6.
functions of Mrp3. Mol Pharmacol, 2005;68:160–8. 93. Oude Elferink, R.P.J., Ottenhoff, R., Liefting, W., Haan, J.D., and Jansen,
70. Huang, W., Zhang, Z., Chua, S.S. et al. Induction of bilirubin clearance by P.L.M. Hepatobiliary transport of glutathione and glutathione conjugate in
the constitutive androstane receptor (CAR). PNAS, 2003;100:4156–61. rats with hereditary hyperbilirubinemia. J Clin Invest, 1989;84:478–83.
71. Okuwaki, M., Takada, T., Iwayanagi, Y. et  al. LXR alpha transactivates 94. Paulusma, C., van Geer, M., Evers, R. et al. Canalicular multispecific organic
mouse organic solute transporter alpha and beta via IR‐1 elements shared anion transporter/multidrug resistance protein 2 mediates low‐affinity trans-
with FXR. Pharm Res, 2007;24:390–8. port of reduced glutathione. Biochem J, 1999;338:393–401.
72. Soroka, C.J., Mennone, A., Hagey, L.R., Ballatori, N., and Boyer, J.L. Mouse 95. Choi, J.H., Ahn, B.M., Yi, J. et al. MRP2 haplotypes confer differential sus-
organic solute transporter alpha deficiency enhances renal excretion of bile ceptibility to toxic liver injury. Pharmacogenet Genomics,2007;17:403–15.
acids and attenuates cholestasis. Hepatology, 2010;51:181–90. 96. Lee, J.M., Trauner, M., Soroka, C.J., Stieger, B., Meier, P.J., and Boyer, J.L.
73. Lam, P., Wang, R., and Ling, V. Bile acid transport in sister of P‐glycoprotein Expression of the bile salt export pump is maintained after chronic cholesta-
(ABCB11) knockout mice. Biochemistry, 2005;44:12598–605. sis in the rat. Gastroenterology, 2000;118:163–72.
31:  Adaptive Regulation of Hepatocyte Transporters in Cholestasis 389

  97. Strautnieks, S.S., Byrne, J.A., Pawlikowska, L. et al. Severe bile salt export 107. Suzuki, M., Suzuki, H., Sugimoto, Y., and Sugiyama, Y. ABCG2 transports sul-
pump deficiency: 82 different ABCB11 mutations in 109 families. fated conjugates of steroids and xenobiotics. J Biol Chem, 2003;278:22644–9.
Gastroenterology, 2008;134:1203–14. 108. Mennone, A., Soroka, C.J., Harry, K.M., and Boyer, J.L. Role of breast
  98. Stieger, B., Fattinger, K., Madon, J., Kullak‐Ublick G‐A., and Meier, P.J. Drug‐ cancer resistance protein in the adaptive response to cholestasis. Drug
and estrogen‐induced cholestasis through inhibition of the hepatocellular bile Metab Dispos, 2010;38:1673–8.
salt export pump (Bsep) of rat liver. Gastroenterology, 2000;118:422–30. 109. Graf, G.A., Yu, L., Li, W.P. et al. ABCG5 and ABCG8 are obligate heter-
  99. Donner, M.G., Schumacher, S., Warskulat, U., Heinemann, J., and Haussinger, odimers for protein trafficking and biliary cholesterol excretion. J Biol
D. Obstructive cholestasis induces TNF‐{alpha}‐ and IL‐1 ‐mediated peripor- Chem, 2003;278:48275–82.
tal downregulation of Bsep and zonal regulation of Ntcp, Oatp1a4, and 110. Kamisako, T. and Ogawa, H. Alteration of the expression of adenosine
Oatp1b2. Am J Physiol Gastrointest Liver Physiol, 2007;293:G1134–46. triphosphate‐binding cassette transporters associated with bile acid and
100. Lee, J.M., Trauner, M., Soroka, C., Steiger, B., Meier, P.J., and Boyer, J.L. cholesterol transport in the rat liver and intestine during cholestasis. J
The molecular expression of the bile salt excretory pump, sister of P‐glyco- Gastroenterol Hepatol, 2005;20:1429–34.
protein (SPGP), is selectively preserved in cholestatic liver injury. 111. Omote, H., Hiasa, M., Matsumoto, T., Otsuka, M., and Moriyama, Y. The
Hepatology, 1998;28(4):429A. MATE proteins as fundamental transporters of metabolic and xenobiotic
101. Kogan, D., Ananthanarayanan, M., Emre, S., Suchy, F.J., and Shneider, organic cations. Trends Pharmacol Sci, 2006;27:587–93.
B.L. The bile salt excretory pump (BSEP/SPGP) is not down‐regulated in 112. Paulusma, C.C., Groen, A., Kunne, C. et  al. Atp8b1 deficiency in mice
human cholestasis associated with extrahepatic biliary atresia (EHBA). reduces resistance of the canalicular membrane to hydrophobic bile salts
Hepatology, 1999;30:468A. and impairs bile salt transport. Hepatology, 2006;44:195–204.
102. Ananthanarayanan, M., Balasubramanian, N., Makishima, M., Mangelsdorf, 113. Bull, L.N., van Eijk, M.J., Pawlikowska, L. et al. A gene encoding a P‐type
D.J., and Suchy, F.J. Human bile salt export pump promoter is transactivated by ATPase mutated in two forms of hereditary cholestasis. Nat Genet,
the farnesoid X receptor/bile acid receptor. J Biol Chem, 2001;276:28857–65. 1998;18:219–24.
103. Sinal, C.J., Tohkin, M., Miyata, M., Ward, J.M., Lambert, G., and Gonzalez, 114. Boyer, J.L. Bile duct epithelium: frontiers in transport physiology. Am
F.J. Targeted disruption of the nuclear receptor FXR/BAR impairs bile acid J Physiol, 1996;270:G1–5.
and lipid homeostasis. Cell, 2000;102:731–44. 115. Hohenester, S., Wenniger, L.M., Paulusma, C.C. et  al. A biliary HCO3‐
104. Song, X., Kaimal, R., Yan, B., and Deng, R. Liver receptor homolog 1 tran- umbrella constitutes a protective mechanism against bile acid‐induced
scriptionally regulates human bile salt export pump expression. J Lipid Res, injury in human cholangiocytes. Hepatology, 2012;55:173–83.
2008;49:973–84. 116. Medina, J.F., Martinez‐Anso, E., Vazquez, J.J., and Prieto, J. Decreased
105. Weerachayaphorn, J., Cai, S.Y., Soroka, C.J., and Boyer, J.L. Nuclear factor anion exchanger 2 immunoreactivity in the liver of patients with primary
erythroid 2‐related factor 2 is a positive regulator of human bile salt export biliary cirrhosis. Hepatology, 1997;25:12–7.
pump expression. Hepatology, 2009;50:1588–96. 117. Concepcion, A.R., Lopez, M., Ardura‐Fabregat, A., and Medina, J.F. Role of
106. Kubitz, R., Droge, C., Stindt, J., Weissenberger, K., and Haussinger, D. The AE2 for pHi regulation in biliary epithelial cells. Front Physiol, 2013;4:413.
bile salt export pump (BSEP) in health and disease. Clin Res Hepatol
Gastroenterol,2012;36:536–53.
SECTION D:
NON‐HEPATOCYTE CELLS
Cholangiocyte Biology
32 and Pathobiology
Massimiliano Cadamuro1,2, Romina Fiorotto2,3, and Mario Strazzabosco2,3
1
Department of Molecular Medicine, University of Padua, Padova, Italy
2
International Center for Digestive Health (ICDH), University of Milan‐Bicocca, Monza, Italy
3
Liver Center and Section of Digestive Diseases, Department of Internal Medicine, Section of Digestive
Diseases, Yale University School of Medicine, New Haven, CT, USA

INTRODUCTION CoH connect the hepatocellular canalicular network, carrying


the primary bile, with the intraportal ramifications of the biliary
Cholangiocytes are the epithelial cells that line the intrahepatic tree. CoH is also considered the site of the putative liver stem
and extrahepatic biliary tree. The last two decades have wit- cell niche. The intrahepatic biliary tree gradually merges from
nessed a significant expansion of our understanding of the func- cholangioles, interlobular, septal, areal, and segmental ducts
tion and dysfunction of this important epithelium. We have also that end in the two principal hepatic ducts and then the common
understood that cholangiocytes are important actors in liver hepatic duct. The latter receives the cystic duct coming from the
repair and progression of liver fibrosis and play a fundamental gallbladder and connects the liver and gallbladder to the intes-
role in liver immunobiology. Several chapters and reviews have tine. The extrahepatic biliary network is surrounded by a capil-
been recently written on a number of aspects of cholangiocyte lary plexus and by peribiliary glands, a stem cell niche of hepatic
biology. In this review we elected to focus on those mechanisms progenitor cells that differs from that of the CoH. Notably, the
that are more pathophysiologically relevant to human diseases intra‐ and extrahepatic biliary tree has a different embryonic ori-
and that may represent possible avenues for basic and transla- gin [5, 6].
tion research in the next few years. Cholangiocytes belonging to different compartments of the
biliary tree are characterized by a specialized morphology,
reflecting distinct physiological function. The smallest biliary
epithelial cells lining the CoH and the distal branches of the
FUNCTIONAL ANATOMY biliary tree have a cuboidal shape with a round basal nucleus,
OF THE BILIARY TREE and are quickly reactive to liver and/or biliary damage. The
large cholangiocytes, lining the bile ducts of higher diameter,
The biliary system is a complex network of tubules coated by are usually columnar and, thanks to their transport abilities,
epithelial cells, or cholangiocytes, which starts from the canals mediate alkalinization, hydration, and modification of the bile
of Hering in the liver lobules (intrahepatic biliary tree), contin- [1, 7]. Cholangiocytes in fact, have both secretory and absorp-
ues outside the liver (extrahepatic biliary tree), and terminates tive functions, being involved in the recirculation of bile acids,
into the ampulla of Vater. The best known function of the biliary glucose, and water. Notably, about 40% of human bile is gener-
tree is the transport and modification of the primary bile secreted ated by biliary epithelial cells, following modification of the
by the hepatocytes, but it is now widely recognized that the bil- primary product of the hepatocytes. Data in animals suggest that
iary tree is actually a functionally complex structure endowed while secretory functions of large cholangiocytes are regulated
with many different biological functions reflected in a morpho‐ mostly by cyclic adenosine monophosphate (cAMP) signaling,
functional specialization of cholangiocytes [1, 2]. The first the most important second messenger in small cholangiocytes is
structure belonging to the biliary tree is represented by the [Ca2+]i [7, 8]. These distinctions are, however, less actual now
canals of Hering (CoH), that is composed 50% by cholangio- that we have a better understanding of the microdomain restric-
cytes and 50% by hepatocytes juxtaposed with each other [3, 4]; tion of second messenger signaling.

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
394 THE LIVER:  DEVELOPMENT OF THE BILIARY TREE

DEVELOPMENT OF THE BILIARY TREE [19–21], periportal hepatoblasts expressing Notch2 are induced
by Jag1+ portal mesenchymal cells to acquire cholangiocyte‐like
The extra‐ and intrahepatic branches of the biliary tree originate morphology and phenotype (DP cells). The interaction between
from different embryological areas, thus explaining their mor- Jag1 and its receptor Notch2 induces the cleavage and nuclear
phological and biological properties. The extrahepatic biliary import of the Notch intracellular domain (NICD) that, in coop-
tree originates from Hex−/Pdx1+/Sox17+ cells in the caudal part eration with TGFβ, binds to the nuclear transcription factor
of the ventral foregut [9, 10]. Studies conducted on Alagille syn- recombinant signal binding protein for immunoglobulin kappa J
drome demonstrate that perturbation of the Notch pathway, one (RBP‐Jk) and activates Hes1 that, in turn promotes the expres-
of the morphogenetic signaling pathways involved in the biliary sion of the biliary epithelial cell‐specific transcription factors
specification of the hepatoblasts, causes intrahepatic ductopenia Hnf1β, Sox4, and Sox9. Notably, mutations on Jag1 or Notch2,
without affecting the extrahepatic biliary structures, and sug- cause Alagille syndrome, a multiorgan disease characterized by
gests that different signals govern the formation of the intra‐ and vanishing of intrahepatic biliary structures, and by the accumu-
extrahepatic structures [11]. lation of cells retaining both hepatic and cholangiocyte phe-
Conversely, intrahepatic bile ducts derive from Hex+/Pdx‐1−/ notype (intermediate hepatobiliary cells) [11, 22]; jaundice,
Sox17−/AFP+ hepatoblasts, bipotent cells present in the fetal itching, and hypercholesterolemia characterize the hepatic
liver bud. Starting from the eighth gestational week (GW), phenotype.
hepatoblasts in contact with mesenchymal cells located in the Recent work by Diehl’s group has clarified the important role
nascent portal tract, form a continue single cells rim of keratin of the Hedgehog (Hh) pathway in biliary morphogenesis, repair,
K8+/18+/19+ cells called the ductal plate (DP). Between twelfth and cancer. This signaling is normally repressed by the interac-
and sixteenth GW, the DP duplicates in discrete areas, represses tion between Patched (Ptc) and Smoothened (Smo) [23–25].
the expression of hepatocellular markers and begins to express Once the Hh ligand (i.e. Sonic, SHh) binds its receptor Patched
K7 and K19, that is, markers of commitment toward a biliary (Ptc), the signal pathway is de‐repressed and Gli3A induces the
phenotype. The duplicating DP progressively forms a lumen, transcription of the downstream effectors Ptc, Glioblastoma
while the residual biliary monolayers are gradually reabsorbed (Gli) 1 and Gli2. Recent data support an important role of Hh in
by apoptosis or by retro‐differentiation to periportal hepato- bile duct morphogenesis; in fact, in the fetal mouse liver, SHh
cytes, as recently shown by lineage tracking experiments [12]. expression together with Gli1, increase at the earliest gesta-
Duplicating DPs are gradually incorporated into the portal mes- tional days (E11.5) and rapidly decrease overtime. Its reported
enchyma where they are joined by a peribiliary capillary plexus, association with the Meckel syndrome, a lethal DP malforma-
which nourishes the nascent biliary tree, and by vascular struc- tion, highlights the morphogenetic properties of Hh signaling.
tures that will become the portal arteries [13]. Hh also has modulatory effects on two other fundamental tran-
The concomitant elongation and maturation of biliary epithe- scription factors, Yes‐associated protein (YAP) and Sox9. In
lium and vascular structures requires the cooperation of differ- mice, YAP is expressed at E15.5 by several cells accumulated in
ent cell types equipped with several specific ligands, receptors, the periportal areas around the nascent portal tract; at E18.5 the
and other morphogens, among them angiogenic growth factors, clear majority of YAP+ cells co‐express Sox9, a marker of
such as vascular endothelial growth factor (VEGF) A, platelet‐ ­biliary differentiation, which is also downstream of Notch
derived growth factor (PDGF), and the angiopoietins, Ang‐1 signaling [26]. The role of YAP as a master gene involved in
and Ang‐2, and their specific receptors, VEGFR1, VEGFR2, differentiation of cholangiocyte precursors is also confirmed by
PDGFRβ, and Tie‐2 [13, 14]. In addition, a range of morphoge- the essential role played by the Hippo pathway in regulating
netic signaling and transcription factors, including Notch and liver size during liver regeneration after partial hepatectomy
WNT/β‐catenin pathways, transforming growth factor‐β (PHx) [27]. Furthermore, recent data [28] show that Sox9 and
(TGF‐β), and hepatocyte nuclear factors (HNFs) are activated Sox4, are involved in the maintenance of the apical–basal
[9, 15]. Furthermore, recent studies outlined the importance of ­polarity of the epithelial sheet, in the correct development of the
miRNAs as regulatory mediators coordinating the interactions primary cilium and, as well as in the normal formation, elongation
among all those peptides. It is important for the experimental as and branching of the biliary tree. This mechanism, which is
well as the clinical hepatologist to acquire a working knowledge known as “planar cell polarity” (PCP), is primarily governed by
of these mechanisms, since an altered development of the bil- the non‐canonical Wnt/β‐catenin pathway, and is altered in
iary tract is responsible for a number of liver malformative dis- some kidney and liver ciliopathies (diseases deriving from a
eases, among which the so‐called DP malformations (DPM) are faulty assembly of primary cilia or from the malfunction of
of particular interest [16–18]. ­proteins expressed in cilia) [29].
A fundamental family of growth factors involved in biliary In summary, the available information highlight the intensive
fate determination is TGFβ, and its receptors; in particular, cross‐talk mechanisms among the different morphogens and
TGFβ type II‐R (TβRII). All TGFβ ligands are highly expressed likely a redundancy of the signals regulating biliary tree archi-
around the nascent portal area, and induce the transition of tecture, in health and in the reparative/regenerative response to
hepatoblasts into cholangiocytes, by binding to TβRII, which is liver damage. More recently, miRNAs have been suggested to
transiently expressed by the monostratified DP cells. This effect play a role in biliary tree specification and development; Rogler
is, at least in part, mediated by the upregulation of Jag1 and by and colleagues showed that the treatment of mice fetuses at
the modulation of the transcription factors Hes1 and Hey1, two E16.5 with miR‐23b, miR‐27b, and miR‐24 antagonists per-
downstream effectors of Notch signaling. As demonstrated by turbed the correct development of the fetal liver and stimulated
studies on different animal models (zebrafish and mice) the expression of keratin 19 in the liver parenchyma due to the
32:  Cholangiocyte Biology and Pathobiology 395

upregulation of the TGFβ signaling [30]. Those miRNAs in fact Na+‐dependent Cl−/HCO3− exchanger (NCHE), and by the Na+/
target different components of the TGFβ pathway, and in par- HCO3− cotransporter (NCB1) present at the basolateral side of
ticular at the level of the small mother against decapentaplegic the cholangiocytes. The difference of potential between the
(SMAD) [31]. intra‐ and extracellular sides of the epithelium is maintained by
two Na+ transport systems: the Na+/K+/2Cl− cotransporter
(NKCC1) and the Na+/K+ adenosine triphosphate (ATP)‐ase [8].
These coordinated mechanisms are necessary to increase the
PHYSIOLOGICAL FUNCTIONS alkalinity and the volume of the bile and their dysfunction can
OF THE BILIARY TREE be responsible for pathologic conditions. In cystic fibrosis (CF),
for example, a genetic disease caused by a genetically transmit-
Within the biliary tree there is a functional and morphological ted defect in the function of CFTR, biliary complications are
specialization between small and large ducts; in particular, the increasingly being recognized. In cystic fibrosis liver disease
main secretory functions are sustained by cholangiocytes lining (CFLD) defective CFTR function alters the coordinated activa-
interlobular, septal, and major ducts, while the ability to react to tion of AE2 and the secretion of bicarbonate and fluid into the
liver damage and possibly function as bipotential progenitor bile. Alterations of the bile composition precipitates a cascade
cells resides in the small ducts and canals of Hering, respec- of events such as the formation of bile plugs, accumulation of
tively [32–36]. toxicants (i.e. toxic bile acids, endotoxins) that damage the epi-
thelium and causes focal biliary inflammation progressing to
sclerosing cholangitis and eventually cirrhosis [43]. Recent
findings (revised in sections on Barrier Function of
SECRETION AND BILE PRODUCTION Cholangiocytes and in Cholangiocytes and Immunity) have
highlighted a novel function of CFTR as a regulator of endo-
The best understood function of cholangiocytes is to modify the toxin tolerance in cholangiocytes, thanks to its interaction with
volume, fluidity, and alkalinity of the primary bile secreted by other proteins (i.e. Src tyrosine kinases), and this will change
the hepatocytes. In addition, the biliary epithelium may reab- our view of CFLD from a classic channelopathy to an inflam-
sorb water, glucose, glutathione, bile acids, and electrolytes matory disease (Figure 32.1).
[37]. Actually, biliary epithelial cells are able to modify the Bicarbonate secretion is not only necessary to drive bile
composition of the primary bile secreted by the hepatocytes into secretory processes but it is also an important defense mecha-
the bile canaliculi and in humans, up to 40% of the bile is pro- nism developed by cholangiocytes to protect themselves from
duced by the biliary epithelium, depending on the digestive several toxicants present in the bile and by the action of apolar
phases. Primary bile flows through the CoH to the biliary net- protonated hydrophobic bile acids. The vast majority of human
work where cholangiocytes lining the larger biliary structures bile salts is glycine‐conjugated and are usually partially proto-
modify it. Changes in bile volume and composition are modu- nated, and apolar, becoming membrane‐permeable at pH 7.4
lated by a complex interplay among different pro‐secretory hor- [44]. Thus, hydrophobic bile salts accumulate into the cell cyto-
monal and paracrine stimuli, among which secretin and ATP are plasm and may cause cell damage, including apoptosis at micro-
important examples [38, 39], and anti‐secretory factors such as molar concentrations. Starting from these observations, it was
endothelin‐1 (ET‐1) [40], gastrin, and somatostatin [41]. These hypothesized that the secretion of HCO3− above the apical mem-
secretory and anti‐secretory stimuli mostly work by increasing brane of cholangiocytes provides a sort of local bicarbonate
or decreasing the levels of cellular cAMP as will be described in shield (or umbrella) able to protect the epithelium. In physio-
detail below. logic conditions, several mechanisms are at play to generate and
The above described secretory mediators are integrated by a maintain this bicarbonate rich microenvironment [45]. In nor-
number of membrane‐bound or soluble adenylyl cyclase (ACs), mal cholangiocytes, the primary bile acid chenodeoxycholic
such as AC4, 5, 6, 7, 8, 9, and SAC (soluble adenyl cyclase), acid (CDCA) interacts with its receptor TGR5, a membrane‐
resulting in given levels of intracellular levels of cAMP, and bound bile acid receptor coupled to a stimulatory G‐protein,
activation of PKA or epithelial Na+ channel (ENAC). Recent localized both on the apical membrane and on the primary cil-
works have highlighted the importance of microdomains in ium. Activation of TGR5 is thought to increase cAMP concen-
cAMP/PKA activation. Some of these microdomains may be tration and activate the excretion of Cl−, together with ATP
associated with a specific ACs that binds PKA to selected pro- through CFTR. ATP released into the bile binds specific surface
teins. Activation of PKA stimulates the cystic fibrosis trans- purinergic receptor P2Y2 that increases intracellular Ca2+ levels
membrane conductance regulator (CFTR) channel [42] that and stimulates by an autocrine loop the secretion of Cl− from
mediates the luminal secretion of Cl−. Cl− secreted into the calcium‐dependent Cl‐channels (i.e. TMEM16A). Cl− is trans-
lumen creates the driving force for the activation of an apically ported back into the cytoplasm by the AE2 exchanger as
located Na+‐independent Cl−/HCO3− exchanger (AE2) that described in the section on Secretion and Bile Production.
extrudes HCO3− in exchange for Cl− from the apical surface of Finally, the action of alkaline phosphatases, located in the apical
the biliary epithelial cells. The concentration of anions into the glycocalyx in close proximity to the P2Y receptors ensures the
lumen of the bile ducts creates an osmotic gradient that attracts correct amount of bicarbonate secretion by degradation of ATP
water into the bile ducts, through aquaporin‐1 and ‐4. To pre- in ADP and AMP [46].
serve cellular homeostasis, the secretion of bicarbonate is coun- It is hypothesized that a “destabilization” of the biliary bicar-
terbalanced by the intracellular import of anions through the bonate shield could generate changes in the luminal pH that
396 THE LIVER:  BILIARY PRIMARY CILIA AND “CILIOPATHIES”

Figure 32.1  Bile secretion and modification in cholangiocytes. Cholangiocytes, through a complex network of pumps, ion exchangers, and
channels are able to modify bile composition, altering its volume, pH, and hydration. Prosecretory hormones (i.e. secretin) bind their respective
receptors stimulating the intracellular increase of cAMP due to the activity of ACs, leading to the activation of CFTR that extrude Cl− anions in the
luminal space together with ATP. Cl− is, in turn, reabsorbed by AE2 extruding HCO3−, responsible for the alkalinization of the bile. cAMP levels are
further sustained by the activation of TGR5 on the apical side of cholangiocytes. The increased osmotic gradient at the apical side of the cholangio-
cytes stimulates H2O flow passively through AQPs (AQP1, AQP4), as well as through a paracellular pathway; ATP, also secreted into the bile by
CFTR, activates the P2Y purinergic receptor and Cl− secretion. Cellular ion homeostasis is maintained by the presence of several ion transporters
at the basal side of cholangiocytes, such as NCHE, NCB1, and Na+/H− pump. This mechanism leads to HCO3− secretion into the bile, which together
with the presence of a glycocalyx, constitutes a well‐known defensive mechanism that is also at the basis of the cholehepatic shunting of weak acids
and bile acids (see [127]), dubbed the “bicarbonate umbrella”. Abbreviations: cAMP, cyclic adenosine monophosphate; ACs, adenylyl cyclase;
CFTR, cystic fibrosis transmembrane conductance regulator; AE2, Na+‐independent Cl−/HCO3− exchanger; AQPs, aquaporins; ENAC, epithelial
Na+ channel; NCB1, Na+/HCO3− cotransporter; NCHE, Na+‐dependent Cl−/HCO3− exchanger; NHE‐1, sodium‐hydrogen antiporter 1.

alter biliary homeostasis and bile salt physicochemical status of HCO3−, through a Ca2+/cPKCα/PKA mechanism and the
and causes toxic effects to the biliary epithelium, thereby trig- P2Y‐dependent signaling and thus reduce the necroinflamma-
gering a chronic cholangitis and scarring and sclerosing cholan- tory damage [48].
gitis akin to the pathophysiological sequence described in the
Mdr3‐knockout (KO) mouse [44, 45, 47].
In chronic cholangitis, AE2 expression seems to be reduced
due to the ex novo expression of miR‐506 that suppresses the BILIARY PRIMARY CILIA
translation of AE2 by binding its 3′UTR region. The reduced AND “CILIOPATHIES”
AE2 activity has a dual effect, that on one side destabilizes the
biliary umbrella and the buffering effect on bile salts, and sec- The apical surface of cholangiocytes presents a non‐motile
ondarily leads to an intracellular accumulation of HCO3− that primary cilium. Recent studies have highlighted its importance
activates the soluble AC (sAC) and sensitizes cholangiocytes to in cholangiocyte biology and pathobiology. Primary, or sensory,
apoptosis. Treatment with ursodeoxycholic acid (UDCA) or the cilia are non‐motile structures, characterized by a 9 + 0 configu-
homolog norUDCA could in part activate the luminal secretion ration, in which, nine doublet microtubules, or axoneme, are
32:  Cholangiocyte Biology and Pathobiology 397

anchored to the basal body [49]. Primary cilia express several play a role the pathogenesis of several liver diseases, including
proteins, including polycystin 1 and 2 (PC1, and PC2), and primary biliary cholangitis (PBC), primary sclerosing cholangi-
fibrocystin (FPC) that could participate in a number of intracel- tis (PSC), and CFLD [56–58].
lular signaling cascades. Primary cilia can act as osmo‐ or The tight junctions (TJs) are among the main structures
mechano‐receptors in biliary structures. Cilia sense the direc- responsible for epithelial barrier function and are of primary
tion of the bile flow and, by bending, activate calcium channels importance for the correct conformation of the biliary struc-
allowing the intracellular influx of Ca2+ ions. These structures tures; unfortunately, to date, very little is known about mecha-
are likely involved in cell proliferation and senescence, activa- nisms controlling their function at the level of canaliculi and the
tion of progenitor cell compartment, regeneration, and develop- bile ducts. In general, assembling and disassembling of TJ are
ment through the modulation of the Hh and canonical WNT modulated by different signaling mechanisms, including protein
pathways. Furthermore, signaling mechanisms located in cilia kinases, and G‐proteins; probiotics, glutamine, and growth fac-
may supervise the correct cell planarity and polarity through tor protect the correct assembly of TJs, while disruption of TJ
non‐canonical WNT signaling [50, 51]. During biliary develop- are induced by pathogens, toxins, and inflammatory cytokines.
ment, the presence and the correct conformation of the primary Experiments on polarized monolayers of cholangiocytes dem-
cilia is necessary for the correct assembly of the biliary struc- onstrated that treatment with TNFα, as well as with lipopolysac-
tures; in fact, its absence is observed in three rare but lethal syn- charides (LPS) and nitric oxide (NO) increase the paracellular
dromes of infants, the Meckel [52], Joubert [53], and Jeune [54] permeability.
syndromes. These are multiorgan diseases and the liver is char- In particular, in human and mouse in vitro cultures of cholan-
acterized by cyst‐like, dysmorphic biliary structures surrounded giocytes carrying mutations for CFTR treated with LPS
by a dense fibrotic stroma, resembling congenital hepatic fibro- increases permeability to dextrans by inducing a derangement
sis (CHF). Several liver diseases, known as ciliopathies, may be of the normal F‐actin cytoskeletal shape, and the delocalization
related to primary cilia defects, including CHF, Caroli’s disease of E‐cadherin, a cytoskeletal structural protein typical of differ-
(CD), autosomal dominant polycystic kidney disease (ADPKD). entiated epithelial cells [58]. In normal cholangiocytes, CFTR is
ADPKD is a genetic liver and kidney disease due to mutation on localized on the apical membrane where it functions as a chan-
the gene encoding for PC1 (80%) and PC2 (20%) characterized nel protein but also interacts in a multiprotein complex and
by progressive and massive enlargement of cyst covered by bil- regulates the function of other proteins. Examples are proteins
iary epithelia; patients with ADPKD may develop severe com- that negatively regulate the activity of Src family tyrosine
plications such as mass effect, cyst hemorrhage, rupture, or kinases (SFK). In cholangiocytes that lack CFTR at the mem-
infection requiring urgent liver transplantation. CHF, CD, and brane, Src is indeed more active and phosphorylates the LPS
autosomal recessive polycystic kidney disease (ARPKD), are receptor TLR4. An aberrant activation of TLR4 in response to
rare inherited diseases of the renal tubular and biliary epithe- LPS and the increased downstream activation of NF‐κB with
lium, characterized by enlargement of the biliary tree with cyst‐ production of inflammatory mediators are directly responsible
like features associated with portal fibrosis and inflammation. for the incorrect reshaping of the F‐actin cytoskeleton and con-
These diseases, all due to mutations in the Pkhd1 gene, encod- sequent increased permeability [58, 59]. Notably, the treatment
ing for FPC cause severe portal hypertension and may be com- with PP2, an SFK inhibitor, prevents the improper assembly of
plicated by acute cholangitis, intrahepatic lithiasis, and also the cytoskeleton and rescues the paracellular permeability of
cholangiocarcinoma. The influences of ciliary protein dysfunc- CF‐KO cells. In contrast to Src, activation of tyrosine kinases
tion in the pathogenesis of cystic liver diseases will be explained (TKs), such as EGFR, have a protective effect on TJ conforma-
in the following chapters. Interestingly, mice carrying mutations tion following hepatic noxae. The activation of the pathway
for the Itf88 gene, also known as Tg737, encoding for the ciliary mediated by EGFR inhibits Src phosphorylation by increasing
protein polaris, display several malformations of the kidney and concentration and activation of the phospholipase Cγ (PLCγ)/
biliary cystic dilatation, accompanied by pericystic fibrosis PKC axis. Altered TJ permeability and back‐diffusion of bile
resembling the ARPKD phenotype [55]. acids may play a pathogenic role in experimental conditions, as
well as in human biliary diseases.

BARRIER FUNCTION
OF CHOLANGIOCYTES CHOLANGIOCYTE REACTION
TO BILIARY DAMAGE
A fundamental function of epithelial cells, including cholangio-
cytes, is to selectively control the diffusion of ions and mole- Biliary epithelial cells are usually quiescent, however, following
cules through the epithelial barrier. Biliary epithelial cells, a liver insult, cholangiocytes activate and/or proliferate as a part
thanks to their ability to secrete ions in a polarized fashion and of the so‐called “hepatic reparative complex”. A typical element
to their selective permeability to solute and water, actively of the hepatic repair response to liver damage is the “ductular
maintain liver homeostasis. Moreover, the biliary epithelium reaction” (DR), a stereotyped histopathological lesion of the bil-
functions as a barrier against the back‐diffusion of xenobiotics, iary epithelium, which plays a fundamental role in the progres-
toxic metabolites, and bile salts from the bile to the interstitial sion of hepatic fibrosis. DR is characterized by a marked
tissue. Primary or secondary changes in the correct formation proliferation of cholangiocytes with poor cytoplasm, arranged
and polarization of the epithelial sheet have been reported to in cell cords without a lumen or in richly anastomosed small
398 THE LIVER:  CHOLANGIOCYTE PROLIFERATION

diameter ducts (<10 μm) with almost unrecognizable lumens consistent with their stem cell nature, such as Pdx1, Sox9,
[51]. An infiltrate composed by innate and adaptive immune Sox17, and EpCAM, together with markers of mature epithelial
cells is invariably present. Reactive ductular cells (RDC) are cells, such as keratin 7 [66, 67]. Notably, these putative stem
activated epithelial cells that secrete a vast array of factors, cells could be isolated through magnetic immunoselection and
including cytokines, chemokines, growth factors, and angio- when cultured by using specific matrix and growth factors, they
genic factors [60]. Different theories, not mutually exclusive, could commit to hepatocyte, cholangiocyte, and pancreatic
have been proposed to explain the genesis of RDCs: it was lineage.
hypothesized that they may derive from hepatocytes undergoing
a process of ductular metaplasia, or from the activation of the
hepatic progenitor cells (HPC) compartment, and/or from pro-
liferation and de‐differentiation of preexisting cholangiocytes CHOLANGIOCYTE PROLIFERATION
[51]. HPC is a population of bipotent stem cells, whose niche is
localized at the level of the CoH, the boundary structure con- Proliferation of cholangiocytes is fundamental not only for the
necting the hepatocytic bile canaliculi with the terminal bile maintenance of the normal homeostasis of the biliary tree, but
ducts. Although this concept is currently being challenged, also in response to liver damage. Proliferation of biliary epithe-
HPCs are thought to be able to expand and differentiate into lial cells (BEC) is mediated by the autocrine and/or paracrine
hepatocytes, through the formation of cells with intermediate action of several growth factors and hormones released in the
phenotype (intermediate hepatobiliary cell, IHBC), or in chol- local microenvironment by inflammatory and mesenchymal
angiocytes, or in RDC as a response to non‐resolving damage cells or by epithelial cells themselves. One of the critical
[51]. RDCs’ phenotype is less differentiated than normal chol- cytokines that stimulate cholangiocyte proliferation is IL6,
angiocytes and expresses molecular markers of immaturity such together with other members of the IL6 family, such as
as chromogranin A, neural cell adhesion molecule (NCAM), Oncostatin M [68, 69]. Other growth factors significantly
and B‐cell protein leukemia lymphoma‐2 (Bcl‐2). Interestingly, involved in BEC proliferation include hepatocyte growth factor
NCAM mediates the interaction between cells and matrix dur- (HGF) [70], insulin‐like growth factor (IGF)‐1 [71], epidermal
ing the development of different epithelial tissues, while Bcl‐2 growth factor (EGF) [72], and VEGF‐A [13, 14]. This response
inhibits apoptosis during development and the growth of new is sustained also by different hormones such as secretin, a major
epithelial structures  [61]. The increase in RDCs is associated activator of intracellular cAMP, or estrogen, testosterone, prol-
with a significant increase in inflammatory infiltrate and portal actin, and follicle stimulating hormone (FSH). This mechanism
fibrosis, as demonstrated in viral hepatic diseases [62], meta- is finely counterbalanced by the action of antiproliferative pep-
bolic disorders [63], or disorders of embryogenesis, such as tides such as gastrin, melatonin, somatostatin, serotonin, and
Alagille syndrome and biliary atresia [22, 63]. Deposition of gamma aminobutyric acid (GABA) [73].
fibrosis is based on a paracrine cross‐talk mediated by the abil- Genetic diseases of the biliary epithelium, such as ADPKD
ity of RDCs to secrete profibrotic and proinflammatory growth may be considered as a model to study the role played by some
factors including interleukin‐6 (IL‐6), TGFβ2, endothelin‐1 of these growth factors in stimulating cholangiocyte prolifera-
(ET‐1), monocyte chemotactic protein‐1 (MCP‐1), platelet‐ tion. VEGF, angiopoietins, and their related receptors are over-
derived growth factor‐B (PDGF‐B), TNF‐α, NO, vascular expressed in the biliary epithelium of ADPKD. ADPKD is
endothelial growth factor (VEGF) [8, 60], and connective tissue caused by mutations in either PKD1 or PKD2 genes, encoding
growth factors (CTGF). These are among many other factors for PC1 and PC2, and are characterized by the formation of
and cyto/chemokines that send signals able to recruit inflamma- multiple cysts in the kidney and in 90% of the cases of bile duct‐
tory and immune cells, mesenchymal cells, and endothelial derived cysts [74]. The production of angiogenic factors by the
cells. RDC are the masterminds of these dynamic interactions cystic epithelium is a recapitulation of biliary development and
and are therefore considered the “pacemakers” of portal fibrosis a feature of de‐differentiation. In fact, during the DP stage, the
[64]. In addition to inflammatory cells, of relevance is the cross‐ developing epithelium secretes VEGF that acts on the endothe-
talk with cells of mesenchymal origin, in particular Kupffer lial cell precursors and promotes the peribiliary vascularization.
cells and portal fibroblasts, which have the role as the main In the same way, in ADPKD the growing cystic epithelium
effectors of fibrosis, as stimulators of the deposition of extracel- secretes VEGF to support the expansion of the surrounding
lular matrix (ECM). In addition, RDC also establish paracrine ­vascular supply [14].
communications with endothelial cells that provide the vascular The relationship between angiogenic signaling and cholan-
support necessary for the growth and arborization of the ductal giocyte proliferation in ADPKD was demonstrated by studies in
structures themselves [13]. Most of these factors are also tran- PC2 defective mice. PC2 is present on the surface of the primary
siently expressed by the DP during fetal development, an obser- cilium and of different organelles of the cholangiocytes of
vation in line with the hypothesis that the DR recapitulates, unclear function, but is involved also in calcium homeostasis in
morphogenetic signaling patterns observed during hepatic the endoplasmic reticulum (ER), cilium, and cytoplasm. In
ontogenesis [13] (Figure 32.2). physiologic conditions, bile flow stimulates the activation of
An important novel stem cell niche is represented by the Ca2+ signaling by a rapid release of Ca2+ from the ER mediated
“peribiliary glands”. These glands are located in the area of the by IP3R, followed by a sustained Ca2+ entry from the plasma
common hepatic duct at the hepatic hilum, in the cystic duct, membrane. This mechanism is also known as store operated
and in the hepatic papilla [65]. These glands contain cells Ca2+ entry (SOCE). When PC2 is defective, store activated cal-
expressing a number of markers and transcription factors cium signaling is altered and cytoplasmic Ca2+ levels are lower.
32:  Cholangiocyte Biology and Pathobiology 399

Figure 32.2  Morphogens involved in epithelial mesenchymal interactions in ductular reaction. In response to liver damage, HPC and RDC com-
partments start to proliferate to restore the normal structure of the liver; the abnormal increase of these structures are responsible for the accumula-
tion of peribiliary fibrosis and of portal inflammatory infiltrates that in turn further stimulate this anomalous process. MF surrounding RDC
stimulates the development of an increased dysmorphic biliary bed by the activation of the Notch2/Jag1 signaling and by secreting Hh ligands that
dock to their receptor Ptc on the surface of the cholangiocytes. Moreover, TGFβ exerts its trophic properties on RDC by binding its specific receptor
TGFβR. Abbreviations: CoH, canal of Hering; Hep, hepatocyte; Hh, hedgehog; HPC, hepatic progenitor cells; BD, bile duct; RDC, reactive ductu-
lar cell; MF, myofibroblast; Ptc, Patched.

This activates adenyl cyclase 5 (AC5) a Ca2+ inhibitable AC, CTGF, leading to the activation of mitogen‐activated protein
which, through PKA, phosphorylates ERK1/2 that upregulates kinase (MAPK) signaling that sustains cell proliferation [78].
HIF1α, a potent inducer of VEGF‐A production by BEC. Notably, perturbation of YAP‐mediated signaling leads to
VEGF‐A, in turn, through an autocrine loop binds its receptor ­biliary paucity in mouse embryos at E18.5, whilst its hyperacti-
VEGFR2 present on the surface of the PC2 defective cholangio- vation, using mutant YAP constitutively expressed into the
cytes, which further sustains the proliferation of the BEC acti- nucleus, stimulates an uncontrolled liver growth following
vating the Raf/MEK/ERK1/2 pathway. Furthermore, PKA ­partial hepatectomy in mice [26].
activation inhibits tuberin, a repressor of mammalian target of
rapamycin (mTOR), a downstream effector of the IGF‐1R that,
following activation by its ligand IGF‐1, stimulates cyclins to
promote BEC proliferation [75–77]. CHOLANGIOCYTES AND IMMUNITY
A novel transcription factor controlling cholangiocyte prolif-
eration (as well as liver size) is YAP, a transcription factor Cholangiocytes are a first line of defense in liver innate immu-
belonging to the Hippo pathway. YAP is present in a phospho- nity and can present the antigen to immune cells, can be a target
rylated and inactive state in the cytosol of cholangiocytes. Once of immune‐mediated aggression, or be the initiators of an
dephosphorylated by LATS1/2 in response to cytoskeletal dis- inflammatory reaction that then progresses to adaptive immune
tress or other stimuli, it translocates into the nucleus and stimu- activation. Contribution of biliary epithelial cells to liver
lates the transcription of several downstream effectors, such as immune responses was believed to be limited to the secretion of
400 THE LIVER:  CHOLANGIOCYTES AND IMMUNITY

immunoglobulin (Ig) A into the bile [79–80], but it is now clear The biliary epithelium is often a target for inflammation or
that the role of cholangiocytes in the immune response is far immune‐mediated injury. In inflammatory cholangiopathies,
more complex [81, 82]. In the liver, IgAs are synthesized by as  discussed above, innate and adaptive immune responses
plasma cells located along the biliary tree, bind to polymeric are sequentially or simultaneously involved. Innate immune
immunoglobulin receptor (pIgR) present on the surface of the responses can be triggered by exogenous or endogenous insults
basolateral membrane of BEC, and then are transported into to cholangiocytes or nearby hepatocytes and if the inflammatory
the cytoplasm of cholangiocytes where they are secreted into the reaction is protracted, it can be perpetuated through adaptive
bile by vesicular transport [83]. immune mechanisms. A number of studies have shown that cell
The biliary epithelium stands as a first line of defense against dysfunction caused by genetic defects, causing cellular or tissue
bacteria, fungi, and other pathogens by secreting antimicrobial malfunction, can generate a chronic inflammation of low mag-
peptides, such as defensin and cathelicidin; normal cholangio- nitude, defined as “parainflammation”, that is an adaptive
cytes, in fact, express high concentration of β‐defensins hBD1 response to a persistent cell dysfunction shifting the homeo-
and hBD3. hBD1 is particularly effective in exerting an antimi- static set points aimed at restoring a normal cell/tissue homeo-
crobial effect against Gram negative bacteria and Pseudomonas stasis [94]. As proposed by Medzihitov, parainflammation in
aeruginosa, whereas hBD3 have a pronounced activity against spite of its homeostatic nature may become maladaptive and
Helicobacter pylori and Staphylococcus aureus [84]. In con- may stimulate a fibrotic response. In this scenario, epithelial
trast, hBD2 is induced by an inflammatory status or by exposure cells secrete a number of molecules and factors able to instruct
to inflammatory peptides such as TNFα and IL1β, or to infec- immune cells or to generate an inflammatory reaction (the
tious agents, like Cryptosporidium parvum, through a TLR2/ epi‐immunome).
TLR4‐dependent NF‐κB activation [85]. An example is CFLD, where a decrease in the LPS tolerance
A major role in epithelial innate immunity in cholangiocytes plays a major role in the development of the disease. Treatment
is played by toll‐like receptors (TLRs) [85], and by nuclear of mice with dextran sulfate sodium (DSS), a chemical that, by
receptors (NR) [86]. TLRs can recognize pathogen‐associated inducing colitis, increases intestinal permeability and transloca-
molecular patterns (PAMPs), namely bacterial structural ele- tion of gut endotoxins to the liver, causes a peribiliary inflamma-
ments such as LPS, DNA, RNA fragments, and flagellin but tion with infiltration of CD45+ cells, mainly neutrophils and
also respond to endogenous components or damage‐associated macrophages that damage the biliary epithelium and promotes
molecular patterns (DAMPs), such as hyaluronan and HMGB1 DR in CFTR‐KO mice but not in their wild‐type (WT) litter-
[87–89] that are released from damaged cells. mates. Cholangiocytes isolated from CFTR‐KO mice and
The TLRs family is composed by nine (1 to 9) transmem- exposed to LPS show an aberrant activation of TLR4/NF‐κB
brane receptors with an intracellular domain, Toll‐IL1 receptor signaling that leads to the increased secretion of proinflamma-
(TIR) that can bind different adaptors, such as MyD88, Mal, tory cytokines such as MCP‐1, G‐CSF, MIP‐2, KC, and lipopol-
TRIF, and TRAM and thereby activate the NF‐κB signaling ysaccharide‐induced CXC chemokine (LIX) [95]. As discussed,
which is then responsible for the secretion of a number of in normal conditions CFTR acts as a docking protein that main-
inflammatory peptides [87–89]. Normal cholangiocytes consti- tains proteins (i.e. Cbp, Csk) involved in the negative regulation
tutively express TLR2, 3, 4, 5, and 9 [90]: TLR2 is a receptor for of Src tyrosine kinase at the membrane. CFTR mutations pre-
the lipoteichoic acid (LTA), a component of the bacterial mem- venting the expression of CFTR at the apical membrane cause
brane, TLR3 senses viral dsRNA, TLR4 is the main sensor for the disaggregation of the complex, and the activation of the
LPS and DAMPS, and TLR9 is a receptor for CpG DNA, a hall- kinase that phosphorylates TLR4 and increases its response to
mark of bacterial invasion. LPS. Notably, the treatment of CF‐KO mice exposed to DSS
TLR4‐mediated signaling is better known and studied in with PP2, an inhibitor of Src family kinases, reduces the accu-
cholangiocytes; once activated by LPS or other ligands, TLR4 mulation of CD45+ cells and the extent of DR. Similar mecha-
activates two different pathways, one mediated by NF‐κB and nisms were further confirmed in human cholangiocytes derived
an alternative one via MAPK/AP‐1. The first one stimulates the from induced pluripotent stem cells (iPSC) of a cystic fibrosis
expression of a number of proinflammatory cyto‐ and patient with the common mutation ΔF508 CFTR [96]. Human
chemokines including TNFα, IL1, IL6, IL8, IL12, G‐CSF, and CF cholangiocytes show an increased activation of Src, aberrant
LIX [89, 91]. The second pathway requires the nuclearization of activation of NF‐κB, sustained secretion of Mcp‐1 and IL‐8, and
the AP‐1 complex [92]. cytoskeletal defects. Treatment with PP2 improves these patho-
In normal cholangiocytes, TLR4 signaling is repressed by logic features and remarkably, increases the efficacy of VX‐770
­protective mechanisms intended to maintain a level of “LPS toler- and VX‐809, two small compounds used in clinic to correct the
ance” such as the expression of negative regulators (i.e. interleu- ΔF508 defect, to restore the secretory function [96] (Figure 32.3).
kin‐1 receptor‐associated kinase M (IRAK‐M)) and PPARγ or by A group of factors whose involvement in the immune
post‐translational regulation of the receptor (i.e. Src mediated response of cholangiocytes has been recognized only recently,
tyrosine phosphorylation) (see below). Also, upregulation of are the nuclear receptors (NRs) that bind the retinoic X receptor
miR‐146 in response to LPS has been described as a mechanism (RXR). This superfamily is composed of several members,
that partially controls the activation of TLR4‐induced NF‐κB‐ among which are the glucocorticoid receptor (GR), the retinoic
mediated cytokine secretion [93]. Since the biliary epithelium is acid receptor (RAR), the vitamin D receptor (VDR), the liver X
continuously in contact with bacterial products coming from the receptors (LXRs), and the peroxisome proliferator‐activated
intestine, changes in one or more regulatory checkpoints may receptors (PPARs). NRs control several cell functions including
elicit an exaggerated inflammatory response in the liver. cell proliferation and apoptosis, cell metabolism, cell–cell
Figure 32.3  Control of innate immune‐mediated inflammation in cholangiocytes is altered in cystic fibrosis cholangiocytes. Healthy cholangiocytes (A), once stimulated by pathogenic noxa, such
as bacterial degradation products (PAMPs), activate a TLR4‐mediated response that transiently allows the secretion of profibrotic and proinflammatory mediators. This response is downmodulated by
several factors, among which, miR‐146, that inhibits TLR4, and by PPARγ, that through its effector IκBα, hinders the nuclear translocation of the p65 subunit of the NF‐κB complex. Furthermore,
polarized normal cholangiocytes express CFTR at the apical membrane and CFTR is able to assemble a macromolecular complex with EBP‐50, Cbp, and Csk, that are able to prevent the tyrosine
kinase Src form phosphorylating TLR4. In CFLD (B), PAMPs (i.e. LPS, or LTA) and DAMPs (i.e. HMGB1) coming from the gut or released by nearby cells, stimulate the secretion of profibrotic and
proinflammatory cyto‐ and chemokines through the TLR4/NF‐κB axis. The absence of CFTR sustains and amplifies this response, allowing the aberrant phosphorylation of Src that, by binding TLR4,
continuously stimulates the secretion of fibroinflammatory mediators and of hBD2. Similarly to TLR4, TLR2 induces the disaggregation of p65, its nuclear import, and its transcriptional activity.
Abbreviations: BEC, biliary epithelial cells; CF, cystic fibrosis; TLR, toll‐like receptor; DAMPs, damage‐associated molecular patterns; PAMPs, pathogen‐associated molecular patterns; LPS,
lipopolysaccharides; LTA, lipoteichoic acid; hBD2, human β‐defensin 2.
402 THE LIVER:  PORTAL AND PERIBILIARY FIBROSIS

interaction, detoxification from bile acids (VDR, FXR), and bile the activation of the P38MAPK/PKC signaling leading to the
secretion (GR, FXR) [86]. Interest in their role in inflammation expression of p16/pRB. Once senescent, cells not only stop their
and innate immunity is raised, particularly for the PPARs (i.e. proliferation, but assume a senescence‐associated secretory
PPARα, PPARβ/δ, and PPARγ). Among the different isoforms, phenotype (SASP) characterized by the secretion of a plethora
PPARγ has been shown to negatively regulate TLR4‐NF‐κB of peptides with profibrogenic, proinflammatory, and tumori-
signaling in CF cholangiocytes. Activation of PPARγ by piogl- genic properties, indicating that senescence could not only act
itazone and rosiglitazone inhibits LPS‐induced activation of as a barrier to tumor growth, but also paracrinally stimulate
NF‐κB and cytokine secretion in Cftr‐KO cholangiocytes and the activation of aberrant reparative/regenerative responses
reduces biliary damage and inflammation in Cftr‐KO mice [103, 104].
treated with DSS. PPARγ can inhibit the transcriptional activity Cholangiocytes can act as antigen presenting cells (APC).
of NF‐κB in two different manners, that is, by inhibiting p65 Cholangiocytes normally express only low levels of human leu-
nuclear entry directly binding the p65/p50 complex, or cocyte antigen (HLA) class I, while HLA class II and the mem-
indirectly, by stimulating IkBα expression that retain p65 in the brane co‐receptors CD80 and CD86 are undetectable. However,
cytoplasm [97]. when cholangiocytes are challenged in vitro with the proinflam-
It is worthy to note that autoimmune responses may be con- matory cytokines IL1, IFNγ, and TNFα or infected by cytomeg-
sequences of alterations of innate immunity. In fact, continuous alovirus, HLA class I are upregulated and HLA class II but not
stress in the absence of correct modulation of the TLR‐mediated CD80 and CD86 can be expressed de novo, allowing the presen-
responses could be the trigger for a chronic inflammatory or tation of the antigens to CD4+ and CD8+ T cells. Interestingly,
autoimmune response. Similarly, PSC that was typically consid- contrary to in vitro cell cultures, human specimens from PBC
ered as a disease with autoimmune traits may in reality be an and PSC show the neo‐expression of CD86 by RDC indicating
autoinflammatory disease on the basis of its increased activation that cholangiocytes can orchestrate the necroinflammatory fea-
of NF‐κB transcription factor and its exaggerated activation of tures of these two diseases [105]. More recently, two papers
the Th17 response to pathogens [98]. Another necroinflamma- show that cholangiocytes from BA display the expression of
tory disease characterized by an increased Th17 response is bil- MHC class I and II, CD40, the co‐stimulatory molecules B7‐1
iary atresia (BA), a severe pediatric disease characterized by and B7‐2 [106], and MHC class I‐like molecule CD1d [107].
fibro‐obliterative vanishing of the bile ducts and inflammation
whose etiopathogenesis is still unclear. In BA, the serum levels
of IL17a and IL23, two typical Th17 response‐related ILs are
augmented, together with the expression of the IL17 transcrip- PORTAL AND PERIBILIARY FIBROSIS
tion factor ROR‐γt, IL‐17α, IL‐1β, IL‐6 [99]. Together with the
Th17 response, in the case of Ross River virus (RRV) infection, Liver fibrosis is the result of an aberrant reparative response to
BA patients could mount a Th2 adaptive response, theoretically liver injury, and the mechanism supporting the clinical and his-
able to respond to the viral infection, but also to cause BA. topathological progression of chronic liver diseases regardless
Recent data have shown that IL‐33 secreted by cholangiocytes of their etiology. Fibrosis is a complex and highly integrated
in responses to RRV infection stimulate innate lymphoid cells pathophysiologic process, which is achieved through the inter-
type 2 (ILC2) to secrete a range of ILs and growth factors such action of multiple cell types and various molecular factors, lead-
as IL‐13, osteopontin, and TGFβ. IL‐13 and IL‐33 behave in a ing to an exaggerated and qualitatively altered deposition of
paracrine loop exerting trophic effects on cholangiocytes and ECM proteins [108].
increasing the extension of ductular reaction in the disease Understanding the fine mechanisms that regulate portal fibro-
[100, 101], while local release of TGFβ stimulates proliferation genesis, is a key step in the effort to identify new therapeutic
and secretion of ECM proteins by myofibroblasts (MF) leading targets for biliary diseases (Figure  32.4). In the context of a
to the obliteration of bile ducts. chronic hepatic insult, the persistent necroinflammatory dam-
In addition, presence of DAMPs and PAMPs could promote age determines the progressive loss of parenchyma by apoptosis
cellular senescence. Cell senescence is a mechanism of irrevers- or necrosis and acts as a stimulus for fibrosis. Necrotic and
ible cell arrest in G1 stage induced by different stimuli, in par- apoptotic cells release cytokines and chemokines and DAMPs
ticular, by oncogenic stressors. The main causes responsible for that induce the recruitment of inflammatory and mesenchymal
the onset of senescence on cells include DNA damage, in par- cells, which in turn secrete a wide range of profibrotic and pro-
ticular, but not exclusively, to the telomers, the activation of inflammatory factors, including TGFβ, CTGF, and TNFα [108,
mitogenic signals induced by the activation of oncogenes, epi- 109]. In addition, the necrotic cells release genetic material and
genetic modifications, and expression of tumor suppressor debris directly into the intercellular space, where they activate
genes. All these signals lead to different physiological responses stellate cells (HSC) and portal fibroblasts that proliferate and
generally leading to tumor suppression, but, in other cases could secrete ECM proteins [110]. More information on epithelial–
also promote cancer development, induce a fibrosing response, mesenchymal cross‐talk in biliary disease can be found in recent
and mediate age‐related degenerative diseases. Regardless of reviews [111, 112].
the stimulus, cell senescence is mediated by two major signal In hepatic fibrogenesis, the main actor is TGFβ1, which is
pathways, p16INK4a/pRB and the p53/p21CIP1/WAF1 [102]. In par- secreted mainly by fibroblasts and macrophages and functions
ticular, genomic or epigenomic damages could induce a persis- to activate HSC, responsible for the deposition of about 80% of
tent DNA damage response (DDR) that leads, on one side, to the the matrix proteins [109]. Under physiological conditions the
direct activation of the p53/p21pathway, and on the other side to ECM is composed mainly of collagens (the most important of
32:  Cholangiocyte Biology and Pathobiology 403

Figure 32.4  Mechanism of peribiliary fibrosis deposition and ECM modification in cholangiopathies. Following cholangiocyte damage, biliary
epithelial cells secrete a wide range of mediators stimulating, in a paracrine fashion, the proliferation, transdifferentiation, and activation of cells of
mesenchymal origin (i.e. HSC and PF) responsible for the deposition of fibrosis. Furthermore, cholangiocytes secrete chemokines involved in
immune cell recruitment. Biliary epithelial cells are also able to modify the structure of the hepatic scaffold by secreting a wide range of metallo-
proteinases and proteolytic enzymes. Abbreviations: DAMPs, damage‐associated molecular patterns; PAMPs, pathogen‐associated molecular pat-
terns; HSC, hepatic stellate cells; PF, portal fibroblast; MF, myofibroblast; Φ, macrophage.

which are types I, III, IV, and VI) and non‐collagen proteins due to an increased and aberrant secretion of fibrosis produced
such as laminin, fibronectin, thrombospondin, and tenascin, and by activated portal myofibroblasts, macrophages (mainly M2),
of various classes of proteoglycans consisting of a protein core and cholangiocytes too [116, 117], and the erosion of the basal
covalently linked to glycosaminoglycan (GAG) such as chon- membrane, composed of collagen IV, laminin, and reticulin,
droitin‐, dermatan‐ and keratan‐sulfate, heparin and heparan‐ surrounding the biliary epithelium [118]. These changes could
sulfate, and hyaluronic acid (HA) [113]. In response to the be persistent as in late fibrosis or cirrhosis, or could be transient
damage, quiescent HSCs transdifferentiate into MF, character- as in early stages of alcoholic liver disease [119]. In addition to
ized by the neo‐expression of α‐actin smooth muscle (α‐SMA), the ECM proteins, MFs are able to produce a broad spectrum of
by sensitivity to fibrogenic, chemoattractant and mitogenic proteins involved in matrix degradation, such as metallopro-
stimuli, and by the ability to actively secrete ECM proteins teases (MMPs) (in particular MMP‐2 and ‐9) and their inhibi-
[109, 114], in particular laminin and fibrillar collagen (type I tors (TIMPs), which with a finely regulated mechanism, lead to
and III). This causes an alteration of the normal composition of continuous remodeling of the matrix during chronic hepatic
the matrix, which is important for the different tissue functions, injury. In the early stages of hepatic injury, the MFs do not
including cell–matrix interactions and the deposition of growth express TIMPs, thus allowing the degradation of the normal
factors [113]. Following intoxication with CCl4 or DDC, com- matrix that will be replaced by the altered one [120].
position of liver ECM components changes dramatically, simi- Cholangiocytes are able to secrete MMPs, in particular
lar to what is observed in human diseases. In particular collagens MMP2, 3, 13, and MT1‐MMP and other enzymes involved in the
I, IV, and V, and fibronectin were increased while content of modification of the hepatic 3D scaffold, such as uPA; notably, in
elastin decreased [115]. Human ARPKD/CHF/CD (see below) a rodent model of fibropolycystic liver diseases, the treatment of
also exhibits a derangement of the normal composition of ECM 8‐week‐old PCK rats with marimastat, a broad spectrum MMP
404 THE LIVER:  REFERENCES

inhibitor, was able to reduce both cyst area and col1a1 expres- TNFα that stimulate the de novo expression of αVβ6 integrin
sion [120]. At a certain point, the production of MMPs is blocked, on the surface of the biliary structures. This integrin is able to
so the remodeling ceases to the advantage of the excessive depo- activate the latent form of TGFβ by the cleavage of the latency‐
sition. In addition, MF further release growth factors that main- associated peptide (LAP). Notably, the treatment of FPC‐KO
tain fibrogenesis, and also intervene in the activation of mice with clodronate, a bisphosphonate that inhibits the mono-
neoangiogenesis. However, HSCs are not the only mesenchymal cyte–macrophage differentiation, improves the main patho-
cells capable of giving rise to MF: alternative sources exist, such logic readouts of the disease, namely, cyst enlargement, fibrosis
as activated portal fibroblasts, circulating mesenchymal precur- accumulation, and development of portal hypertension [116].
sors of medullary origin (5–7%) and fibrocytes (4–6%) [114]. Interestingly, the secretion of CXCL10 was induced not only
Whether epithelial–mesenchymal transition (EMT) [121] by a β‐catenin‐dependent mechanism but also by the secretion
plays a role in liver fibrosis similar to other organs like the kid- of IL1β, thanks to the NF‐κB dependent activation of the
ney has been hotly debated. Most evidence indicates that during NLRP3 inflammasome complex [126]. Moreover, the treat-
liver repair, cholangiocytes lose part of their epithelial charac- ment of Pkhd1del4/del4 mice with AMG‐487 an inhibitor of the
teristics to acquire some biological and immunophenotypic receptor for CXCL10 is able to reduce the extent of pericystic
properties of mesenchymal cells, such as motility and the ability macrophages together with cystic areas and peribiliary fibrosis
to depose collagen, but there is not a transition to a full mesen- in a similar manner to clodronate, demonstrating that mac-
chymal phenotype [122] and cholangiocytes simply acquire a rophage accumulation is one of the major drivers of the disease
few functional properties that favor the repair of the wound. in CHF [126].
CTGF is another growth factor capable of modulating the
transduction of extracellular signals by interacting with mem-
brane glycoproteins (integrins), growth factors (TGFβ1), and
with different components of the extracellular matrix (fibronec- CONCLUSIONS
tin). Once secreted, CTGF can stimulate the proliferation, adhe-
sion, and recruitment of different cell types [123], including In this chapter, we have summarized some of the recent
macrophages; CTGF can play a fundamental role in supporting advances in the understanding of cholangiocyte biology. We
the inflammatory response driven by macrophages. PBC and PSC elected to focus our discussion on those mechanisms that have
are characterized by increased CTGF secretion by different cell relevance for understanding the pathophysiology of the most
milieu, among which are cholangiocytes, in which expression relevant genetic or acquired cholangiopathies. We believe that
sustains the deposition of ECM proteins and fibrosis. Recently, a the reader will have a working understanding of the function
paper by Pi and colleagues [124] demonstrates that in the liver, and dysfunction of this important epithelium of the liver and
CTGF, in association with integrin αvβ6, is expressed by HPCs to focus their research and interest on finding novel therapeutic
and RDCs after experimental biliary damage, where they regulate targets that are urgently needed. We expect that future research
the activation of HPC and deposition of fibrosis, by interacting into the pathophysiology of these fascinating diseases will be
with fibronectin and TGF‐β1. Notably, in a mouse model of CCl4 rewarding.
intoxication, the treatment with siRNA against CTGF was able to
revert fibrosis and reduce collagen deposition [125].
Among genetic cholangiopathies of particular interest are
fibropolycystic diseases, as they are characterized by exuberant
ACKNOWLEDGMENTS
biliary fibrosis, but in the absence of cholangiocyte necrosis/
Fondazione Cariplo, grant number 2014–1099, to M. Cadamuro;
apoptosis. This is a group of monogenic diseases with autoso-
NIH grant DK‐034989 (Silvio O. Conte Digestive Diseases
mal recessive transmission and variable phenotype that includes
Research Core Centers) to M. Strazzabosco; NIH grants DK‐079005
the autosomal recessive disease of polycystic kidney (ARPKD),
and DK‐096096, and PSC Partners Seeking a Cure and Connecticut
congenital hepatic fibrosis (CHF), and Caroli’s disease (CD),
Innovations (16‐RMA‐YALE‐26) to M. Strazzabosco.
all due to the mutation, in different loci, of the same gene,
which encodes for fibrocystin (FPC). Fibropolycystic liver dis-
eases are characterized by ductal dysgenesia resulting in biliary
microhamartomas and segmental dilations, with a phenotype REFERENCES
retaining a fetal configuration (DP malformations). Biliary
dysgenesis is typically associated with progressive accumula- 1. Strazzabosco, M., Spirlí, C., and Okolicsanyi L. Pathophysiology of the intra-
tion of portal fibrosis eventually leading to portal hypertension. hepatic biliary epithelium. J Gastroenterol Hepatol, 2000;15:244–53.
Interestingly, these diseases are characterized by “parainflam- 2. Cheung, A.C., Lorenzo Pisarello, M.J., and LaRusso, N.F. Pathobiology of
biliary epithelia. Biochim Biophys Acta, 2018;1864:1220–1231.
mation”, a smoldering and unresolved inflammatory process 3. Crawford, A.R., Lin, X.Z., and Crawford, J.M. The normal adult human liver
indicating a maladaptive response to persistent tissue dysfunc- biopsy: a quantitative reference standard. Hepatology, 1998;28:323–31.
tion [94]. Increased nuclear expression of β‐catenin is respon- 4. Roskams, T.A., Theise, N.D., Balabaud, C. et al. Nomenclature of the finer
sible for the secretion of several chemokines, among which are branches of the biliary tree: canals, ductules, and ductular reactions in human
livers. Hepatology, 2004;39:1739–45.
CXCL1, CXCL10, and CXCL12. Chemokines secreted in the
5. Lemaigre, F.P. Molecular mechanisms of biliary development. Prog Mol Biol
hepatic microenvironment are responsible for the recruitment Transl Sci, 2010;97:103–26.
of inflammatory cells around the cysts, and in particular of M1 6. Ober, E.A. and Lemaigre, F.P. Development of the liver: insights into organ
and M2 macrophages. Macrophages locally secrete TGFβ and and tissue morphogenesis. J Hepatol, 2018;68:1049–62.
32:  Cholangiocyte Biology and Pathobiology 405

7. Lazaridis, K.N., Strazzabosco, M., and Larusso N.F. The cholangiopathies: 32. Kanno, N., LeSage, G., Glaser, S., Alvaro, D., and Alpini, G. Functional
disorders of biliary epithelia. Gastroenterol, 2004;127:1565–77. heterogeneity of the intrahepatic biliary epithelium. Hepatology, 2000;
8. Strazzabosco, M., Fabris, L., and Spirli, C. Pathophysiology of cholangiopa- 31:555–61.
thies. J Clin Gastroenterol, 2005;39:S90–102. 33. Marzioni, M., Glaser, S.S., Francis, H., Phinizy, J.L., LeSage, G., and Alpini,
9. Raynaud, P., Carpentier, R., Antoniou, A., and Lemaigre, F.P. Biliary differentia- G. Functional heterogeneity of cholangiocytes. Semin Liver Dis,
tion and bile duct morphogenesis in development and disease. Int J Biochem 2002;22:227–40.
Cell Biol, 2011;43:245–56. 34. Glaser, S., Francis, H., Demorrow, S. et al. Heterogeneity of the intrahepatic
10. Gordillo, M., Evans, T., and Gouon‐Evans V. Orchestrating liver develop- biliary epithelium. World J Gastroenterol, 2006;12:3523–36.
ment. Development, 2015;142:2094–108. 35. Strazzabosco, M. and Fabris, L. Functional anatomy of normal bile ducts.
11. Libbrecht, L., Spinner, N.B., Moore, E.C., Cassiman, D., Van Damme‐ Anat Rec, 2008;291:653–60.
Lombaerts, R., and Roskams, T. Peripheral bile duct paucity and cholestasis 36. Han, Y., Glaser, S., Meng, F. et al. Recent advances in the morphological and func-
in the liver of a patient with Alagille syndrome: further evidence supporting tional heterogeneity of the biliary epithelium. Exp Biol Med, 2013;238:549–65.
a lack of postnatal bile duct branching and elongation. Am J Surg Pathol, 37. Boyer J.L. Bile formation and secretion. Compr Physiol, 2013;3:1035–78.
2005;29:820–6. 38. Nathanson, M.H., Burgstahler, A.D., Mennone, A., and Boyer J.L.
12. Carpentier, R., Suñer, R.E., van Hul, N. et al. Embryonic ductal plate cells Characterization of cytosolic Ca2+ signaling in rat bile duct epithelia. Am J
give rise to cholangiocytes, periportal hepatocytes, and adult liver progenitor Physiol, 1996;271:G86–96.
cells. Gastroenterol, 2011;141:1432–8. 39. Zsembery, A., Spirlì, C., Granato, A., et al. Purinergic regulation of acid/base trans-
13. Fabris, L., Cadamuro, M., Libbrecht, L. et al. Epithelial expression of angio- port in human and rat biliary epithelial cell lines. Hepatology, 1998;28:914–20.
genic growth factors modulate arterial vasculogenesis in human liver devel- 40. Caligiuri, A., Glaser, S., Rodgers, R.E. et al. Endothelin‐1 inhibits secretin‐
opment. Hepatology, 2008;47:719–28. stimulated ductal secretion by interacting with ETA receptors on large chol-
14. Fabris, L., Cadamuro, M., Fiorotto, R. et  al. Effects of angiogenic factor angiocytes. Am J Physiol, 1998;275:G835–46.
overexpression by human and rodent cholangiocytes in polycystic liver diseases. 41. Gong, A.Y., Tietz, P.S., Muff, M.A. et al. Somatostatin stimulates ductal bile
Hepatology, 2006;43:1001–12. absorption and inhibits ductal bile secretion in mice via SSTR2 on cholan-
15. Lemaigre F.P. Mechanisms of liver development: concepts for understanding giocytes. Am J Physiol Cell Physiol, 2003;284:C1205–14.
liver disorders and design of novel therapies. Gastroenterol, 2009;137:62–79. 42. Mennone, A., Biemesderfer, D., Negoianu, D. et al. Role of sodium/hydro-
16. Nakanuma, Y., Harada, K., Sato, Y., and Ikeda, H. Recent progress in the gen exchanger isoform NHE3 in fluid secretion and absorption in mouse and
etiopathogenesis of pediatric biliary disease, particularly Caroli’s disease rat cholangiocytes. Am J Physiol Gastrointest Liver Physiol, 2001;280:
with congenital hepatic fibrosis and biliary atresia. Histol Histopathol, G247–54.
2010;25:223–35. 43. Elborn, J.S. Cystic fibrosis. Lancet, 2016;388:2519–31.
17. Desmet, V.J. Ductal plates in hepatic ductular reactions. Hypothesis and 44. Beuers, U., Hohenester, S., de Buy Wenniger, L.J., Kremer, A.E., Jansen,
implications. II. Ontogenic liver growth in childhood. Virchows Arch, P.L., and Elferink, R.P. The biliary HCO(3)(‐) umbrella: a unifying hypothesis
2011;458:261–70. on pathogenetic and therapeutic aspects of fibrosing cholangiopathies.
18. Desmet, V.J. Ductal plates in hepatic ductular reactions. Hypothesis Hepatology, 2010;52:1489–96.
and implications. III. Implications for liver pathology. Virchows Arch, 45. Hohenester, S., Maillette de Buy Wenniger, L., Jefferson, D.M., Oude Elferink,
2011;458:271–9. R.P., and Beuers, U. Biliary bicarbonate secretion constitutes a protective
19. Simons, M., Gloy, J., Ganner, A. et al. Inversin, the gene product mutated in mechanism against bile acid‐induced injury in man. Dig Dis, 2011;29:62–5.
nephronophthisis type, I.I., functions as a molecular switch between Wnt 46. Beuers, U., Maroni, L., and Elferink, R.O. The biliary HCO(3)(‐) umbrella:
signaling pathways. Nat Genet, 2005;37:537–43. experimental evidence revisited. Curr Opin Gastroenterol, 2012;28:253–7.
20. Lozier, J., McCright, B., and Gridley, T. Notch signaling regulates bile duct 47. Hohenester, S., Wenniger, L.M., Paulusma, C.C. et  al. A biliary HCO3‐
morphogenesis in mice. PLoS One, 2008;3:e1851. umbrella constitutes a protective mechanism against bile acid‐induced injury
21. Zong, Y., Panikkar, A., Xu, J. et al. Notch signaling controls liver development in human cholangiocytes. Hepatology, 2012;55:173–83.
by regulating biliary differentiation. Development, 2009;136:1727–39. 48. Beuers, U., Trauner, M., Jansen, P., and Poupon, R. New paradigms in the
22. Fabris, L., Cadamuro, M., Guido, M. et al. Analysis of liver repair mecha- treatment of hepatic cholestasis: from UDCA to FXR., PXR and beyond.
nisms in Alagille syndrome and biliary atresia reveals a role for notch signal- J Hepatol, 2015;62:S25–37.
ing. Am J Pathol, 2007;171:641–53. 49. Ainsworth, C. Cilia: tails of the unexpected. Nature, 2007;448:638–41.
23. Omenetti, A. and Diehl A.M. The adventures of sonic hedgehog in develop- 50. Bisgrove, B.W. and Yost H.J. The roles of cilia in developmental disorders
ment and repair. II. Sonic hedgehog and liver development, inflammation, and disease. Development, 2006;133:4131–43.
and cancer. Am J Physiol Gastrointest Liver Physiol, 2008;294:G595–8. 51. Strazzabosco, M. and Fabris, L. Development of the bile ducts: essentials for
24. Choi, S.S., Omenetti, A., Syn, W.K., and Diehl A.M. The role of Hedgehog the clinical hepatologist. J Hepatol, 2012;56:1159–70.
signaling in fibrogenic liver repair. Int J Biochem Cell Biol, 2011;43:238–44. 52. Clotman, F., Libbrecht, L., Killingsworth, M.C., Loo, C.C., Roskams, T., and
25. Machado, M.V. and Diehl A.M. Hedgehog signalling in liver pathophysiol- Lemaigre, F.P. Lack of cilia and differentiation defects in the liver of human
ogy. J Hepatol, 2018;68:550–62. foetuses with the Meckel syndrome. Liver Int, 2008;28:377–84.
26. Yi, J., Lu, L., Yanger, K. et al. Large tumor suppressor homologs 1 and 2 regu- 53. Kagan, K.O., Dufke, A., and Gembruch, U. Renal cystic disease and associated
late mouse liver progenitor cell proliferation and maturation through antago- ciliopathies. Curr Opin Obstet Gynecol, 2017;29:85–94.
nism of the coactivators YAP and TAZ. Hepatology, 2016;64:1757–72. 54. Schmidts M. Clinical genetics and pathobiology of ciliary chondrodyspla-
27. Swiderska‐Syn, M., Xie, G., Michelotti, G.A. et al. Hedgehog regulates yes‐ sias. J Pediatr Genet, 2014;3,46–94.
associated protein 1 in regenerating mouse liver. Hepatology, 2016; 55. Lehman, J.M., Michaud, E.J., Schoeb, T.R., Aydin‐Son, Y., Miller, M., and
64:232–44. Yoder, B.K. The Oak Ridge polycystic kidney mouse: modeling ciliopathies
28. Poncy, A., Antoniou, A., Cordi, S., Pierreux, C.E., Jacquemin, P., and of mice and men. Dev Dyn, 2008;237:1960–71.
Lemaigre F.P. Transcription factors SOX4 and SOX9 cooperatively control 56. Sambrotta, M., Strautnieks, S., Papouli, E. et al. Mutations in TJP2 cause
development of bile ducts. Dev Biol, 2015;404:136–48. progressive cholestatic liver disease. Nat Genet, 2014;46:326–28.
29. Mansini, A.P., Peixoto, E., Thelen, K.M., Gaspari, C., Jin, S., and Gradilone, 57. Lee, S.H. Intestinal permeability regulation by tight junction: implication on
S.A. The cholangiocyte primary cilium in health and disease. Biochim inflammatory bowel diseases. Intest Res, 2015;13:11–8.
Biophys Acta, 2018;1864:1245–53. 58. Fiorotto, R., Villani, A., Kourtidis, A. et al. The cystic fibrosis transmembrane
30. Rogler, C.E., Matarlo, J.S., Kosmyna, B., Fulop, D., and Rogler, L.E. conductance regulator controls biliary epithelial inflammation and permeabil-
Knockdown of miR‐23, miR‐27, and miR‐24 alters fetal liver development ity by regulating Src tyrosine kinase activity. Hepatology, 2016;64:2118–34.
and blocks fibrosis in mice. Gene Expr, 2017;17:99–114. 59. Fiorotto, R., Scirpo, R., Trauner, M. et  al. Loss of CFTR affects biliary
31. Rogler, C.E., Levoci, L., Ader, T., Massimi, A., Tchaikovskaya, T., Norel, R., epithelium innate immunity and causes TLR4‐NF‐κB‐mediated inflamma-
and Rogler L.E. MicroRNA‐23b cluster microRNAs regulate transforming tory response in mice. Gastroenterol, 2011;141:1498–508.
growth factor‐beta/bone morphogenetic protein signaling and liver stem cell 60. Fabris, L. and Strazzabosco, M. Epithelial‐mesenchymal interactions in
differentiation by targeting Smads. Hepatology, 2009;50:575–84. biliary diseases. Semin Liver Dis, 2011;31:11–32.
406 THE LIVER:  REFERENCES

61. Fabris, L., Strazzabosco, M., Crosby, H.A. et al. Characterization and isola- 86. Firrincieli, D., Zuniga, S., Poupon, R., Housset, C., and Chignard, N. Role
tion of ductular cells coexpressing neural cell adhesion molecule and Bcl‐2 of nuclear receptors in the biliary epithelium. Dig Dis, 2011;29:52–7.
from primary cholangiopathies and ductal plate malformations. Am J Pathol, 87. Barton, G.M. and Medzhitov, R. Toll‐like receptors and their ligands. Curr
2000;156:1599–612. Top Microbiol Immunol, 2002; 270:81–92.
62. Clouston, A.D., Powell, E.E., Walsh, M.J., Richardson, M.M., Demetris, 88. Iwasaki, A. and Medzhitov, R. Toll‐like receptor control of the adaptive
A.J., and Jonsson J.R. Fibrosis correlates with a ductular reaction in hepatitis immune responses. Nat Immunol, 2004; 5:987–995.
C: roles of impaired replication, progenitor cells and steatosis. Hepatology, 89. Seki, E. and Brenner, D.A. Toll‐like receptors and adaptor molecules in
2005;41:809–18. liver disease: update. Hepatol, 2008; 48:322–35.
63. Richardson, M.M., Jonsson, J.R., Powell, E.E. et al. Progressive fibrosis in 90. Di Gioia, M. and Zanoni, I. Toll‐like receptor co‐receptors as master regula-
non‐alcoholic steatohepatitis: association with altered regeneration and a tors of the immune response. Mol Immunol, 2015;63:143–52.
ductular reaction. Gastroenterol, 2007;133:80–90. 91. Akira, S., Takeda, K., and Kaisho T. Toll‐like receptors: critical proteins
64. Desmet, V.J. and Roskams, T. Cirrhosis reversal: a duel between dogma and linking innate and acquired immunity. Nat Immunol, 2001;2:675–80.
myth. J Hepatol, 2004;40:860–7. 92. Whitmarsh, A.J. and Davis, R.J. Transcription factor AP‐1 regulation by
65. de Jong, I.E.M., van Leeuwen, O.B., Lisman, T., Gouw, A.S.H., and Porte, mitogen‐activated protein kinase signal transduction pathways. J Mol Med
R.J. Repopulating the biliary tree from the peribiliary glands. Biochim (Berl), 1996;74:589–607.
Biophys Acta, 2018;1864:1524–31. 93. Soares, J.B., Pimentel‐Nunes, P., Roncon‐Albuquerque, R., and Leite‐
66. Cardinale, V., Wang, Y., Carpino, G. et al. Multipotent stem/progenitor cells Moreira, A. The role of lipopolysaccharide/toll‐like receptor 4 signaling in
in human biliary tree give rise to hepatocytes, cholangiocytes, and pancreatic chronic liver diseases. Hepatol Int. 2010;4:659–72.
islets. Hepatology, 2011;54:2159–72. 94. Medzhitov, R. Origin and physiological roles of inflammation. Nature,
67. Carpino, G., Cardinale, V., Renzi, A. et  al. Activation of biliary tree stem 2008;454, 428–35.
cells within peribiliary glands in primary sclerosing cholangitis. J Hepatol, 95. Scirpo, R., Fiorotto, R., Villani, A., Amenduni, M., Spirli, C., and
2015;63:1220–8. Strazzabosco, M. Stimulation of nuclear receptor peroxisome proliferator‐
68. Meng, F., Yamagiwa, Y., Ueno, Y., and Patel, T. Over‐expression of interleu- activated receptor‐γ limits NF‐κB‐dependent inflammation in mouse cystic
kin‐6 enhances cell survival and transformed cell growth in human malig- fibrosis biliary epithelium. Hepatol, 2015;62:1551–62.
nant cholangiocytes. J Hepatol, 2006;44:1055–65. 96. Fiorotto, R., Amenduni, M., Mariotti, V., Fabris, L., Spirli, C., and
69. Znoyko, I., Sohara, N., Spicer, S.S., Trojanowska, M., and Reuben, A. Strazzabosco, M. Src kinase inhibition reduces inflammatory and cytoskel-
Expression of oncostatin M and its receptors in normal and cirrhotic human etal changes in ΔF508 human cholangiocytes and improves cystic fibrosis
liver. J Hepatol,2005;43:893–900. transmembrane conductance regulator correctors efficacy. Hepatology,
70. Joplin, R., Hishida, T., Tsubouchi, H et  al. Human intrahepatic biliary 2018;67:972–88.
epithelial cells proliferate in vitro in response to human hepatocyte growth 97. Baeuerle, P.A. and Baltimore, D. NF‐kappa B: ten years after. Cell,
factor. J Clin Invest, 1992;90:1284–9. 1996;87:13–20.
71. Alvaro, D., Metalli, V.D., Alpini, G., et al. The intrahepatic biliary epithe- 98. Katt, J., Schwinge, D., Schoknecht, T. et  al. Increased T helper type 17
lium is a target of the growth hormone/insulin‐like growth factor 1 axis. J response to pathogen stimulation in patients with primary sclerosing chol-
Hepatol, 2005;43:875–83. angitis. Hepatology, 2013;58:1084–93.
72. Glaser, S.S., Gaudio, E., Miller, T., Alvaro, D., and Alpini, G. Cholangiocyte 99. Yang, Y., Liu, Y.J., Tang, S.T. et al. Elevated Th17 cells accompanied by
proliferation and liver fibrosis. Expert Rev Mol Med, 2009;11:e7. decreased regulatory T cells and cytokine environment in infants with bil-
73. Hall, C., Sato, K., Wu, N. et al. Regulators of cholangiocyte proliferation. iary atresia. Pediatr Surg Int, 2013;29:1249–60.
Gene Expr, 2017;17:155–71. 100. Li, J., Razumilava, N., Gores, G.J. et al. Biliary repair and carcinogenesis
74. Strazzabosco, M. and Somlo, S. Polycystic liver diseases: congenital disorders are mediated by IL‐33‐dependent cholangiocyte proliferation. J Clin Invest,
of cholangiocyte signaling. Gastroenterol, 2011;140:1855–9. 2014;124:3241–51.
75. Gatto, M., Drudi‐Metalli, V., Torrice, A. et al. Insulin‐like growth factor‐1 101. Gieseck, R.L. 3rd, Ramalingam, T.R., Hart, K.M. et al. Interleukin‐13 acti-
isoforms in rat hepatocytes and cholangiocytes and their involvement in pro- vates distinct cellular pathways leading to ductular reaction, steatosis, and
tection against cholestatic injury. Lab Invest, 2008;88:986–94. fibrosis. Immunity, 2016;45:145–58.
76. Spirli, C., Okolicsanyi, S., Fiorotto, R. et  al. ERK1/2‐dependent vascular 102. d’Adda di Fagagna, F. Living on a break: cellular senescence as a DNA‐
endothelial growth factor signaling sustains cyst growth in polycystin‐2 damage response. Nat Rev Cancer, 2008;8:512–22.
defective mice. Gastroenterol, 2010;138:360–71. 103. Muñoz‐Espín, D. and Serrano, M. Cellular senescence: from physiology to
77. Spirli, C., Okolicsanyi, S., Fiorotto, R. et al. Mammalian target of rapamycin pathology. Nat Rev Mol Cell Biol, 2014;15:482–96.
regulates vascular endothelial growth factor‐dependent liver cyst growth in 104. Strazzabosco, M., Fiorotto, R., Cadamuro, M. et  al. Pathophysiologic
polycystin‐2‐defective mice. Hepatology, 2010;51:1778–88. implications of innate immunity and autoinflammation in the biliary epithe-
78. Piccolo, S., Dupont, S., and Cordenonsi, M. The biology of YAP/TAZ: hippo lium. Biochim Biophys Acta, 2018;1864:1374–9.
signaling and beyond. Physiol Rev, 2014;94:1287–312. 105. Tsuneyama, K., Yasoshima, M., Harada, K., Hiramatsu, K., Gershwin,
79. Nagura, H., Smith, P.D., Nakane, P.K., and Brown, W.R. IGA in human bile M.E., and Nakanuma, Y. Increased CD1d expression on small bile duct
and liver. J Immunol, 1981;126:587–95. epithelium and epithelioid granuloma in livers in primary biliary cholangi-
80. Sugiura, H. and Nakanuma, Y. Secretory component and immunoglobulins tis. Hepatology, 1998;28:620–3.
in the intrahepatic biliary tree and peribiliary gland in normal livers and 106. Barnes, B.H., Tucker, R.M., Wehrmann, F., Mack, D.G., Ueno, Y., and
hepatolithiasis. Gastroenterol Jpn, 1989;24:308–14. Mack C.L. Cholangiocytes as immune modulators in rotavirus‐induced
81. Chuang, Y.H., Lan, R.Y., and Gershwin, M.E. The immunopathology murine biliary atresia. Liver Int, 2009;29:1253–61.
of human biliary cell epithelium. Semin Immunopathol, 2009;31: 107. Schrumpf, E., Tan, C., Karlsen, T.H. et al. The biliary epithelium presents anti-
323–31. gens to and activates natural killer T cells. Hepatology, 2015;62:1249–59.
82. Syal, G., Fausther, M., Dranoff J.A. Advances in cholangiocyte immunobiol- 108. Weiskirchen, R., Weiskirchen, S., and Tacke, F. Organ and tissue fibrosis:
ogy. Am J Physiol Gastrointest Liver Physiol, 2012;303:G1077–86. molecular signals, cellular mechanisms and translational implications. Mol
83. Wu, C.T., Davis, P.A., Luketic, V.A., and Gershwin, M.E. A review of the Aspects Med, 2018;65:2–15.
physiological and immunological functions of biliary epithelial cells: targets 109. Kisseleva, T. and Brenner, D.A. Mechanisms of fibrogenesis. Exp Biol Med
for primary biliary cholangitis, primary sclerosing cholangitis and drug‐ (Maywood), 2008;233:109–22.
induced ductopenias. Clin Dev Immunol, 2004;11:205–13. 110. Mehal, W. and Imaeda, A. Cell death and fibrogenesis. Semin Liver Dis,
84. Menendez, A. and Brett Finlay, B. Defensins in the immunology of bacterial 2010;30:226–31.
infections. Curr Opin Immunol, 2007;19:385–91. 111. Gressner, O.A. and Gao, C. Monitoring fibrogenic progression in the liver.
85. Chen, X.M., O’Hara, S.P., Nelson, J.B. et al. Multiple TLRs are expressed in Clin Chim Acta, 2014;10;433:111–22.
human cholangiocytes and mediate host epithelial defense responses to 112. Fabris, L., Spirli, C., Cadamuro, M., Fiorotto, R., and Strazzabosco, M.
Cryptosporidium parvum via activation of NF‐kappaB. J Immunol, Emerging concepts in biliary repair and fibrosis. Am J Physiol Gastrointest
2005;175:7447–56. Liver Physiol, 2017;313:G102–16.
32:  Cholangiocyte Biology and Pathobiology 407

113. Batzios, S.P., Zafeiriou, D.I., and Papakonstantinou, E. Extracellular matrix 121. Zeisberg, M., Yang, C., Martino, M. et al. Fibroblasts derive from hepato-
components: an intricate network of possible biomarkers for lysosomal cytes in liver fibrosis via epithelial to mesenchymal transition. J Biol Chem,
storage disorders? FEBS Lett, 2013;587:1258–67. 2007;282:23337–47.
114. Mallat, A. and Lotersztajn, S. Cellular mechanisms of tissue fibrosis. 5. Novel 122. Kisseleva, T. and Brenner, D.A. Is it the end of the line for the EMT?
insights into liver fibrosis. Am J Physiol Cell Physiol, 2013;305:C789–99. Hepatology, 2011;53:1433–5.
115. Klaas, M., Kangur, T., Viil. J. et al. The alterations in the extracellular matrix 123. Charrier, A., Chen, R., Kemper, S., and Brigstock, D.R. Regulation of pan-
composition guide the repair of damaged liver tissue. Sci Rep, 2016;6:27398. creatic inflammation by connective tissue growth factor (CTGF/CCN2).
116. Locatelli, L., Cadamuro, M., Spirlì, C. et al. Macrophage recruitment by Immunology, 2014;141:564–76.
fibrocystin‐defective biliary epithelial cells promotes portal fibrosis in con- 124. Pi, L., Robinson, P.M., Jorgensen, M. et al. Connective tissue growth factor
genital hepatic fibrosis. Hepatology, 2016;63:965–82. and integrin αvβ6: a new pair of regulators critical for ductular reaction and
117. Sato, Y., Harada, K., Ozaki, S. et  al. Cholangiocytes with mesenchymal biliary fibrosis in mice. Hepatology, 2015;61:678–91.
features contribute to progressive hepatic fibrosis of the polycystic kidney 125. Li, G., Xie, Q., Shi, Y. et al. Inhibition of connective tissue growth factor by
rat. Am J Pathol, 2007;171:1859–71. siRNA prevents liver fibrosis in rats. J Gene Med, 2006;8:889–900.
118. Yasoshima, M., Sato, Y., Furubo, S. et  al. Matrix proteins of basement 126. Kaffe, E., Fiorotto, R., Pellegrino, F. et al. β‐Catenin and interleukin‐1β‐
membrane of intrahepatic bile ducts are degraded in congenital hepatic dependent chemokine (C‐X‐C motif) ligand 10 production drives progres-
fibrosis and Caroli’s disease. J Pathol, 2009;217:442–51. sion of disease in a mouse model of congenital hepatic fibrosis.
119. Poole, L.G. and Arteel, G.E. Transitional remodeling of the hepatic extracellu- Hepatology, 2018;67:1903–19.
lar matrix in alcohol‐induced liver injury. Biomed Res Int, 2016;2016:3162670. 127. Strazzabosco, M., Sakisaka, S., Hayakawa, T., and Boyer, J.L. Effect of
120. Urribarri, A.D., Munoz‐Garrido, P., Perugorria, M.J. et  al. Inhibition of UDCA on intracellular and biliary pH in isolated rat hepatocyte couplets
metalloprotease hyperactivity in cystic cholangiocytes halts the develop- and perfused livers. Am J Physiol, 1991;260(1):G58–69.
ment of polycystic liver diseases. Gut, 2014;63:1658–67.
Polycystic Liver Diseases:
33 Genetics, Mechanisms,
and Therapies
Tatyana Masyuk, Anatoliy Masyuk, and Nicholas LaRusso
Division of Gastroenterology and Hepatology, Mayo Clinic College of Medicine, Rochester, MN, USA

INTRODUCTION therapeutic interventions in PLD are aimed at reducing liver


volume and improving quality of life. These procedures include
Polycystic liver disease (PLD), a genetic cholangiociliopathy, aspiration, sclerotherapy, cyst resection, fenestration, and liver
is characterized by the presence of multiple liver cysts of dif­ transplantation [4]. However, while these procedures might
ferent shape and size filled with cystic fluid [1]. PLD exists as relieve the symptoms and improve quality of life, they are inva­
an extra‐renal manifestation of autosomal dominant polycys­ sive, only partially effective, have high recurrence risk, and do
tic kidney disease (ADPKD) and autosomal recessive PKD not block disease progression.
(ARPKD) or as isolated autosomal dominant PLD (ADPLD). The only currently available pharmacological therapy in PLD
In PLD associated with ADPKD and ARPKD liver cysts is off‐label use of somatostatin analogs which reduce total liver
coexist with kidney cysts while in isolated ADPLD the majority volume and improve quality of life. While the beneficial value
of the cysts are localized to the liver with very few or no cysts of somatostatin analogs was shown in multiple clinical trials,
in the kidney. Clinically, PLD is arbitrarily defined by the their effects are modest, not all patients respond to treatment,
presence of more than 20 liver cysts as measured by ultra­ and the cost is very high [5, 6]. Clinical testing of UDCA did not
sonography, computed tomography, or magnetic resonance show reduction in the total liver volume in human subjects with
imaging [2]. Liver volume and number of cysts in patients isolated ADPLD, however, liver cyst volume was decreased in
with isolated ADPLD is generally larger than in patients with patients with ADPKD‐associated PLD [7]. Therefore, the search
ADPKD‐associated PLD [2]. for new therapies is necessary and ongoing.
Familial studies indicate that around 20% of patients with The discovery of PLD‐related genes and the identification of
mutations in PLD‐related genes have mild or no disease [2]. In mechanisms underlying cyst progression have provided a better
the remaining 80% of patients, progressive enlargement of liver understanding of the pathogenesis of this disease leading to new
volume results in PLD‐related symptoms caused by liver com­ perspectives for therapeutic interventions. The current knowl­
pression of adjacent organs. Based on the volume of the liver, edge related to the genetics of PLD, the mechanisms involved,
PLD is classified as mild (height‐corrected liver volume less and potential therapeutic targets identified are discussed in this
than 1600 mL), moderate (height‐corrected liver volume chapter.
between 1600–3200 mL), and severe disease (height‐corrected
liver volume more than 3200 mL) [3]. Severity of disease
­correlates with symptom burden and declined quality of life.
Patients with mild PLD seldomly develop complications. GENETICS
Patients with moderate and especially with severe disease
­experience pressure‐related symptoms such as pain, abdominal As of today, there are nine PLD‐causative genes that could be
distention and bloating, gastroesophageal reflux, nausea, and classified as: (i) exclusively mutated in ADPKD‐related PLD
dyspnea. Multiple complications might occur including cyst (i.e. PKD1 and PKD2), (ii) exclusively mutated in isolated
infection, hemorrhage, and rupture [2]. Currently available ADPLD (PRKCSH, SEC63, ALG8, LRP5, and SEC61B), and

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
33:  Polycystic Liver Diseases: Genetics, Mechanisms, and Therapies 409

Figure 33.1  PLD‐related genes and proteins. Hepatic cystogenesis in ADPKD‐associated PLD, ARPKD‐associated PLD, and in isolated ADPLD
is triggered by mutations in nine genes. PKD1 and PKD2 are causative for ADPKD‐associated PLD. Mutations in GANAB are associated with both
ADPKD‐associated PLD and isolated ADPLD. PKHD1 is implicated in development of both ADPKD and ADPLD. Five genes (i.e. PRKCSH, SEC63,
ALG8, LRP5, and SEC61B) are responsible for hepatic cystogenesis in ADPLD. Despite the genetic heterogeneity, in all forms of PLD multiple cysts
occupied liver parenchyma. Photograph of the resected liver from Pkhd1del2/del2 mice, light microscopic image of the liver section from the PCK rat and
scanning electron micrographs of the liver from Pkd2WS25/− mice demonstrate the presence of cysts in different animal models of PLD.

(iii) commonly mutated in PKD‐associated PLD and in isolated encodes the catalytic alpha subunit of glucosidase II (GΙΙα,
ADPLD (i.e. GANAB and PKHD1) [2, 4, 8, 9]. Despite the also known as PKD3) and is a one of the member of the gly­
genetic heterogeneity and complexity, phenotypically all PLDs cosyl hydrolase family of 31 proteins. GIIα resides in ER and
are characterized by the presence of numerous liver cysts of plays a role in protein folding and quality control (Figures 33.1
variable size and shape that occupy the majority of the liver and 33.2a) [4, 8, 9].
parenchyma (Figure 33.1).
ARPKD‐associated PLD
ADPKD‐associated PLD ARPKD (prevalence is 1 : 10 000–20 000) causes significant
The most common PLD is associated with ADPKD. The prev­ renal and liver‐related morbidity and mortality in children.
alence of ADPKD is 1 : 400 and up to 90% of ADPKD patients ARPKD is characterized by enlarged kidneys with various
develop PLD [2, 9, 10]. The kidney and liver cysts in this dis­ degrees of renal dysfunction, bile duct dilatation, and congenital
order are the result of mutations in three genes  –  PKD1, hepatic fibrosis that lead to portal hypertension [11]. The num­
PKD2, and GANAB. PKD1 and PKD2 genes encode, respec­ ber of affected individuals surviving the neonatal period that
tively, polycystin 1 (PC1) and polycystin 2 (PC2) proteins might require liver transplantation increases significantly with
(Figure  33.1). PC1 is expressed on the plasma membrane age. The disease is caused by mutation in a single PKHD1 gene
and along primary cilia (Figure  33.2a). PC2 is present in which encodes the protein fibrocystin (FC, Figure 33.1). FC is
endoplasmic reticulum (ER), plasma membrane, and primary expressed at plasma membrane and primary cilia (Figure 33.2a).
cilia (Figure 33.2a). The exact functions of PC1 and PC2 are While the functions of FC remained largely unclear, experimen­
not well understood. PC1 is known to interact with PC2 tal data suggest that this protein may act as a membrane‐bound
­regulating cell/cell–matrix interactions, cell proliferation, and receptor involved in regulation of the cell–cell adhesion and
G‐­protein‐coupled signal‐transduction pathways [4, 8, 9]. GANAB proliferation [11, 12].
410 THE LIVER:  GENETICS

Figure 33.2  Interactions of genes and proteins in PLD. (a) Schematic diagram illustrates the site of expression of PLD‐related proteins. PKD‐
associated PLD is known as a cholangiociliopathy indicating the nature of disease origin (i.e. cholangiocytes) and organelles involved (i.e.
­primary cilia). Indeed, the protein products of the genes mutated in PKD‐associated PLD are all localized to primary cilia. Proteins encoded by
the genes involved in isolated ADPLD are localized to ER but play an important role in targeting of PC1, PC2, and FC to the plasma membrane
and cilia. (b) Mutations in ADPLD‐related genes reduce functional dosage of PC1 resulting in hepatic cystogenesis. Abbreviations: FC, ­fibrocystin;
PC1, polycystin 1; PC, polycystin 2; LRP5, low‐density lipoprotein receptor‐related protein 5; GIIα, glucosidase II subunit alpha; GIIβ, glucosi­
dase II subunit beta; ALG8, alpha‐1,3‐glucosyltransferase; GANAB, glucosidase II alpha subunit; SEC63, protein‐transporting protein SEC63;
SEC61, translocon subunit SEC61; PRKCSH, protein kinase C substrate 80K‐H; PKHD1, polycystic kidney hepatic disease 1. Reproduced from [9]
with permission.

Isolated ADPLD grow. In line with this, a recent study in mice demonstrated that
deletion of the mitochondrial uncoupling protein 2 gene (i.e.
Isolated ADPLD (prevalence 1 : 100 000) is characterized by Ucp2) led to a spontaneous development of hepatic cysts that
the  predominant presence of liver cysts [5, 13]. ADPLD is a pathologically resembled human PLD [21].
result of heterozygous loss‐of‐function mutations in at least five
genes  –  PRKCSH, SEC63, GANAB, ALG8, and SEC61B
(Figure  33.1). There is some evidence that mutations in
Genetic interaction network in PLD
PRKCSH (responsible for ~15% of clinical cases) and SEC63 As mentioned, the genetic landscape of PLD includes nine
are associated with early onset of disease and more severe clini­ causative genes. Recent data clearly demonstrate a genetic
cal symptoms [2]. interaction network within PLD‐related genes [9, 22]. First,
The protein products of PRKCSH (the regulatory beta‐subu­ mutations in GANAB cause both ADPKD and isolated ADPLD
nit of glucosidase II, GIIβ), SEC63 (SEC63), GANAB (GIIα), while PKHD1 was found to be mutated in patients with
ALG8 (α‐1,3‐glucosyltransferase), and SEC61B (beta‐subunit ARPKD and ADPLD (Figure 33.1) [8, 16, 23]. Second, stud­
of the SEC61 protein translocation apparatus) are integral ER ies in animal models demonstrate that loss‐of‐function in
membrane residents (Figure 33.2a) that facilitate translocation PRKCSH, SEC63, GANAB, ALG8, and SEC61B genes com­
and transmembrane insertion of multiple glycoproteins bined with the functional dosage of PC1 act as a predictor of
­including PC1 and PC2 [8, 14–18]. In addition, LRP5 has been severity of hepatic and renal cystogenesis (Figure 33.2b) [8,
implicated in development of isolated ADPLD; however, no 16, 22, 23]. Third, GANAB is crucial for PC1 and PC2 matu­
clear loss‐of‐function variants in the LRP5 gene have been ration and localization to the cell surface and cilia [16].
found in ADPLD patients [19]. The protein product of LRP5 Fourth, GΙΙβ and SEC63 are required for formation of PC1/
(i.e. LRP5) is located to the cell membrane where it interacts PC2 functional complex with PC1 being the rate‐limiting
with Frizzled in the canonical Wnt signaling pathway component of this process [23]. Fifth, mutations in ALG8,
(Figures 33.1 and 33.2a). Upon binding of Wnt to the Frizzled/ GANAB, and SEC61B reduce steady‐state levels of PC1;
LRP5 complex, the canonical Wnt/beta‐catenin pathway is among them, SEC61B causes the most notable quantitative
activated subsequently enhancing cholangiocyte proliferation. reduction of PC1 (Figure 33.2b) [8]. Sixth, mutations in the
Finally, mutations in the ARPKD‐associated gene, PKHD1, PKHD1 gene likely contribute to hepatic cystogenesis through
was shown as disease‐causative not only in ARPKD but in aberrant interaction of PC1 with FC without affecting biogen­
ADPLD as well (Figure 33.1) [20]. esis of PC1 [8]. Finally, missense variants in LRP5 coexist
Seven genes associated with isolated ADPLD account for with PKHD1 variants suggesting that interaction of these two
approximately 50% of genetically identifiable cases; therefore genes contribute to hepatic cystogenesis in patients with
the list of genes causative for this form of PLD is expected to mutated PKHD1 [8].
33:  Polycystic Liver Diseases: Genetics, Mechanisms, and Therapies 411

MECHANISMS OF HEPATIC Morphological abnormalities in primary cilia, basal bodies,


CYSTOGENESIS and/or centrosomes coupled with disturbances in ciliary func­
tions accelerate cholangiocyte proliferation and cyst expan­
Cholangiocytes lining liver cysts are enlarged with thickened sion [5, 9, 27, 31]. Peculiarities in primary cilia and centrosomes
basement membrane [24–26]. While the majority of hepatic are also accompanied by dysregulated cell cycle machinery
cysts are lined by single layer of cholangiocytes, around 10% of and anomalous expression of the cell cycle‐related proteins
cysts in animal models of PLD (i.e. the PCK rat and [26, 33, 34].
Pkd2WS25/−mice) and around 20% of cysts in ADPKD patients
are multilayered [25].
Intracellular cAMP and [Ca2+]i signaling
pathways
Ductal plate malformation
Elevated levels of cAMP in PLD cholangiocytes is considered
Ductal plate malformation (i.e. the embryological arrest of to be one of the major mechanisms that drives hepatic cystogen­
ductal plate development) contributes to hepatic cystogenesis in esis [5, 14, 31]. This second messenger was shown to regulate
PLD [5, 27]. Incomplete development of ductal plates leads to many cellular functions including autophagy, the cell cycle,
the presence of their remnants along the portal tract margins in fluid secretion, and cell proliferation via its downstream effec­
ARPKD‐associated PLD [28]. ADPKD‐associated PLD and tors Rap guanine nucleotide exchange factor 3 also known as
isolated ADPLD are also linked to ductal plate malformation; exchange factor directly activated by cAMP (EPAC), protein
however, the pathological mechanisms involved are unknown kinase A (PKA), and mitogen‐activated protein kinase/extracel­
[5, 27]. In addition, many regulators of ductal plate remodeling lular signal‐regulated kinase (MAPK/ERK) [29, 35–38]. The
(such as growth factors, transcriptional factors, and miRNAs) mechanisms behind cAMP elevations in PLD cholangiocytes
are abnormally expressed in cystic cholangiocytes, therefore are still not well understood. However, recent data suggest that
further connecting ductal plate malformation and hepatic cys­ the overexpression of cAMP‐linked bile acid‐responsive
togenesis [5, 27]. G‐­protein coupled receptor, TGR5, combined with previously
published observations that the concentration of endogenous
TGR5 ligands is increased in cystic cholangiocytes, might con­
Cholangiocyte primary cilia tribute to cAMP overproduction in PLD [38, 39].
PKD‐associated PLD is often referred as a cholangiociliopathy PLD cholangiocytes are also characterized by the reduced
[29, 30] accentuating the cell of origin and affected organelle in concentration of intracellular calcium ([Ca2+]i) believed to result
hepatic cystogenesis. In control cholangiocytes primary cilia from a dysfunctional PC1/PC2 ciliary complex [32]. Mutational
(around 7 μm in length) extend from the apical membrane into defects in PC2 reduce entry of extracellular calcium that leads
the bile duct lumen and function as chemo‐, osmo‐ and mecha­ to depletion of cytoplasmic and ER calcium stores altering
nosensors detecting and transducing extracellular stimuli into the homeostasis of this signaling molecule [27, 32, 40]. Decreased
cell interior to regulate cholangiocyte proliferation, planar cell [Ca2+]i may contribute to cAMP elevations by activating c­ alcium
polarity, intracellular cyclic adenosine monophosphate (cAMP) inhibitable adenylyl cyclase 5/6 (AC5/6) and by preventing
and calcium signaling and many other cellular functions [29, 30]. cAMP degradation by phosphodiesterase 1 (PDE1) and PDE3
The protein products of genes that cause ADPKD‐ and [41]. Alternatively, due to existing crosstalk between these two
ARPKD‐linked PLD (i.e. PKD1, PKD2, and PKHD1) are all signaling pathways, cAMP can influence calcium signaling via
expressed in primary cilia and play an important role in preser­ PKA activation [10, 27].
vation of ciliary structure and functional integrity (Figure 33.2a)
[31]. Indeed, mutations in PKD1, PKD2, and PKHD1 are asso­
ciated with various ciliary abnormalities including aberrantly
Cholangiocyte autophagy
shaped, shortened, and malformed cholangiocyte cilia in vivo Recently we discovered that autophagy, an evolutionary con­
and in vitro (Figure 33.3a,b) [12, 24, 25]. In addition, basal bodies served pathway involved in regulation of cellular homeostasis
(i.e. the nucleation site for the growth of primary cilia) and cen­ via degradation and recycling of cytoplasmic constituents, is
trosomes are also defective in PLD as evident by their aberrant abnormal in PLD [37]. The morphologic appearance of orga­
positioning and elongated shape [26]. The presence of supernu­ nelles involved in autophagy is strikingly different from those in
merary centrosomes in cholangiocytes within hepatic cysts control cholangiocytes; for example, the number of autophago­
results in the formation of extra cilia in up to 15% of the cells somes and their size was around 2–3 times greater in cultured
[26]. Malfunctioned planar cell polarity (i.e. improper align­ PCK rat and human ADPKD cholangiocytes (Figure 33.3c) as
ment of the mitotic spindle during cell division which leads to well as in cholangiocytes lining liver cysts in animal models of
longitudinal rather than horizontal cell growth) has also been PLD and human patients with PLD compared to respective con­
implicated [32]. On the other hand, the protein products of trols [37]. Changes in autophagosome morphology was accom­
genes which are causative for isolated ADPLD are not present in panied by similar changes (i.e. increased number and size) in
cilia (Figure  33.2a); however, several studies clearly demon­ lysosomes and autolysosomes, the organelles critical for proper
strated that they are crucial for biogenesis of proteins associated functioning of autophagic machinery [37]. Global analysis of
with PKD‐related PLD (i.e. PC1, PC2, and FC) and their proper the PLD cholangiocyte transcriptome followed by functional
translocation to the cholangiocyte primary cilia [8, 16, 23]. pathway cluster analysis showed that the autophagy–lysosomal
Figure 33.3  PLD cholangiocytes are characterized by a number of morphological abnormalities. (a) In cholangiocytes lining liver cysts in an
animal model of PLD, the PCK rat, primary cilia are shorter and malformed as assessed by scanning electron microscopy (SEM), transmission
electron microscopy (TEM), and immunofluorescent confocal microscopy (IMF) compared to cholangiocytes lining bile ducts in wild‐type rat.
(b) (Consistent with in vivo data, shorter and defective cilia are also present in cultured cholangiocytes derived from the PCK rat as compared to
cholangiocytes derived from wild‐type rat (NRC). Reproduced from [101] with permission. (c) Transmission electron microscopic images demon­
strate that the number and size of autophagosomes (depicted by black arrows) are greater in cystic cholangiocytes derived from the PCK rat and
human patients with ADPKD compared to control cholangiocytes derived, respectively, from the wild‐type rat (NRC) and healthy humans (NHC).
33:  Polycystic Liver Diseases: Genetics, Mechanisms, and Therapies 413

pathway is one of the most altered pathways in PLD. Cystic might contribute to enhanced fluid secretion in PLD. However,
cholangiocytes have aberrantly expressed autophagy‐related no such experimental evidence exists so far.
proteins and enhanced autophagic flux. Furthermore, altered
autophagy in PLD cholangiocytes is accompanied by activation
of the cAMP–PKA–cAMP response element binding protein
miRNAs
(CREB) signaling pathway and enhanced cell proliferation [37]. As of today a single experimental study described the role of
It has been also reported that mutations in PRKCSH induces miRNAs in PLD progression [34]. It was shown that miR‐15a is
autophagy although a different mechanism (i.e. mTOR‐­ one of the most downregulated miRNA in rat cystic cholangio­
mediated) was implicated [42]. In line with our observations, cytes. This miRNA targets the cell cycle master phosphatase,
the absence of endogenous hepatocystin (i.e. the protein product cell division cycle 25 (Cdc25A), the expression of which is
of PRKCSH gene) or the presence of mutated hepatocystin in markedly increased in PLD. Experimental upregulation of miR‐
an  in  vitro experimental system induces autophagy [42]. 15a in cystic cholangiocytes decreases levels of Cdc25A, inhib­
Together  these data emphasize the important contribution of its cell proliferation, and reduces cyst growth in vitro. In
autophagy in hepatic cystogenesis and suggest autophagy as a contrast, downregulation of miR‐15a in control cholangiocytes
therapeutic target. increases expression of Cdc25A, enhances cell proliferation,
and accelerates cyst growth in culture [34]. Therefore, these
data indicate that alterations in miR‐15a/Cdc25A axis contrib­
Cholangiocyte proliferation
ute to the molecular pathogenesis of PLD.
Cholangiocyte proliferation is considered to be one of the We also found that besides miR‐15a, expression of additional
important mechanisms of hepatic cystogenesis and in many pre­ 67 miRNAs was dysregulated in PLD cholangiocytes [56].
clinical studies was shown to be a marker of disease progression Importantly, the mRNA targets of these miRNAs include those
or regression [24, 29, 33–35, 39, 43–51]. Under basal con­ related to cycle machinery, fluid secretion, cAMP pathway, and
ditions, cholangiocytes are dormant but become hyperprolifera­ calcium intracellular signaling further indicating that miRNAs
tive during cyst expansion as indicated by their positive staining are specific regulators of affected intracellular pathways in PLD.
for proliferating cell nuclear antigen (PCNA) [48–50, 52].
However in well‐established disease, proliferation of cystic
cholangiocytes diminishes [27, 53]. Different cytokines and
Other mechanisms of hepatic cystogenesis
growth factors (such as interleukin 6 [IL‐6], epidermal growth In addition to the above mentioned mechanisms implicated in PLD
factor [EGF], vascular epidermal growth factor [VEGF], and pathogenesis, many more have been experimentally proven to pro­
insulin‐like growth factor [IGF]) accelerate cholangiocyte pro­ mote hepatic cystogenesis (Figure  33.4a). These mechanisms
liferation via ERK and phosphatidylinositol‐4,5‐bisphosphate include, in particular: (i) enhanced angiogenesis and cyst wall neo­
3‐kinase (PI3‐kinase)–protein kinase B (AKT)–mTOR path­ vascularization, (ii) extracellular matrix remodeling, and (iii) global
ways. These cytokines and growth factors are present in the changes in mRNA and protein expression [10, 32, 49, 57].
cystic fluid and/or secreted by cystic cholangiocytes [27, 32]. In We recently performed mRNA profiling analysis of human
PLD linked to ARPKD, a proliferation‐independent mechanism cholangiocytes derived from healthy individuals and from
(i.e. excessive differentiation of biliary precursors during ductal patients with ADPKD and ADPLD using the NexGen sequenc­
plate remodeling) appears to play a role in early stages of cyst ing (NGS) technology. The expression patterns of mRNAs dif­
development followed by proliferation‐dependent cyst expan­ fer strikingly in control and cystic cholangiocytes. Many
sion in the postnatal period [54]. transcripts are up‐ or downregulated in PLD. Importantly,
ADPKD and ADPLD cholangiocytes share a number of com­
monly expressed genes (Figure  33.4b). Clustering of these
Fluid secretion
genes by functional pathway cluster analysis reveals previously
Increased secretion of fluid into the lumen of hepatic cysts is identified cellular pathways and exposes some novel pathways
also an important contributing mechanism in PLD progression. as well (Figure 33.4c).
In cystic cholangiocytes of animal models and humans with
PLD, cystic fibrosis transmembrane conductance regulator
(CFTR; i.e. the chloride channel), anion exchange protein 2
(AE2; i.e. the chloride/bicarbonate exchanger), and aquaporin 1 THERAPIES
(AQP1; i.e. the water channel) are all overexpressed and mislo­
calized. Activation of cAMP machinery in cultured cholangio­ Accumulating evidence suggest that mutations in PLD‐causa­
cytes expedites insertion of these proteins into the apical tive genes initiate the formation of hepatic cysts that continue to
membrane and thus facilitates water influx [43]. In line with this grow because of disturbances in cellular functions and signaling
observation, an overexpression of AQP1 was recently reported pathways. The efficacy of targeting of these processes for poten­
in PLD patients [55]. tial therapeutic applications has been tested in cultured cystic
The concentration of bile acids is increased in the liver of cholangiocytes and in animal models of PLD [2, 4, 9, 10, 27].
PCK rats and in cystic fluid of PLD patients [39]. Given the Several of these preclinical studies have been translated into
importance of these molecular species in bile flow regulation clinical trials. Specifically, the role of the somatostatin analogs,
and the overexpression of bile acid‐responsive receptor TGR5 octreotide and pasireotide, and the effects of UDCA were exam­
in cystic cholangiocytes, it has been suggested that bile acids ined in patients with ADPKD‐associated PLD and isolated
414 THE LIVER:  SINGLE DRUG THERAPIES

Figure 33.4  Cholangiocyte morphological defects and abnormal signaling pathways underlie hepatic pathology in PLD. (a) Cystic cholangio­
cytes are characterized by structurally deformed primary cilia, increased number of autophagic organelles and mitochondria, and numerous dys­
regulated cellular functions. (b) Using the NextGen sequencing approach, 2310 and 1914 dysregulated transcripts were detected, respectively, in
cholangiocytes derived from human patients with ADPKD‐associated PLD and isolated ADPLD. These genes were differentially expressed (i.e.
up‐ and downregulated) in cystic cholangiocytes compared to cholangiocytes derived from healthy individuals (fold changed great than 2.0, P
< 0.05). Using a Venn diagram analysis, 1078 transcripts were commonly present in cystic cholangiocytes. (c) The pathway cluster analysis of these
1078 genes identified pathways associated with PLD pathogenesis.

ADPLD [2, 4–7, 58]. In clinical trials, total liver volume is a SINGLE DRUG THERAPIES
prognostic marker of disease and the primary endpoint for
assessing drug effects in patients. In preclinical trials, liver
cystic areas and the rate of cholangiocyte proliferation are usu­
Somatostatin analogs
ally used to assess disease progression/regression. The majority Synthetic analogs of somatostatin currently represent the main
of preclinical studies have focused on the evaluation of a single pharmacological treatment options for PLD patients and are
drug; however, studies that examine combinational strategies prescribed off‐label [5, 27, 58]. Out of five somatostatin recep­
have also begun to emerge. Here we discuss current promising tors (SSTR) known to be present in cholangiocytes octreotide
monotherapies and combinational therapies in PLD treatment binds to SSTR2, SSTR3, and SSTR5. Another somatostatin
which are summarized in Table 33.1. analog, pasireotide, is more potent than octreotide, has a broader
33:  Polycystic Liver Diseases: Genetics, Mechanisms, and Therapies 415

Table 33.1  Therapeutic targets in polycystic liver disease (PLD)


Target Studies Drug Outcome Ref
SINGLE DRUG THERAPIES
cAMP Preclinical Octreotide, pasireotide Decreased cholangiocyte proliferation in vitro and hepatic [33, 47]
cystogenesis in PCK rats and Pkd2WS25/− mice
Clinical Octreotide, lanreotide Decreased total liver volume and increased quality of life [59, 62–69,
in patients with isolated ADPLD and ADPKD‐associated 71, 73, 74]
PLD
Pasireotide A clinical trial involving patients with isolated ADPLD
and ADPKD‐associated PLD is underway at the Mayo
Clinic (NCT01670110107)
Pasireotide No improvement in cyst reduction after aspiration [75]
sclerotherapy
Intracellular Preclinical UDCA Decreased cholangiocyte proliferation in vitro and hepatic [39]
calcium cystogenesis in PCK rats
Clinical UDCA No changes in total liver volume; decreased liver cystic [7]
volume in patients with ADPKD‐associated PLD (but not
in ADPLD patients)
Tendency to decrease the total liver volume in patients [80]
with PLD
Autophagy Preclinical HCQ Decreased cholangiocyte proliferation in vitro and hepatic [37]
cystogenesis in PCK rats
TGR5 Preclinical SBI‐115 Decreased cholangiocyte proliferation and cysts growth [38]
in vitro
AC5 Preclinical SQ22536 Decreased cholangiocyte proliferation in vitro and hepatic [83]
cystogenesis in PC2‐defective mice
PKA, MEK Preclinical PKA & MEK inhibitors Decreased cholangiocyte proliferation and cysts growth [29, 84]
in vitro
FSHR Preclinical Reduction by shRNA Decreased cholangiocyte proliferation in vitro [85, 100]
VR2 Preclinical Tolvaptan, OPC‐31260 Decreased cholangiocyte proliferation in vitro [86]
Single case Tolvaptan Decreased total liver volume in a single patient with [88]
report severe PLD
TRPV4 Preclinical GSK1016790A Decreased cholangiocyte proliferation in vitro and hepatic [98]
cystogenesis in PCK rats
BNP Preclinical BNP Decreased cholangiocyte proliferation in vitro and hepatic [89]
cystogenesis in PCK rats
HSP90 Preclinical STA‐2842 Decreased cholangiocyte proliferation in vitro and hepatic [90]
cystogenesis in Pkd1KO mice
CDC25A Preclinical Menadione, PM‐20 Decreased cholangiocyte proliferation in vitro and hepatic [48]
cystogenesis in PCK rats and Pkd2WS25/− mice
VEGF Preclinical SU5416 Decreased cholangiocyte proliferation in vitro and hepatic [50, 92]
cystogenesis in PC2‐deficient mice
PPARγ Preclinical Pioglitazone, telmisartan Decreased hepatic cystogenesis in PCK rats [93, 94]
HDAC6 Preclinical Tubastatin‐A, tubacin, Decreased cholangiocyte proliferation in vitro and hepatic [35, 44]
panobinostat, ACY‐1215, cystogenesis in PCK rats
ACY‐738, ACY‐241
MMP Preclinical Marimastat Decreased cholangiocyte proliferation in vitro and hepatic [95]
cystogenesis in PCK rats
mTOR Preclinical Rapamycin Decreased cholangiocyte proliferation in vitro and hepatic [49]
cystogenesis in Pkd2KO mice
Sirolimus No effects on hepatic cystogenesis in PCK rats [96]
Everolimus Decreased hepatic cystogenesis in PCK rats [97]
c‐Src Preclinical SKI‐606 Decreased hepatic cystogenesis in PCK rats [51]
COMBINATIONAL DRUG THERAPIES
cAMP Clinical Everolimus No synergistic effects on total liver volume [98]
+ +
mTOR octreotide
cAMP Preclinical Octreotide Octreotide abolished the paradoxical effects of sorafenib [84]
+ + on hepatic cystogenesis in PC2‐deficient mice
RAF sorafenib
cAMP Preclinical Octreotide No additive effects on hepatic cystogenesis in PCK rats [99]
+
pasireotide
cAMP Preclinical Pasireotide Additive effects on cholangiocyte proliferation and cyst [38]
+ growth in vitro
SBI‐115
cAMP Preclinical Pasireotide Additive effects on cholangiocyte proliferation in vitro and [37]
+ + hepatic cystogenesis in PCK rats
autophagy HCQ
cAMP Preclinical Pasireotide Additive effects on cholangiocyte proliferation in vitro and [35]
+ + hepatic cystogenesis in PCK rats
HDAC6 ACY‐1215
416 THE LIVER:  SINGLE DRUG THERAPIES

spectrum of binding (i.e. binds to SSTR1, SSTR2, SSTR3, and UDCA as therapeutic agent in PLD has been tested in preclini­
SSTR5) and has around six times the half‐life [33, 58]. In cal and clinical studies.
­clinical trials, three somatostatin analogues (namely, octreotide,
lanreotide, and pasireotide) have been used. A dose of 120 mg Preclinical studies
of lanreotide is equivalent to 60 mg of octreotide [59].
In PCK rats (that have increased cAMP, decreased intracellular
calcium, and elevated concentration of toxic bile acids), UDCA
Preclinical studies
inhibits hepatic cystogenesis, fibrosis, and cholangiocyte
In cultured rat and human cystic cholangiocytes, both octreotide hyperproliferation and improves their physical fitness. UDCA
and pasireotide decreased cAMP levels and cell proliferation with also normalized bile acid concentrations in bile. These effects
pasireotide being more effective [33, 47]. Consistent with this, a overlap with restoration of the intracellular calcium in cystic
more substantial reduction of hepatic cystogenesis was observed cholangiocytes via a PI3K–AKT–MEK/ERK‐dependent mech­
in animal models of PLD (i.e. PCK rats and Pkd2WS25/− mice) after anism [39].
treatment with pasireotide than with octreotide [33, 47]. The
reduced cyst growth was coupled with decreased cAMP levels Clinical trials
and rate of cholangiocyte proliferation. These studies provided a
Given the positive outcome of preclinical studies, a clinical trial
strong justification for assessing the potential benefits of somato­
was performed to assess the effects of UDCA in patients with
statin analogs in humans.
ADPKD‐associated PLD and isolated ADPLD. UDCA was
well tolerated with no major side‐effects. While no changes
Clinical trials
in  total liver volume were observed in response to UDCA
A number of clinical trials have evaluated the efficacy of soma­ ­treatment, liver cystic volume was decreased in patients with
tostatin analogs in patients with isolated ADPLD and ADPKD‐ ADPKD‐associated PLD (but not in ADPLD patients). The
associated PLD. The overall summary of the findings is as levels of the γ‐glutamyl transpeptidase (GGT) and alkaline
­
follows. First, octreotide and lanreotide modestly (by around phosphatase (ALP) were reduced in all patients [7]. In another
5%) decreased total liver volume when administrated for 6–12 small trial, UDCA treatment also reduced the levels of biliary
months; this decrease remained stable for 12 months post‐ther­ enzymes (i.e. GGT and ALP) and had a tendency (although not
apy. Second, with continuous octreotide usage, the reduction is statistically significant) to decrease the total liver volume [80].
liver volumes is maintained; however, when treatment is Thus, UDCA might be a pharmacological option for patients
stopped, liver volume begins to increase toward baseline. Third, with ADPKD‐associated PLD [5, 7, 57, 80].
patients with the most severe PLD appear to respond to treat­
ment more efficiently than those with modest disease. Fourth, in
some patients (around 15%) treatment has no effects on total
Autophagy
liver volume and quality of life. Fifth, somatostatin analogues Recent data strongly suggest that altered autophagy in PLD
are well tolerated and improve quality of life. Unfortunately, cholangiocytes contributes to disease progression and might
somatostatin analogs are very expensive, US$7000  –  US$11 serve as a potential therapeutic target [37, 42]. Indeed, treatment
000 per month for patients with health insurance who have to of cultured rat and human cholangiocytes with inhibitors of
obtain pre‐approval usually on a case‐by‐case basis [6, 59–74]. autophagy, bafilomycin A1 and hydroxychloroquine (HCQ),
A clinical trial (NCT01670110107) to examine the efficacy decreased their proliferation and cyst growth in vitro. However,
of a more potent somatostatin analog pasireotide in ADPKD‐ in PLD cholangiocytes with genetically reduced expression of
associated PLD and isolated ADPLD is now underway at the one of the major autophagy‐related proteins, ATG7, these effects
Mayo Clinic College of Medicine, USA. In line with this, in of bafilomycin A1 and HCQ were abolished. Finally, in PCK
PLD patients who underwent aspiration sclerotherapy of a large, rats HCQ treatment attenuated hepatic cystogenesis by decreas­
single, symptomatic, hepatic cyst, pasireotide did not improve ing cAMP levels and proliferation of cholangiocytes lining liver
the reduction in cyst size. Inadequate dose of pasireotide or cysts [37].
­limited timing of drug administration (administered two weeks
before and two weeks after procedure) might explain the nega­
tive result [75].
cAMP‐linked bile acid receptor TGR5
TGR5 is expressed in different cell types regulating multiple
cellular processes and recently has been considered as a poten­
Ursodeoxycholic acid tial therapeutic target in metabolic, inflammatory, and digestive
Bile acids have been recognized as signaling molecules regulat­ diseases [38]. Under basal conditions TGR5 is localized to dif­
ing cell proliferation, apoptosis, and liver regeneration [76]. ferent cholangiocyte compartments including primary cilia and
Ursodeoxycholic acid (UDCA; 3α,7β‐dihydroxy‐5β‐cholanoic apical membrane [36, 81]. In PLD cholangiocytes, TGR5 is
acid), an endogenous hydrophilic bile acid, stimulates a bicar­ overexpressed and the concentrations of endogenous TGR5
bonate‐rich choleresis protecting both hepatocytes and cholan­ agonists (e.g. lithocholic, chenodeoxycholic, deoxycholic, and
giocytes against the cytotoxicity of hydrophobic bile acids by cholic acids) are increased in livers and serum of PCK rats
increasing levels of intracellular calcium [77–79]. Taking into [38, 39]. Activation of cultured rat and human cystic cholangio­
consideration that cystic cholangiocytes are characterized by cytes with TGR5 natural or synthetic agonists further increased
decreased levels of intracellular calcium a potential role of cAMP levels, cell proliferation, and cyst growth. Furthermore,
33:  Polycystic Liver Diseases: Genetics, Mechanisms, and Therapies 417

the TGR5 agonist, xenobiotic oleanolic acid, worsened hepatic its agonist, vasopressin, increased cell proliferation and cAMP
cystogenesis in PCK rats and increased proliferation of cystic levels which were blocked by preincubation of cholangiocytes
cholangiocytes in vivo. Genetic reduction of TGR5 in vitro or with two VR2 antagonists, tolvaptan and OPC‐31260 [86]. As
complete elimination of this G protein‐coupled receptor (GPCR) of today, effects of VR2 targeting on hepatic cystogenesis have
in vivo by generating TGR5−/−; Pkhd1del2/del2 double mutant mice not been tested preclinically.
prevented the effects of TGR5 activation of cyst growth and
reduced cAMP production. Together, these data indicate that Single clinical case
TGR5 contributes to cAMP elevation and to hepatic cystogene­
An inhibitor of VR2, tolvaptan, was shown to decrease kidney
sis in PLD representing a potential therapeutic target. Indeed, a
volume in patients with PKD in clinical trials and has recently
novel small molecule TGR5 antagonist, SBI‐115, decreases
been approved by the Food and Drug Administration (FDA) for
cAMP levels, rate of proliferation, and cyst growth in cultured
treatment of renal (but not hepatic) cystogenesis in ADPKD
cholangiocytes emphasizing the beneficial potential of TGR5
patients [87]. A single clinical report also demonstrates that
inhibition in PLD [38].
tolvaptan decreased total liver volume after 17 months of treat­
ment in a patient with severe PLD [88]. As of today, no other
Adenylyl cyclase 5 (AC5) data related to efficacy of tolvaptan in PLD patients have been
reported.
Several studies implicated AC5/6 in the pathogenesis of PLD.
Cholangiocyte primary cilia possess a protein complex consist­
ing of AC5/6 and the A kinase anchor protein AKAP150. TRPV4
Activation of this complex by luminal flow triggers changes in The calcium entry channel transient receptor potential vanilloid 4
intracellular calcium and cAMP levels [82]. In PC2‐defective (TRPV4) is overexpressed at the mRNA and protein levels in
cells, inhibition of AC5 reduced cAMP production, expression PLD cholangiocytes. Activation of TRPV4 in these cells increased
of cAMP downstream effector ERK, VEGF secretion, and levels of intracellular calcium subsequently inhibiting cell prolif­
growth of biliary organoids. Finally, an AC5 inhibitor, SQ22536, eration and cyst growth in vitro by activating AKT and inhibiting
decreased liver cystic area and cell proliferation in PC2‐­ the BRAF–ERK signaling pathway. Partial silencing of TRPV4
defective mice [83]. by small interfering RNA blocked effects of TRPV4 activation on
these processes. Finally, a TRPV4 activator, GSK1016790A,
cAMP downstream effectors decreased hepatic cystogenesis in PCK rats [44].
cAMP machinery has been shown to pay an important role in
PLD pathogenesis. One of the PKA regulatory subunits, PKA‐ BNP B‐type natriuretic peptide (BNP)
RIb, is overexpressed in cystic cholangiocytes and its activation
BNP, a guanylyl cyclase A agonist, triggers cyclic guanosine
triggers cell proliferation and cyst growth in vitro [29]. These
monophosphate (cGMP) production demonstrating antifibrotic,
PKA‐stimulated effects were blocked by the PKA‐ and MEK‐spe­
antihypertensive, and vasopressin‐suppressive properties. BNP
cific inhibitors decreasing ERK phosphorylation, rate of prolifera­
decreased renal cystic epithelial cell proliferation, suppressed
tion, and cyst growth in 3D cultures [29, 84]. Given the recent
fibrotic gene expression, and increased intracellular calcium in
recognition that PKA targeting might be beneficial in other prolif­
vitro. Continuous BNP production in PCK rats (in response to
erative diseases such as cancer, it would be of value to test the
adeno associated viral serotype 9 [AAV9] vectors that carry
effects of PKA inhibitors on hepatic cystogenesis in vivo.
cytomegalovirus [CMV]‐driven codon‐optimized proBN­
PcDNA) inhibited hepatic cystogenesis. Thus, targeting cGMP
Follicle‐stimulating hormone receptor (FSHR) in PLD may provide a novel therapeutic strategy [89].
FSHR is a cAMP‐associated receptor involved in follicular
growth. FSHR is present in cystic cholangiocytes derived from HSP90
livers of patients with ADPKD‐associated PLD and its stimula­
Heat shock protein 90 (HSP90) is a chaperone protein that con­
tion increases cell proliferation in a cAMP–ERK‐dependent
trols proper folding of its clients preventing their degradation. In
manner; these effects were abolished after genetic reduction of
conditional Pkd1 KO mice, inhibitor of HSP90 STA‐2842 (a
FSHR. Nevertheless, therapeutic efficacy of FSHR has not yet
resorcinolic triazole molecule) prevented growth of hepatic
been tested in animal models of PLD [85].
cysts and decreased cholangiocyte proliferation. In addition,
expression and activity of HSP90 client proteins, EGFR and
Vasopressin ERK1/2, was reduced [90].

Preclinical studies
Recently, vasopressin receptor 2 (VR2; one of the three known
CDC25A
arginine vasopressin receptors – VR1A, VR1B, and VR2) was In PLD cholangiocytes overexpression of the master cell cycle
found to be present in control cholangiocytes [86]. Moreover, regulator Cdc25A is associated with cell cycle dysregulation
overexpressed VR2 was detected in cystic cholangiocytes of and increased numbers of cells with multiple centrosomal
ADPKD patients. Activation of this cAMP‐linked receptor with abnormalities [26, 34]. Genetic depletion of Cdc25A from
418 THE LIVER:  COMBINATIONAL DRUG THERAPIES

cultured cholangiocytes or complete elimination of this protein mTOR


by generating Cdc25A+/−;Pkhd1del2/del2 double mutant mice nor­
malized these processes [26, 48]. In addition, suppression of The mammalian target of rapamycin (mTOR) has been impli­
Cdc25A by increasing levels of its regulator, miR‐15A, inhib­ cated in the regulation of a variety of fundamental cellular pro­
ited cholangiocyte proliferation and cyst growth in vitro [34]. cesses including cell growth, metabolism, protein synthesis and
These observations provided the rationale to study the therapeu­ autophagy. In PLD, a dysregulated mTOR signaling network
tic potential of Cdc25A targeting in PCK rats and Pkd2WS25/− contributes to disease progression by affecting a spectrum of
mice. Two Cdc25A inhibitors (i.e. menadione and PM‐20) intracellular pathways, including growth factor signaling [49].
decreased cyst growth in these models of PLD by reducing Overexpression of mTOR in cystic cholangiocytes is accompa­
activity and expression of Cdc25A [48]. nied by upregulation of VEGF, IGF1, IGF1R, and PKA.
Rapamycin, an inhibitor of mTOR, lessened cyst growth in
Pkd2 KO mice by decreasing cholangiocyte proliferation. The
VEGF effects of rapamycin were coupled with accumulation of HIF1α
Vascular endothelial growth factor (VEGF)/VEGF receptor 2 (the major transcription factor of VEGF) and inhibition of
(VEGFR2) plays an important role in hepatic cystogenesis. VEGF secretion with involvement of the MEK–ERK pathway
Decreased levels of intracellular calcium and increased [49]. On the other hand, the mTOR inhibitor, sirolimus, had no
­production of cAMP in PLD cholangiocytes stimulate mTOR/ effects on hepatic and renal cystogenesis in PCK rats [96].
hypoxia‐inducible factor 1‐alpha (HIFα) axis via the PKA– Finally, an alternative mTOR inhibitor, everolimus, halted
RAS–RAF–ERK pathway leading to overproduction of VEGF hepatic cystogenesis and reduced hepatic fibrosis in PCK rats
[49, 50]. The VEGFR1 and VEGFR2 are also upregulated in via PI3K–AKT–mTOR signaling [97]. In this study, the conse­
liver cysts and their activation enhanced cholangiocyte prolifera­ quences of everolimus treatment on disease progression were
tion in a cAMP‐dependent fashion [91]. In contrast, an inhibitor evaluated by magnetic resonance imaging approach using liver
of VEGFR2, SU5416, reduced hepatic cystogenesis in PC2‐defi­ volume as endpoint in contrast to previous studies in which
cient mice [50, 92]. cystic areas were measured in thin liver sections. Therefore, the
role of mTOR inhibitors as a therapeutic entity in PLD remains
unclear due to limited number of experimental reports and con­
PPARγ troversial outcomes.
Several studies suggest that pioglitazone (a full agonist of
PPARγ) and telmisartan (a partial PPARγ agonist) decreased Proto‐oncogene tyrosine‐protein kinase Src
liver cyst growth and elevated blood pressure in PCK rats
which is accompanied by inhibited expression of the trans­
(c‐Src)
forming growth factor beta 1 (TGFβ) in cystic cholangiocytes c‐Src, which phosphorylates tyrosine residues of different pro­
[93, 94]. teins, was shown to promote angiogenesis and proliferation in
different cell types. Cystic cholangiocytes of PCK rats have
increased c‐Src activity and a pharmacologic c‐Src inhibitor,
Histone deacetylase 6 (HDAC6)
SKI‐606, ameliorated biliary duct abnormalities in this model of
HDAC6 (in contrast to other members of the HDAC family that PLD [51].
epigenetically control gene expression) is predominantly local­
ized in the cytoplasm and is involved in regulation of multiple
cellular process including cell cycle progression, autophagy,
and ciliary disassembly. HDAC6 is upregulated in PLD cholan­ COMBINATIONAL DRUG THERAPIES
giocytes and its inhibition by tubastatin‐A, tubacin, and
ACY‐1215 decreased cell proliferation in vitro in a dose‐ and cAMP and mTOR
time‐dependent manner. Several HDAC6 inhibitors (i.e. panobi­
nostat, ACY‐1215, ACY‐738, and ACY‐241) reduced hepatic The effects of the somatostatin analog, lanreotide, and the
cystogenesis in PCK rats; however, ACY‐1215 exhibits stronger mTOR inhibitor, everolimus, on disease pathogenesis were
suppressive effects [35, 44]. evaluated in patients with ADPKD‐associated PLD after kidney
transplants. While lanreotide decreased total liver volume as
previously reported [64], no synergistic reduction on disease
Metalloproteinases (MMP) progression was achieved in response to drug combination [98].
Increased levels of different MMPs have been observed in PLD
cholangiocytes. Enhanced secretion of IL‐6, IL‐8, and 17β‐
cAMP and RAF
estradiol by cystic cholangiocytes promote in an autocrine/par­
acrine fashion hyperactivity of MMPs that, in turn, stimulate Given that in PC2‐deficient mice cAMP together with its down­
the digestion and remodeling of extracellular matrix [47]. When stream effector PKA activates RAS–RAF–MEK–ERK pathway
secretion of IL‐6 and/or IL‐8 was blocked, enhanced resulting in overgrowth of hepatic cysts, a therapeutic potential
­proliferation of cystic cholangiocytes was reduced. Importantly, of RAF inhibition by sorafenib was examined. Unexpectedly,
MMP inhibitor, marimastat, attenuated hepatic cystogenesis in sorafenib further accelerated hepatic cystogenesis and ERK
PCK rats [95]. phosphorylation by activating RAF1 in PC2‐deficient mice.
33:  Polycystic Liver Diseases: Genetics, Mechanisms, and Therapies 419

Combinational administration of octreotide and sorafenib abol­ cystogenesis facilitated the discovery of multiple potential
ished the paradoxical effects of the latter [84]. therapeutic targets for disease treatment. Many of these tar­
gets have been tested preclinically showing beneficial
effects on suppression of cell proliferation and cyst growth
cAMP in different experimental systems (i.e. cultured cystic chol­
Octreotide + pasireotide angiocytes, in vitro models of cystogenesis, and animal
models of PLD). However, these targets still have to be eval­
The effects of coadministration of octreotide and pasireotide on
uated in clinical trials. Mounting evidence suggests that
hepatic cystogenesis were examined in PCK rats. As previously
drug combinations might be more effective that monothera­
reported [33, 47], each somatostatin analog decreased hepatic
pies in PLD treatment. Clinically, somatostatin analogs and
cystogenesis; however, no additive effects in reduction of cyst
UDCA have been evaluated in patients with ADPKD‐­
growth were observed in response to drug combination [99].
associated PLD and isolated ADPLD, yet only somatostatin
analogs showed beneficial effects and are used as pharmaco­
Pasireotide + SBI‐115
therapy in PLD.
We recently reported that elevated cAMP in PLD cholangio­
cytes might be a result of overexpression of the cAMP‐linked
GPCR, TGR5 [38]. We also showed that TGR5 antagonist,
REFERENCES
SBI‐115, decreased cholangiocyte proliferation and growth of
cystic structures in vitro. Therefore, we tested how targeting of 1. Masyuk, T., Masyuk, A., and LaRusso, N. Cholangiociliopathies: genetics,
cAMP machinery by the somatostatin analog, pasireotide, and molecular mechanisms and potential therapies. Curr Opin Gastroenterol,
the TGR5 inhibitor, SBI‐115, affected these processes. More 2009;25(3):265–71.
substantial inhibition of cAMP levels, rate of proliferation, and 2. van Aerts, R.M.M. et  al. Clinical management of polycystic liver disease.
J Hepatol, 2017.
cyst growth in vitro was observed in response to concurrent drug 3. Neijenhuis, M.K. et al. Impact of liver volume on polycystic liver disease‐
administration compared to each drug alone [38]. Genetic related symptoms and quality of life. United European Gastroenterol J,
depletion of TGR5 in cultured cholangiocytes abolished the 2018;6(1):81–8.
effects of SBI‐115. While this result is encouraging, further 4. van de Laarschot, L.F.M. and Drenth, J.P.H. Genetics and mechanisms of
studies in animal models are needed. hepatic cystogenesis. Biochim Biophys Acta, 2018;1864(4 Pt B):1491–7.
5. Santos‐Laso, A.,et al. New advances in polycystic liver diseases. Semin Liver
Dis, 2017;37(1):45–55.
cAMP and autophagy 6. Khan, S., Dennison, A., and Garcea, G. Medical therapy for polycystic liver
disease. Ann R Coll Surg Engl, 2016;98(1):18–23.
When combined, pasireotide and HCQ reduced cAMP levels 7. D’Agnolo, H.M. et  al. Ursodeoxycholic acid in advanced polycystic liver
disease: a phase 2 multicenter randomized controlled trial. J Hepatol,
and decreased proliferation of cultured cystic cholangiocytes to
2016;65(3):601–7.
a greater extent compared to each drug alone. Consistent with 8. Besse, W. et  al. Isolated polycystic liver disease genes define effectors of
this, co‐treatment of PCK rats with pasireotide and HCQ showed polycystin‐1 function. J Clin Invest, 2017;127(9):3558.
additive effects of the drug combination on reduction in liver 9. Masyuk, T.V., Masyuk, A.I., and LaRusso, N.F. Polycystic liver disease: the
weight, hepatic cystic and fibrotic areas, and rate of cell prolif­ interplay of genes causative for hepatic and renal cystogenesis. Hepatology,
2018;67(6):2462–4.
eration within the liver cysts [37].
10. Masyuk, T.V., Masyuk, A.I., and LaRusso, N.F. Therapeutic targets in poly­
cystic liver disease. Curr Drug Targets, 2017;18(8):950–957.
11. Bergmann, C. Genetics of autosomal recessive polycystic kidney disease and
cAMP and HDAC6 its differential diagnoses. Front Pediatr, 2017;5:221.
Among several HDAC6 inhibitors (i.e. panobinostat, ACY‐1515, 12. Woollard, J.R. et  al. A mouse model of autosomal recessive polycystic
­kidney disease with biliary duct and proximal tubule dilatation. Kidney Int,
ACY‐738, and ACY‐241) tested in PCK rats, ACY‐1215 was 2007;72(3):328–36.
most effective. Therefore, the synergistic effects of ACY‐1215 13. Wills, E.S., Roepman, R., and Drenth, J.P. Polycystic liver disease:
and HDAC6 in combination on disease progression in vitro and ductal  plate malformation and the primary cilium. Trends Mol Med,
in vivo were tested. In cultured cystic cholangiocytes, a mixture 2014;20(5):261–70.
of ACY‐1215 and pasireotide more effectively decreased cell 14. Cnossen, W.R. and Drenth, J.P. Polycystic liver disease: an overview of
pathogenesis, clinical manifestations and management. Orphanet J Rare
proliferation and reduced cAMP levels compared to each drug Dis, 2014;9:69.
alone. In PCK rats, co‐administration of two drugs synergisti­ 15. Davila, S. et al. Mutations in SEC63 cause autosomal dominant polycystic
cally attenuated hepatic cystogenesis and increased length of liver disease. Nat Genet, 2004;36(6):575–7.
primary cilia showing additive effects of combinational therapy 16. Porath, B. et  al. Mutations in GANAB, encoding the glucosidase II alpha
subunit, cause autosomal‐dominant polycystic kidney and liver disease.
on these processes [35].
Am J Hum Genet, 2016;98(6):1193–207.
17. Drenth, J.P. et  al. Germline mutations in PRKCSH are associated with
­autosomal dominant polycystic liver disease. Nat Genet, 2003;33(3):345–7.
18. Waanders, E. et  al. Cysts of PRKCSH mutated polycystic liver disease
CONCLUSION patients lack hepatocystin but express Sec63p. Histochem Cell Biol,
2008;129(3):301–10.
19. Cnossen, W.R., et al. Whole‐exome sequencing reveals LRP5 mutations and
Considerable progress in our understanding of the genetic canonical Wnt signaling associated with hepatic cystogenesis. Proceedings
landscape in PLD and identification of dysregulated signal­ of the National Academy of Science,Science,,s of the United States of
ing pathways and abnormal cellular functions of hepatic America, 2014;111(14):5343–8.
420 THE LIVER:  REFERENCES

20. Besse, W. et al. A noncoding variant in GANAB explains isolated polycystic 45. Gradilone, S.A. et al. Cholangiocyte cilia express TRPV4 and detect changes
liver disease (PCLD) in a large family. Hum Mutat, 2018;39(3):378–82. in luminal tonicity inducing bicarbonate secretion. Proc Natl Acad Sci U S A,
21. Hirose, M. et  al. The mitochondrial uncoupling protein 2 gene is causal 2007;104(48):19138–43.
for  the spontaneous polycystic liver diseases in mice. Mitochondrion, 46. Gradilone, S.A. et  al. Activation of Trpv4 reduces the hyperproliferative
2017;42:50–3. phenotype of cystic cholangiocytes from an animal model of ARPKD.
­
22. Bergmann, C. and Weiskirchen, R. It’s not all in the cilium, but on the road Gastroenterol, 2010;139(1):304–14.
to it: genetic interaction network in polycystic kidney and liver diseases and 47. Masyuk, T.V. et  al. Octreotide inhibits hepatic cystogenesis in a rodent
how trafficking and quality control matter. J Hepatol, 2012;56(5):1201–3. model of polycystic liver disease by reducing cholangiocyte adenosine 3’,5’‐
23. Fedeles, S.V. et al. A genetic interaction network of five genes for human cyclic monophosphate. Gastroenterol, 2007;132(3):1104–16.
polycystic kidney and liver diseases defines polycystin‐1 as the central deter­ 48. Masyuk, T.V. et al. Inhibition of Cdc25A suppresses hepato‐renal cystogen­
minant of cyst formation. Nature Genetics, 2011;43(7):639–47. esis in rodent models of polycystic kidney and liver disease. Gastroenterol,
24. Masyuk, T.V. et  al. Biliary dysgenesis in the PCK rat, an orthologous 2012;142(3):622–33.
model  of autosomal recessive polycystic kidney disease. Am J Pathol, 49. Spirli, C. et al. Mammalian target of rapamycin regulates vascular endothe­
2004;165(5):1719–30. lial growth factor‐dependent liver cyst growth in polycystin‐2‐defective
25. Stroope, A. et  al. Hepato‐renal pathology in pkd2ws25/‐ mice, an animal mice. Hepatology, 2010;51(5):1778–88.
model of autosomal dominant polycystic kidney disease. Am J Pathol, 50. Spirli, C. et al. ERK1/2‐dependent vascular endothelial growth factor signal­
2010;176(3):1282–91. ing sustains cyst growth in polycystin‐2 defective mice. Gastroenterol,
26. Masyuk, T.V. et  al. Centrosomal abnormalities characterize human and 2010;138(1):360–71.
rodent cystic cholangiocytes and are associated with Cdc25A overexpres­ 51. Sweeney, W.E., Jr. et al. Src inhibition ameliorates polycystic kidney d­ isease.
sion. Am J Pathol, 2014;184(1):110–21. J Am Soc Nephrol, 2008;19(7):1331–41.
27. Perugorria, M.J. et al. Polycystic liver diseases: advanced insights into the 52. Alvaro, D. et al. Morphological and functional features of hepatic cyst epi­
molecular mechanisms. Nat Rev Gastroenterol Hepatol, 2014;11(12): thelium in autosomal dominant polycystic kidney disease. Am J Pathol,
750–61. 2008;172(2):321–32.
28. Raynaud, P. et  al. Biliary differentiation and bile duct morphogenesis in 53. Waanders, E. et  al. Disrupted cell adhesion but not proliferation mediates
development and disease. Int J Biochem Cell Biol, 2011;43(2):245–56. cyst formation in polycystic liver disease. Mod Pathol, 2008;21(11):
29. Banales, J.M. et al. The cAMP effectors Epac and protein kinase a (PKA) are 1293–302.
involved in the hepatic cystogenesis of an animal model of autosomal reces­ 54. Beaudry, J.B. et  al. Proliferation‐independent initiation of biliary cysts in
sive polycystic kidney disease (ARPKD). Hepatology, 2009;49(1):160–74. polycystic liver diseases. PLoS One, 2015;10(6):e0132295.
30. Masyuk, A.I., Masyuk, T.V., and LaRusso, N.F. Cholangiocyte primary cilia 55. Li, D. et al. Overexpression of aquaporin 1 on cysts of patients with polycys­
in liver health and disease. Dev Dyn, 2008;237(8):2007–12. tic liver disease. Rev Esp Enferm Dig, 2016;108(2):71–8.
31. Masyuk, T.V. et al. Defects in cholangiocyte fibrocystin expression and cili­ 56. Masyuk, T., Masyuk, A., and LaRusso, N., MicroRNAs in cholangiociliopa­
ary structure in the PCK rat. Gastroenterol, 2003;125(5):1303–10. thies. Cell Cycle, 2009;8(9):1324–8.
32. Strazzabosco, M. and Somlo, S. Polycystic liver diseases: congenital disor­ 57. Perugorria, M.J. et  al. Bile acids in polycystic liver diseases: triggers of
ders of cholangiocyte signaling. Gastroenterol, 2011;140(7):1855–9. ­disease progression and potential solution for treatment. Dig Dis, 2017;35(3):
33. Masyuk, T.V. et al. Pasireotide is more effective than octreotide in reducing 275–81.
hepatorenal cystogenesis in rodents with polycystic kidney and liver dis­ 58. Larusso, N.F., Masyuk, T.V., and Hogan, M.C. Polycystic liver disease:
eases. Hepatology, 2013;58(1):409–21. the  benefits of targeting cAMP. Clin Gastroenterol Hepatol, 2016;14(7):
34. Lee, S.O. et al. MicroRNA15a modulates expression of the cell‐cycle regula­ 1031–4.
tor Cdc25A and affects hepatic cystogenesis in a rat model of polycystic 59. Hogan, M.C. et al. Randomized clinical trial of long‐acting somatostatin for
kidney disease. J Clin Invest, 2008;118(11):3714–24. autosomal dominant polycystic kidney and liver disease. J Am Soc Nephrol,
35. Lorenzo Pisarello, M. et al. Combination of a histone deacetylase 6 inhibitor 2010;21(6):1052–61.
and a somatostatin receptor agonist synergistically reduces hepatorenal cys­ 60. Gevers, T.J. and Drenth, J.P., Somatostatin analogues for treatment of poly­
togenesis in an animal model of polycystic liver disease. Am J Pathol, cystic liver disease. Curr Opin Gastroenterol, 2011;27(3):294–300.
2018;188(4):981–94. 61. Temmerman, F. et al. Safety and efficacy of different lanreotide doses in the
36. Masyuk, A.I. et al. Ciliary subcellular localization of TGR5 determines the treatment of polycystic liver disease: pooled analysis of individual patient
cholangiocyte functional response to bile acid signaling. Am J Physiol data. Aliment Pharmacol Ther, 2013;38(4):397–406.
Gastrointest Liver Physiol, 2013;304(11):G1013–24. 62. Pisani, A. et al. Long‐term effects of octreotide on liver volume in patients
37. Masyuk, A.I. et  al. Cholangiocyte autophagy contributes to hepatic cys­ with polycystic kidney and liver disease. Clin Gastroenterol Hepatol,
togenesis in polycystic liver disease and represents a potential therapeutic 2016;14(7):1022–30.
target. Hepatology, 2018;67(3):1088–108. 63. van Keimpema, L., de Man, R.A. and Drenth, J.P., Somatostatin analogues
38. Masyuk, T.V. et al. TGR5 contributes to hepatic cystogenesis in rodents with reduce liver volume in polycystic liver disease. Gut, 2008;57(9):1338–9.
polycystic liver diseases through cyclic adenosine monophosphate/Galphas 64. van Keimpema, L. et al. Lanreotide reduces the volume of polycystic liver:
signaling. Hepatology, 2017;66(4):1197–218. a  randomized, double‐blind, placebo‐controlled trial. Gastroenterol,
39. Munoz‐Garrido, P. et al. Ursodeoxycholic acid inhibits hepatic cystogenesis 2009;137(5):1661–8.
in experimental models of polycystic liver disease. J Hepatol, 2015;63(4): 65. Caroli, A. et  al. Reducing polycystic liver volume in ADPKD: effects of
952–61. somatostatin analogue octreotide. Clin J Am Soc Nephrol, 2010;5(5):
40. Masyuk, A.I., Gradilone, S.A., and LaRusso, N.F. Calcium signaling in cilia 783–9.
and ciliary‐mediated intracellular calcium signaling: are they independent or 66. Hogan, M.C. et al. Somatostatin analog therapy for severe polycystic liver
coordinated molecular events? Hepatology, 2014;60(5):1783–5. disease: results after 2 years. Nephrology, dialysis, transplantation: Nephrol
41. Spirli, C. et al. Altered store operated calcium entry increases cyclic 3′,5′‐ Dial Transplant, 2012;27(9):3532–9.
adenosine monophosphate production and extracellular signal‐regulated 67. Neijenhuis, M.K. et al. Somatostatin analogues improve health‐related qual­
kinases 1 and 2 phosphorylation in polycystin‐2‐defective cholangiocytes. ity of life in polycystic liver disease: a pooled analysis of two randomised,
Hepatology, 2012;55(3):856–68. placebo‐controlled trials. Aliment Pharmacol Ther, 2015;42(5):591–8.
42. Yang, J. et  al. Deficiency of hepatocystin induces autophagy through an 68. Gevers, T.J. et  al. Effect of lanreotide on polycystic liver and kidneys in
mTOR‐dependent pathway. Autophagy, 2011;7(7):748–59. autosomal dominant polycystic kidney disease: an observational trial. Liver
43. Banales, J.M. et al. Hepatic cystogenesis is associated with abnormal expres­ Int, 2015;35(5):1607–14.
sion and location of ion transporters and water channels in an animal 69. Hogan, M.C. et al. Efficacy of 4 years of octreotide long‐acting release ther­
model  of autosomal recessive polycystic kidney disease. Am J Pathol, apy in patients with severe polycystic liver disease. Mayo Clin Proc,
2008;173(6):1637–46. 2015;90(8):1030–7.
44. Gradilone, S.A. et al. HDAC6 is overexpressed in cystic cholangiocytes and 70. Drenth, J.P. et al. Medical and surgical treatment options for polycystic liver
its inhibition reduces cystogenesis. Am J Pathol, 2014;184(3):600–8. disease. Hepatology, 2010;52(6):2223–30.
33:  Polycystic Liver Diseases: Genetics, Mechanisms, and Therapies 421

71. Chrispijn, M. et al. The long‐term outcome of patients with polycystic liver 88. Mizuno, H. et al. Tolvaptan for the treatment of enlarged polycystic liver
disease treated with lanreotide. Aliment Pharmacol Ther, 2012;35(2):266–74. disease. Case Rep Nephrol Dial, 2017;7(3):108–11.
72. Gevers, T.J. et al. Young women with polycystic liver disease respond best to 89. Holditch, S.J. et al. B‐type natriuretic peptide overexpression ameliorates
somatostatin analogues: a pooled analysis of individual patient data. hepatorenal fibrocystic disease in a rat model of polycystic kidney disease.
Gastroenterol, 2013;145(2):357–65. Kidney Int, 2017;92(3):657–68.
73. Temmerman, F. et al. Safety and efficacy of different lanreotide doses in the 90. Smithline, Z.B. et al. Inhibiting heat shock protein 90 (HSP90) limits the
treatment of polycystic liver disease: pooled analysis of individual patient formation of liver cysts induced by conditional deletion of Pkd1 in mice.
data. Aliment Pharmacol Ther, 2013;38(4):397–406. PLoS One, 2014;9(12):e114403.
74. Temmerman, F. et  al. Lanreotide reduces liver volume, but might not 91. Fabris, L. et al. Effects of angiogenic factor overexpression by human and
improve muscle wasting or weight loss, in patients with symptomatic poly­ rodent cholangiocytes in polycystic liver diseases. Hepatology,
cystic liver disease. Clin Gastroenterol Hepatol, 2015;13(13):2353–9. 2006;43(5):1001–12.
75. Wijnands, T.F. et al. Efficacy and safety of aspiration sclerotherapy of simple 92. Amura, C.R. et  al. VEGF receptor inhibition blocks liver cyst growth in
hepatic cysts: a systematic review. Am J Roentgenol, 2017;208(1):201–7. pkd2(WS25/‐) mice. Am J Physiol Cell Physiol, 2007;293(1):C419–28.
76. Marin, J.J. et al. Bile acids in physiology, pathology and pharmacology. Curr 93. Yoshihara, D. et al. Telmisartan ameliorates fibrocystic liver disease in an
Drug Metab, 2015;17(1):4–29. orthologous rat model of human autosomal recessive polycystic kidney
77. Poupon, R. Ursodeoxycholic acid and bile‐acid mimetics as therapeutic ­disease. PloS One, 2013;8(12):e81480.
agents for cholestatic liver diseases: an overview of their mechanisms of 94. Yoshihara, D. et al. PPAR‐gamma agonist ameliorates kidney and liver dis­
action. Clin Res Hepatol Gastroenterol, 2012;36(1):S3–12. ease in an orthologous rat model of human autosomal recessive polycystic
78. Beuers, U. et al. New paradigms in the treatment of hepatic cholestasis: from kidney disease. Am J Physiol Renal Physiol, 2011;300(2):F465–74.
UDCA to FXR, PXR and beyond. J Hepatol, 2015;62(1):S25–37. 95. Urribarri, A.D. et al. Inhibition of metalloprotease hyperactivity in cystic
79. Glaser, S. et al. Adrenergic receptor agonists prevent bile duct injury induced cholangiocytes halts the development of polycystic liver diseases. Gut,
by adrenergic denervation by increased cAMP levels and activation of Akt. 2014;63(10):1658–67.
Am J Physiol Gastrointest Liver Physiol, 2006;290(4):G813–26. 96. Renken, C. et  al. Inhibition of mTOR with sirolimus does not attenuate
80. Iijima, T. et al. Ursodeoxycholic acid for treatment of enlarged polycystic progression of liver and kidney disease in PCK rats. Nephrol Dial
liver. Ther Apher Dial, 2016;20(1):73–8. Transplan, 2011;26(1):92–100.
81. Keitel, V. and Haussinger, D. TGR5 in cholangiocytes. Curr Opin 97. Temmerman, F. et  al. Everolimus halts hepatic cystogenesis in a rodent
Gastroenterol, 2013;29(3):299–304. model of polycystic‐liver‐disease. World J Gastroenterol, 2017;23(30):
82. Masyuk, A.I. et al. Cholangiocyte cilia detect changes in luminal fluid flow 5499–507.
and transmit them into intracellular Ca2+ and cAMP signaling. Gastroenterol, 98. Chrispijn, M. et al. Everolimus does not further reduce polycystic liver vol­
2006;131(3):911–20. ume when added to long acting octreotide: results from a randomized con­
83. Spirli, C. et al. Adenylyl cyclase 5 links changes in calcium homeostasis to trolled trial. J Hepatol, 2013;59(1):153–9.
cAMP‐dependent cyst growth in polycystic liver disease. J Hepatol, 99. Kugita, M. et al. Beneficial effect of combined treatment with octreotide
2017;66(3):571–80. and pasireotide in PCK rats, an orthologous model of human autosomal
84. Spirli, C. et al. Cyclic AMP/PKA‐dependent paradoxical activation of Raf/ recessive polycystic kidney disease. PLoS One, 2017;12(5):e0177934.
MEK/ERK signaling in polycystin‐2 defective mice treated with sorafenib. 100. Mancinelli, R. et al. Follicle‐stimulating hormone increases cholangiocyte
Hepatology, 2012;56(6):2363–74. proliferation by an autocrine mechanism via cAMP‐dependent phospho­
85. Onori, P. et al. Role of follicle‐stimulating hormone on biliary cyst growth in rylation of ERK1/2 and Elk‐1. Am J Physiol Gastrointest Liver Physiol,
autosomal dominant polycystic kidney disease. Liver Int, 2013;33(6):914–25. 2009;297(1):G11–26.
86. Mancinelli, R. et al. Vasopressin regulates the growth of the biliary epithe­ 101. Muff, M., Masyuk, T., Stroope, A. et al. Development and characterization
lium in polycystic liver disease. Lab Invest, 2016;96(11):1147–55. of a cholangiocyte cell line from the PCK rat, an animal model of autoso­
87. Edwards, M.E. et al. Long‐term administration of tolvaptan in autosomal dom­ mal recessive polycystic kidney disease. Laboratory Investigation 2006;
inant polycystic kidney disease. Clin J Am Soc Nephrol, 2018;13(8):1153–61. 86:940–950. PMID:16783394.
The Liver Sinusoidal
34 Endothelial Cell: Basic
Biology and Pathobiology
Karen K. Sørensen and Bård Smedsrød
Vascular Biology Research Group, Department of Medical Biology, UiT The Arctic University of Norway,
Tromsø, Norway

INTRODUCTION for example, atherosclerosis, systemic lupus erythematosus


(SLE), and glomerular fibrotic nephrosis.
Endothelial cells constitute a heterogenous cell population This update starts by outlining central aspects of the basic
with distinct structural and functional characteristics depend- biology of the LSEC. The importance of LSECs in the etiology
ing on tissue type. Different from other microvascular of selected pathologies of the liver and extrahepatic sites will
endothelial cells, the liver sinusoidal endothelial cells then be described. A section on LSEC‐mediated liver toxicity is
(LSECs) are highly perforated and lack an organized basal also included.
lamina [1]. This signature morphological feature of the LSEC
allows free passage of macromolecules and colloids, such as
lipoproteins, across the sinusoidal lining. The functional char-
acteristics of LSECs include their active endocytosis recep- STRUCTURE–FUNCTION
tors endowing the cells with an extraordinary high endocytic
CONSIDERATIONS
capability [2]. This makes these cells the body’s most effec-
tive scavenger cells, clearing blood‐borne macromolecules The hepatocytes represent the major cell type in the liver (92.5%
and nanosized compounds, thus contributing importantly to of the total liver cell volume), with LSECs, Kupffer cells, and
maintaining homeostasis. A special metabolic feature of stellate cells making up 3.3%, 2.5%, and 1.7% of the cell v­ olume
LSECs is their very high production of lactate signifying their [9]. In spite of the volume‐wise predominance of the hepato-
largely anaerobic metabolism, which is also reflected in their cytes, the minor cell populations play pivotal functional roles
low number of mitochondria. LSECs represent an important due to the anatomical and morphological organization of the
player in immunity: they express the endocytic Fc‐gamma organ. The liver is highly vascularized, receiving about a quarter
receptor IIb2 (CD32b, FcγRIIb2), several pattern recognition of cardiac output. The blood that enters liver is a mixture of 25%
receptors (PRRs), for example, various toll‐like receptors oxygen‐rich blood from the hepatic artery and 75% oxygen‐poor
(TLRs), as well as the mannose receptor (MR) and several blood that enters via the portal vein. The portal vein carries gut‐
scavenger receptors (SRs), of which stabilin‐1 and stabilin‐2 derived nutrients and potentially or overtly harmful s­ ubstances
represent the most important LSEC SRs. LSECs further dis- that will be surveilled and if needed eliminated by the liver,
play features of adaptive immunity, contributing to hepatic ensuring that blood leaving the liver is compatible with normal
immune tolerance [3]. homeostasis. This efficient filtering capability of the liver can be
In liver pathobiology, LSECs play central roles in liver fibro- mainly ascribed to the following three structure/function features
sis [4, 5], cancer metastasis [6, 7], and liver toxicology [8]. In of the liver sinusoid:
addition to being involved in the etiology of pathologies of the
liver, it is noteworthy that the LSECs are also believed to have a 1. The cells facing the blood flowing through the sinusoids
role in the development of complications at extrahepatic sites, include mainly the LSECs that make up the wall of the

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
34:  The Liver Sinusoidal Endothelial Cell: Basic Biology and Pathobiology 423

(a)

SC
LSEC WBC

RBC
KC

Space of Disse

HC

(b)
KC
LSEC

Virus Bacteria
Oligonucleotides
Nanoparticles

Proteins Red blood cells

1nm 10 nm 100 nm 1μm 10 μm

Figure 34.1  (a) The liver sinusoid with its main cell types. The cartoon illustrates the localization of hepatocytes (HC), liver sinusoidal endothelial
cells (LSEC), Kupffer cells (KC), and stellate cells (SC). The Kupffer cells are normally located at the luminal aspect of the sinusoid, whereas the
stellate cells are located between the LSECs and hepatocytes, in the space of Disse. RBC, red blood cell; WBC, white blood cell. Blue and green
dots represent soluble macromolecules. The figure is reproduced from [2] with permission. (b) Share of scavenging workload between LSECs and
Kupffer cells. Together LSECs and Kupffer cells make up the largest population of scavenger cells in the body [10]. LSECs are geared to effective
clathrin‐mediated endocytosis and avidly endocytose a wide range of soluble macromolecules and nanoparticles. LSECs are essentially non‐phago-
cytic, and larger particles, such as bacteria and aged and damaged red blood cells are therefore cleared by the Kupffer cells, illustrating “the dual
cell principle of waste handling” [11]. The figure is a courtesy from Dr Kjetil Elvevold, D’Liver AS Tromsø, Norway.

numerous hepatic sinusoids (Figure 34.1a), and the Kupffer THE LSEC SCAVENGER FUNCTION
cells, representing by far the largest population of mac- AND ITS IMPLICATIONS
rophages in our body. Both of these cell types have unsur-
passed ability to recognize and endocytose soluble and
particulate material that is incompatible with normal homeo-
Historical considerations
stasis. Together the two cell types make up the largest scaven- It was only after detailed ultrastructural studies carried out by
ger cell system in the body [10, 11] (Figure 34.1b). Eddie Wisse in the early 1970s that LSECs were distinguished
2. The LSECs are perforated with pores 100–150 nm in as a cell type distinct from the Kupffer cells [1]. Development of
diameter, commonly referred to as fenestration, and occu- a method for mass isolation of LSECs, Kupffer cells, and hepat-
pying as much as 20% of the LSEC [12]. The LSECs lack ocytes from a single rat liver, allowing establishment of primary
a basal lamina and form the only fenestrated endothelium cultures of purified cells, enabled the discovery that the extra-
that offers open passage of macromolecules and nanopar- cellular matrix polysaccharide hyaluronan was avidly and spe-
ticles between the blood and underlying parenchyma cifically endocytosed and degraded mainly in LSECs, but not in
(Figure 34.2). other types of liver cells [14]. This assigned a role for the LSEC
3. Also contributing to the high blood filtering capability of the as a novel type of physiologically important scavenger cell [15].
liver is the large total surface area of the sinusoidal endothe- Over the years that followed, the list of macromolecules found
lium. Data from stereological analysis of the rat liver [13] to be cleared mainly by LSECs increased steadily, and it is now
can be used to calculate that the total LSEC surface area of generally accepted that LSECs are far more important than
an average normal adult human liver is approx. 210 m2 (i.e. Kupffer cells in the clearance of soluble macromolecules and
80% of a tennis court). nanomaterial from the circulation [2] (Table 34.1). This notion
424 THE LIVER:  THE LSEC SCAVENGER FUNCTION AND ITS IMPLICATIONS

(a) (b)

HC

LSEC SD HC

HC

LSEC
1µm HC 1µm

(c) (d)

PV

50 µm 2 µm

Figure 34.2  Morphology of the liver sinusoidal endothelial cell. (a) Scanning electron micrograph of mouse liver showing the high porosity of
the liver sinusoidal endothelium. The fenestrae of the LSECs are typically organized in groups, termed sieve plates (encircled by a white dashed
line). HC, hepatocyte; SD, space of Disse. (b) Transmission electron micrograph of a rat liver sinusoid. The very thin cytoplasm of the LSEC con-
tains many fenestrations (arrow heads), and there is no organized basal membrane beneath the endothelium. The fenestrae provide open channels
allowing bidirectional traffic of fluids, solutes, and small particles between the blood and hepatocytes. (c) Immune fluorescence micrograph of
mouse liver, showing uptake of BK polyomavirus‐like particles (green fluorescence, arrow heads) in the sinusoidal endothelium 15 minutes after
intravenous injection, visualized by an anti‐BK polyomavirus antibody. PV, portal vein. The experiment was carried out as part of the study
described in [69]. (d) Scanning electron micrograph of rat LSEC. Freshly isolated LSECs were plated on collagen coated tissue culture plastic and
fixed one hour after seeding. The cells are highly fenestrated; the white dashed line encircles a sieve plate.

came as a surprise to many researchers in the field because it in Kupffer cells. To resolve this misconception experiments
deviated from the paradigm that hepatic blood clearance is a have been performed in exactly the same way as described by
function of the reticuloendothelial system (RES), a term that has Aschoff and his contemporaries, and with vital stains prepared
been erroneously used synonymously with the macrophage, or according to the original protocols as those used 100–140 years
the mononuclear phagocyte system. The name and notion of ago. Thus revisiting these early studies about 75 years after
RES, launched nearly 100 years ago by Ludwig Aschoff [16], Aschoff launched his concept of RES revealed that the uptake of
was based on the consistent observation that different types of the most frequently used vital stain, lithium‐carmine, was pre-
vital stains (i.e. colloidal, nanosized compounds) accumulated dominantly in LSECs [18].
in specific anatomical locations, and most notably along the
hepatic sinusoids. Although Aschoff himself did not link RES
specifically to macrophage or phagocytosis, scientists gradually
The dual cell principle of waste clearance
came to view the RES as a cellular system responsible for Thanks to the studies reported in the above paragraph we know
phagocytic uptake of blood‐borne colloidal vital stains mainly today that the liver scavenger system consists of both the Kupffer
or exclusively in macrophages. The reason why this happened cells, that are geared mainly to phagocytic uptake of particles
must be understood in light of the fact that although the terms greater than 200 nm, and LSECs that clear blood‐borne macromol-
macrophage and phagocytosis had been introduced by ecules and nanoparticles less than 200 nm (Figure 34.1b). This con-
Metchnikoff [17] several years prior to Aschoff’s publication, cept, referred to as the dual cell principle of waste clearance [11]
the perception of these terms have been considerably developed must be taken into consideration when aiming to control or predict
up through the years. Yet, even at the present time it is not uptake of blood‐borne compounds in liver. Of note, the principle
uncommon to find scientific papers and updated textbooks applies to animal species of all vertebrate classes. However, the
where hepatic uptake is described merely as phagocytic uptake scavenger endothelial cells are mainly in liver in land‐based
34:  The Liver Sinusoidal Endothelial Cell: Basic Biology and Pathobiology 425

Table 34.1  Ligands cleared from the blood circulation mainly by the LSECs


Ligand Receptor Reference
Tissue turnover waste products Hyaluronan Stabilin‐2 [14, 25]
Chondroitin sulfate Stabilin‐2 [14, 139]
Nidogen SR [140]
Heparin Stabilin‐2 [139, 141]
Serglycin SR [142]
N‐terminal propeptides of procollagen (I, III) SR, stabilin‐2 [25, 143]
Collagen alpha chains (I, II, III, IV, V, XI) Mannose receptora [15, 39, 144]
C‐terminal propeptide of procollagen type I Mannose receptor [36]
Tissue plasminogen activator Mannose receptor [37]
Lysosomal enzymes Mannose receptor [22]
Salivary amylase Mannose receptor [145]
Modified proteins and Oxidized LDL SR, stabilin‐1, stabilin‐2 [31, 146]
lipoproteins AGE‐albumin SR, stabilin‐2 (stabilin‐1b) [132, 147]
Immune complexes Soluble IgG immune complexes Fc‐γRIIb2 [44]
Ligands of non‐mammalian CpG oligodeoxynucleotides SR [33]
origin Invertase Mannose receptor [148]
Mannan Mannose receptor [41]
Ovalbumin Mannose receptor [21]
Ricin Mannose receptor [68]
Horseradish peroxidase Mannose receptor [149]
Aminated β(1–3) glucan Unknown [150]
Non‐physiological model ligands Formaldehyde‐treated albumin SR, stabilin‐1, stabilin‐2 [31, 151]
Acetylated LDL SR (stabilin‐1/ stabilin‐2b) [152]
Agalacto‐orosomucoid Mannose receptor [149]
Ahexosamino‐orosomucoid Mannose receptor [149]
Lithium carmine Unknown [18]
The table is modified from [11] with permission. SR, scavenger receptor; AGE, advanced glycation end‐product; LDL, low density lipoprotein. Stabilin‐1 synonyms: SR‐H1,
FEEL‐1 (fasciclin endothelial growth factor (EGF)‐like, laminin‐type EGF‐like and link domain‐containing scavenger receptor 1) [153], CLEVER‐1 (common lymphatic endothelial
and vascular endothelial receptor‐1) [154]. Stabilin‐2 synonyms: SR‐H2, FEEL‐2 [155], HA/SR (hyaluronan/SR)[25], HARE (hyaluronan receptor for endocytosis) [106].
a
 The reported LSEC receptor for collagen α‐chains is identical to the mannose receptor on these cells [39]. Collagen α‐chains and mannose‐terminated ligands have affinity to
non‐overlapping binding sites on the receptor [38].
b
 Affinity of AGE‐albumin to stabilin‐1, and acetylated LDL to stabilin‐1 and stabilin‐2 have been studied in transfected cell lines [153, 155].

vertebrates (i.e. mammals, birds, reptiles, and amphibia), whereas Receptors important for LSEC scavenger
in phylogenetically older vertebrates the scavenger endothelial function
cells are located in heart atrium and/or kidney (bony fishes) or gills
(cartilaginous fishes, lamprey, and hagfish) [10]. Several LSEC receptor species have been reported to bind
blood‐borne macromolecules. In this section only major endo-
cytically active receptors known to play a chief role in the
Additional evidence that LSECs are geared LSEC‐mediated blood clearance will be highlighted
to effective uptake of blood‐borne (Table 34.1). Other LSEC receptors for which important scav-
macromolecules and nanosized material enger function is not obvious will be briefly mentioned at the
end of the section.
LSECs are packed with organelles involved in endocytosis, and
intracellular transport and endo‐lysosomal processing of inter-
Scavenger receptors (SRs)
nalized cargo. In spite of their low proportion of the total liver
cell volume (3.3%), LSECs contain an impressive 45% of the Scavenger receptors (SRs) represent important LSEC clearance
organ’s total mass of pinocytic vesicles and 17% of the lysoso- receptors. The concept of SR has evolved considerably since the
mal volume [13]. Moreover, LSECs, compared to hepatocytes receptor for acetylated low‐density lipoprotein (LDL) was dis-
and Kupffer cells, contain twice as many clathrin‐coated pits per covered in macrophages and denoted “scavenger receptor” [24].
membrane unit [19]. The fact that LSECs contain high amounts The common denominator of all SRs is their affinity for
of proteins involved in membrane traffic (clathrin, alpha and ­polyanionic ligands. The LSEC is reported to express SRs class
beta‐adaptins, Rab4, Rab5, Rab7, and rabaptin‐5) [19, 20] A, B, E, and H [11]. Of these, SR‐H1 (stabilin‐1) and SR‐H2
strengthens the notion that the cells are indeed geared to clath- (stabilin‐2) are instrumental for LSEC‐mediated scavenging
rin‐mediated endocytosis. Another factor that is rarely dis- [11]. In the liver stabilin‐2 is expressed on LSECs only, and thus
cussed, but nevertheless contributes to the superefficient is a signature receptor than can be used as a specific LSEC
endocytosis in LSECs is the much more rapid receptor recycling marker [25–28]. Interestingly the Stab2 gene is highly con-
time in LSEC (seconds) [21] compared to most other cell types served, and antibodies to human stabilin‐2 stain the same protein
(minutes). These features, along with the fact that the specific in fish (Atlantic cod) [11]. Moreover, in zebrafish where the
activity of several lysosomal enzymes is higher in LSECs than Stab2 gene had been knocked out, clearance of ligands for stabi-
in hepatocytes and Kupffer cells [22, 23] indicate that LSECs lin‐2 in scavenger endothelial cells was completely lost [29].
play a major role in the hepatic blood clearance function. Blood‐borne ligands cleared by LSEC stabilins include several
426 THE LIVER:  THE LSEC SCAVENGER FUNCTION AND ITS IMPLICATIONS

molecules that are products of physiological turnover of connec- The Fc‐gamma receptor IIb2
tive tissue [2] (Table  34.1). Physiologically modified proteins
The Fc‐gamma receptor IIb2 (FcγRIIb2, CD32b) is highly and
such as advanced glycation end products (AGEs) and oxidized
uniquely expressed in LSECs in liver, enabling efficient endocy-
LDL (oxLDL) are also cleared by LSEC stabilins [30, 31]. Of
tosis of soluble IgG‐antigen complexes [44]. The first hint that
note, mildly oxLDL (the major circulating form of oxLDLs) is
not only Kupffer cells, but also LSECs bind IgG immune com-
endocytosed in LSECs only, mainly via stabilin‐1, whereas non‐
plexes was reported in the early 1980s, when it was found that
physiological heavily oxidized oxLDL is endocytosed by both
isolated LSECs formed rosettes with IgG‐coated red blood cells
LSECs and Kupffer cells [31], pointing to the importance of
[45]. Another study showed that LSECs at high capacity could
LSECs in preventing cholesterol accumulation. SR class B, type
bind, but not phagocytose the same test particles [46]. The find-
I (SR‐BI), responsible for mediating the selective uptake of high‐
ing that soluble IgG immune complexes could compete with
density lipoprotein (HDL) cholesteryl esters in hepatocytes, has
IgG‐red blood cells for binding to the LSECs strengthened the
been identified in LSECs as well, where the receptor may be
notion that the binding of IgG‐red cells was FcγR‐mediated
responsible for the efflux of cellular cholesterol to HDL, deliver-
[47]. At the same time it was reported that small circulating IgG
ing cholesterol from LSECs to hepatocytes [32]. Nucleotides, for
immune complexes were taken up in LSECs, and it was noted
example, oligonucleotides containing unmethylated CpG motifs
that the relative proportions removed by Kupffer cells and
[33], anti‐sense‐oligonucleotides [34], and plasmid DNA [35]
LSECs depended on the size of the immune complex: the larger
represent a different group of molecules that are cleared by SR‐
the complex the more was directed to the Kupffer cells [48]. Of
mediated uptake in LSECs.
note, the LSECs account for at least 70% of the uptake of small
IgG immune complexes in mouse liver sinusoids [49]. Thus
The mannose receptor FcγR‐mediated uptake agrees with the dual cell principle of
The mannose receptor (MR, CD206, Mrc1), a 175–180 kDa waste clearance, in that small soluble immune complexes are
type I integral membrane protein belonging to the C‐type lec- cleared by LSECs, whereas larger immune complexes are endo-
tin family, has traditionally been associated with macrophages. cytosed by Kupffer cells. More recent studies have revealed that
However, several other cell types express this receptor, not the FcγRIIb2 is the main receptor for the efficient uptake of immune
least the LSECs, which is the main carrier of this receptor in complexes in rat LSECs, via clathrin‐mediated endocytosis
liver [2]. The importance of the LSEC MR is reflected by the [44]. The authors reported that the receptor was present on
consistent observation that intravenously administered solu- LSECs only, but not in Kupffer cells or other types of liver cells.
ble/macromolecular MR ligands are rapidly cleared from the Consistent with these findings in rats it has later been shown
circulation mainly by uptake in the LSECs [11]. The reason that FcγRIIb2 in LSECs is responsible for the clearance of solu-
for the remarkably efficient MR‐mediated uptake by LSECs is ble immune complexes in mice as well, with 75% and 90% of
its rapid receptor recycling time (only 10 seconds) [21]. The total body and total liver FcγRIIb2 expressed in LSECs [50].
MR is highly versatile in its function as a clearance receptor: Due to the consistent observation that FcγRIIb2 removes blood‐
several C‐type lectin‐like domains displayed by the receptor borne soluble immune complexes, several studies have impli-
protein are responsible for its binding to mannose and cated the LSECs as important in the etiology of autoimmune
N‐acetylglucosamine (GlcNAc), enabling LSECs to scavenge diseases, for example, systemic lupus erythematosus (SLE).
an array of ­different macromolecules carrying mannose or This will be dealt with in the section on the Role of LSECs in
GlcNAc in the glycosylation terminal position. The binding the etiology of autoimmune disorders.
specificity of the MR is essentially independent of the class of
Other endocytosis receptors in LSECs
macromolecule that carries the glycosylation side‐chains.
Important physiological ligands include C‐terminal propep- Compared to the relatively rich literature on LSEC SRs (stabi-
tides of collagens [36] and tissue plasminogen activator [37]. lin‐1 and stabilin‐2), MR, and FcγRIIb2, little is known about
Interestingly, MR‐mediated uptake of blood‐borne lysosomal the functional significance of other LSEC receptors, including
enzymes represents a physiological route to maintain the high LYVE‐1 (lymphatic vessel endothelial hyaluronan receptor‐1)
specific activity of these enzymes in LSECs [22]. The MR also [51], and the C‐type lectins L‐SIGN (liver/lymph node‐specific
expresses a fibronectin‐type II domain, which mediates the ICAM‐3 grabbing nonintegrin; syn. DC‐SIGNR) and LSECtin
scavenging of blood‐borne collagen alpha‐chains [38], result- (liver and lymph node sinusoidal endothelial cell C‐type lectin;
ing from physiological turnover processes of bone and other syn. DC‐SIGN) [52, 53]. L‐SIGN is involved in recognition and
connective tissues (estimated to be in the order of grams per uptake of human immune deficiency virus and hepatitis C virus
day) [39]. Yet a third ligand binding site on the MR is the in LSEC [52, 54], whereas LSECtin regulates viral immune
­outermost cysteine‐rich N‐terminal domain, which recognizes responses by interacting with L‐SIGN [55], and has also been
specific sulfated sugars [40]. Cytokines and inflammatory reported to have a role in adherence of cancer metastasis in liver
stimuli affect the expression of LSEC MR: upregulation has [7]. LYVE‐1 is a hyaluronan binding protein expressed in lym-
been reported after exposure to IL‐1 [6, 41, 42], IL‐10, or IL‐4 phatic vascular endothelium and sinusoidal endothelial cells of
in combination with IL‐13 [43]. In a mouse colon carcinoma lymph nodes, liver, and spleen [28]. In adult liver, LYVE‐1 is
liver metastasis model MR expression was observed to increase only expressed in the LSECs [51], and is distributed all along
along the sinusoids, which was suggested to contribute to the the sinusoids [56], making it a good LSEC marker. LYVE ‐1
hepatic prometastatic effects of IL‐1, cyclic oxygenase‐2, and clears hyaluronan from lymph, but its function in liver is unclear.
ICAM‐1 [6]. Stabilin‐2 shows a high capacity for hyaluronan‐uptake in
34:  The Liver Sinusoidal Endothelial Cell: Basic Biology and Pathobiology 427

LSECs, and the relative contribution of LYVE‐1 and stabilin‐2 a useful way of identifying the cells is to determine their ability
to this clearance is unknown. to accumulate ligands known to be recognized by any of these
receptors. To this end FITC‐labeled formaldehyde‐modified
serum albumin (FITC‐FSA), a ligand for stabilin‐1 and stabi-
LSEC markers
lin‐2 [31], has been frequently used [60]. In liver FITC‐FSA
The gold standard of LSEC markers is the characteristic open accumulates predominantly in LSECs when administered
fenestrae that, until recently, could only be detected in the elec- intravenously at low dose (< 2 μg g−1 body weight), and moni-
tron microscope [5]. In spite of the small pore size of the tored within 10–15 minutes [69]. LSECs with an intact ability
fenestrae, preventing their detection in conventional light and to internalize bound ligand and process it intracellularly, will
fluorescence microscopes, recent development has made it accumulate the probe and stain positively. Hence, uptake of
­possible to observe fenestrae in specially equipped light micro- FSA or other ligands that are specifically endocytosed via any
scopes, for example, structured illumination microscopy, direct of the LSEC signature receptors will yield information on both
stochastic optical reconstruction microscopy, atomic force LSEC identity and integrity. To assess the identity and purity of
microscopy, and stimulated emission depletion (STED) micros- LSECs in vitro it is important to use FSA at low concentration
copy [57–59]. Structured illumination microscopy, atomic force (< 10 μg mL−1) and short incubation times (10–30 minutes).
microscopy, and STED microscopy offer the important possibil- Unspecific staining of non‐LSECs may result if FITC‐FSA is
ity of studying the dynamics of the fenestrae opening and clos- applied at higher concentrations or longer incubation times.
ing [57–59]. In spite of these new developments, the most Uptake of acetylated LDL (AcLDL) is sometimes used to
reliable way of quantifying the number, size, and membrane identify LSECs. However, this ligand is taken up by many
organization of fenestrae is still examination in the electron types of endothelial cells and is not a totally reliable LSEC‐
microscope. Apart from using the fenestration as an identifier, specific probe [5].
LSECs may be distinguished by a set of markers that can be
detected in the light/fluorescence microscope. These markers
include antigens that are specifically expressed on the LSEC LSEC heterogeneity of marker molecules across
surface and can be marked by specific antibodies. In addition,
the liver lobule
LSEC signature endocytosis receptors may be utilized function-
ally to determine specific uptake of macromolecular ligands LSEC features such as cell size, number, and size of fenestra-
known to be taken up by these receptors. It is however, utterly tion, expression of intracellular and surface markers, expres-
important to be aware of the challenges of interpreting the sion of lectins and von Willebrand factor, endocytosis, and
results using these markers, of which some are quite specific for susceptibility to acetaminophen toxicity vary across the
LSECs, whereas others are not [60]. Moreover, there is a clear hepatic lobule [61]. Lobular zonation has not been reported for
need to standardize methods for identification and purity assess- the LSEC markers stabilin‐2 and MR in human or rodent mod-
ment of isolated LSECs, as well for identification of LSECs in els, but heterogeneic expression along sinusoids in the human
the intact liver. In a recently published update on LSECs it was liver was reported for CD32/FcγRIIb, ICAM‐1, and LYVE‐1,
advocated that cell identification should include validation of with low or absent expression of these markers in acinar zone
the identity of isolated cells assessed by both ultrastructure and 1 [70]. The authors mentioned the possibility that the low
purity based on evidence of high‐affinity endocytosis of specific staining intensity of CD32/FcγRIIb in zone 1 may be the result
ligands [61]. of systemic endothelial cells radiating from the portal tract. It
Von Willebrand factor, a marker of vascular endothelial is important to know whether positive selection using an
cells, is not normally expressed in LSECs in young individuals LSEC‐specific marker like CD32b will exclude LSECs from
[15], but may be expressed in old livers [62]. CD31 is expressed acinar zone 1, and if there are species differences. In rat,
in all endothelia, including human [63], mouse [64], and rat FcγRIIb2 expression was shown to be continuous along the
LSECs [27], but at variance from other endothelia the expres- sinusoids and similar to stabilin‐2 [71], when using the
sion of CD31 on the LSEC surface is low or varying in young FcγRIIb2‐specific SE‐1 antibody [72, 73]. In accordance with
individuals, but upregulated in liver fibrosis [65]. Hence other this finding, studies in mice on binding of IgG immune com-
LSEC markers should be chosen. Stabilin‐2, FcγRIIb2, and plexes to localize sinusoidal FcγR showed continuous pres-
LYVE‐1 are uniquely expressed by LSECs in adult liver, and ence of the probe along the sinusoidal wall, but no binding to
are to date the most reliable surface markers available for the vascular endothelium of the central and portal areas. The
immune‐identification of LSECs [2]. Other proteins that can be finding that far more probe bound to LSECs than to Kupffer
used for LSEC identification are VEGFR3 (Flt‐4) [66] and cells, combined with present knowledge that FcγRIIb2 is the
coagulation factor VIII, the latter being uniquely synthesized only FcγR present on LSEC, indicate continuous presence of
by LSEC in the liver [67]. The MR is highly expressed on FcγRIIb2 along the sinusoid [74, 75], but that the density of
LSECs but also shows a variable expression in Kupffer cells, receptors may vary in acinar zone 1.
which may be species dependent. MR expression has been The common leukocyte antigen CD45 is often used as a
reported to be absent in human and mouse Kupffer cells [22, negative selection marker when isolating human and mouse
28] and expressed at a low level in rat Kupffer cells [68] com- LSECs by fluorescent‐ or magnetic‐activated cell sorting.
pared to LSECs. However, rat LSEC is reported to be CD45 positive [76–78]
Since the functional hallmark of LSECs is their super active with high expression in zone 1, low in zone 2, and negative in
and specific endocytosis via their signature clearance receptors, zone 3 [76]. Studies in rat have also shown recruitment of
428 THE LIVER:  THE LSEC SCAVENGER FUNCTION AND ITS IMPLICATIONS

CD133+ CD45+ CD31+ bone marrow‐derived progenitor cells Adaptive immune functions
for LSECs after liver injury [79, 80]. Using CD45 as a negative
Instead of eliciting an immune response in the liver intraportally
selection criterion when isolating LSECs may therefore lead to
administered antigens results in specific tolerance [88]. Without
a biased selection of cells. Clearly, the CD45 expression pat-
a mechanism that elicits tolerogenic response in the liver, the
tern in LSECs needs to be further investigated in different
many waste molecules from the intestines and the general circu-
species.
lation would readily cause devastating immune responses in the
organ. Several studies have reported on the role of the LSECs in
the generation of hepatic immune tolerance [3, 89–92]. One
LSEC metabolism hypothesis suggests that the LSECs express MHC II molecules
Less than 20% of ATP generated by LSECs and Kupffer cells enabling them to present exogenous antigens resulting in toler-
from exogenous substrates is derived from glucose metabo- ance [92, 93]. However, some studies have failed to detect the
lism, whereas the main energy sources in these cells are oxida- presence of MHC II in LSECs [47, 90]. Another hypothesis
tion of glutamine and palmitate [81]. Studies on the metabolic advocates LSEC in cross‐presentation, where LSECs present
fate of endocytosed macromolecules following their endocyto- exogenous antigens via MHC I to CD8+ T helper cells [91].
sis by LSECs revealed that lactate and acetate are major degra-
dation products that are transported out of the cells. It has been
Role of LSECs as a sink for blood‐borne virus
speculated that the very high release of lactate from the LSECs
serves to fuel the ATP production in hepatocytes [82]. On a per A vast number of virus particles gain access to our circulation
cell basis hepatocytes oxidize 136 times and 45 times more lac- [94]. From the gut alone it is estimated that every day 31 × 109
tate than do LSECs and KCs, meaning that practically all oxi- phage particles find their way across the epithelial cell layer
dation of lactate in the liver is by the hepatocytes. This is in [95]. Yet few virus particles accumulate in the circulation.
accordance with the report that 98.8% of the mitochondria Hence there must be a very effective clearance mechanism elim-
in liver is in the hepatocytes, with 0.5% in LSECs and 0.4% in inating these virus particles from the blood. It has been hypoth-
Kupffer cells [13]. esized that blood‐borne virus is cleared by non‐phagocytic
The tension of O2, an important modulator of metabolic pro- uptake in LSECs, that are geared to endocytosis of material of
cesses, ranges from 60–70 mmHg periportally to 25–35 mmHg less than approx. 200 nm. And indeed, two recent studies have
pericentrally. Most studies with cultured LSECs have been per- shown that intravenously administered adenovirus and BK and
formed in incubators using a standard gas mixture of 5% CO2 JC polyomavirus‐like particles are efficiently taken up in the
and 95% air, which represents 150 mmHg O2 (as in the hepatic mouse liver, with LSECs as the main cellular site of uptake [69,
artery). Hence, cultivation of LSECs in a conventional incubator 96] (Figure  34.2c). If uptake of blood‐borne virus in LSECs
represents a hyperoxic environment for these cells. A study turns out to represent a general route of virus elimination, it will
­performed to determine the effect of lowering O2 tension on assign an important novel role for these cells in the antivirus
­primary cultured LSECs gave the following results [83]: (i) cell defense apparatus.
longevity in culture increased, (ii) endocytic activity was pre- Studies on how duck hepatitis B virus infects hepatocytes have
served for longer periods, (iii) proinflammatory IL‐6 was shown that virus is first endocytosed by the LSECs, before a few
reduced, (iv) anti‐inflammatory cytokine IL‐10 was increased, virus particles are rescued from the endo‐lysosomal route and trans-
(v) production of lactate doubled, (vi) generation of H2O2 was ferred to the hepatocytes [97]. Future studies on the early interaction
drastically reduced. Based on these results we and some other of blood‐borne viruses with LSECs should be carried out to gain
laboratories routinely cultivate LSECs at normoxic conditions more knowledge on how some viruses escape metabolism by
(5%, or 52 mmHg O2). LSECs and produce infection of liver and other organs.

Role of LSECs in host defense and immune Origin and renewal of LSECs


regulatory functions There appears to be two populations of LSEC progenitor cells:
bone marrow‐derived progenitors and resident or intrahepatic
Innate immune functions
LSEC progenitors [79, 80]. Normal LSEC turnover is main-
PRRs that recognize pathogen associated molecular patterns tained by the liver resident population of LSEC progenitor cells;
(PAMPs) or own molecules that may cause damage, called in addition the liver is able to recruit bone marrow‐derived cells
damage‐associated molecular patterns (DAMPs), are hall- to help replenish the LSEC population when needed [98].
marks of cells involved in innate immune functions [84]. PRRs After  induction of liver injury or partial hepatectomy LSEC
in LSECs include the MR and stabilin‐1 and ‐2, which mediate ­repopulation with CD133+ CD45+ CD31+ bone marrow‐derived
the clearance of an array of PAMPs and DAMPs from the progenitor occurred rapidly [77, 79, 80]. Hepatic VEGF regu-
­circulation. Additional PRRs expressed by LSECs are TLR2, lates the recruitment of bone marrow cells to liver [80].
3, 4, 6/2, 8, and 9 [33, 85, 86]. Of note, LSECs produce cytokines Hepatocyte growth factor (HGF) is essential for liver regenera-
TNF‐alpha, IL‐6, and IL‐1beta when challenged with TLR tion. In normal liver, the stellate cells are the main producer of
agonists. The inflammasome molecules NLRP‐1, NLRP‐3, HGF, with little expression in LSECs. However, bone marrow
and AIM2 are highly expressed in both LSECs and Kupffer LSECs repopulating the liver after liver injury are rich in HGF,
cells, with much lower, or absent expression in other types of suggesting an essential role of bone marrow‐derived cells in
liver cells [87]. liver regeneration [79].
34:  The Liver Sinusoidal Endothelial Cell: Basic Biology and Pathobiology 429

PATHOBIOLOGY of the fenestral pores and preventing larger particles from passing
through the pores. This filtering is responsible for the production
In this part of the chapter we have chosen to focus mainly on of 50% of all lymph produced in the body [108]. LSEC porosity
selected pathological conditions known to involve signature is key to understanding the lipoprotein traffic between the blood
structure and functions of LSECs, namely their fenestration and and the hepatocytes [12]. Spherical particles of triglyceride‐rich
their scavenger activity. spherical lipoproteins, generated by intestinal epithelium from
dietary lipids, are too big (100–1000 nm) to pass through the
LSEC function in inflammation fenestral pores [109], and thus stay in the circulation. The 30–80
nm chylomicron remnant particles resulting from size reduction
and liver fibrosis by lipoprotein lipase at the surface of extrahepatic endothelium,
A recent, excellent review has outlined in detail how LSECs pass through the LSEC fenestrae, and are endocytosed by the
influence the immune microenvironment within the liver and hepatocytes [110]. This explains why reduced porosity, as in liver
discusses their contribution to immune‐mediated liver diseases cirrhosis, diabetes mellitus, and old age leads to prolonged post-
and the complications of fibrosis and carcinogenesis [99]. This prandial lipoproteinemia and increased circulatory cholesterol
topic will therefore be only briefly discussed here. levels, with consequential increased risk for development of
Fibrosis is a reversible scarring process resulting from chronic atherosclerosis [110]. Due to the apparent correlation between
liver injury and characterized by excess deposition of extracellular reduced LSEC porosity and probability of developing atheroscle-
matrix (ECM) [100]. Stellate cells are the primary source of ECM rosis, effort has been invested into studying the dynamics and
both in normal and fibrotic liver, whereas LSECs and Kupffer cells molecular structure of the pores. Cholinergic agonists, vasoactive
stimulate this process through TGFβ‐activation and signaling, and intestinal peptide, glucagon, and increased intrasinusoidal blood
production of proinflammatory molecules [101, 102]. Liver fibrosis pressure dilate fenestrae, whereas alpha‐adrenergic agonists and
is associated with decreased LSEC fenestration, and appearance of serotonin give constriction [12, 111]. Such studies are also
an organized basal lamina in the space of Disse, a process called ­anticipated to yield results that can help controlling the delivery
capillarization. The capillarization of LSECs is reported to precede of drugs to hepatocytes.
the onset of liver fibrosis and suggested to act as a gatekeeper event
for the progression of fibrosis [5]. Normally differentiated LSECs Role of LSECs in liver toxicology
prevent hepatic stellate cell activation and promote reversion to qui- and drug distribution
escence, whereas capillarized LSECs do not [103].
In normal liver, LSECs are tolerogenic [91]. Following fibrotic Although hepatotoxicity is generally associated with the effect
hepatotoxic injury, LSECs develop a proinflammatory phenotype of toxic compounds on the hepatocytes, LSECs are sometimes
able to induce an immunogenic T cell phenotype and enhance cyto- the initial target of injury in a condition referred to as sinusoidal
toxic T‐cell responses [64]. During the progression of liver fibrosis obstruction syndrome (SOS, formerly hepatic veno‐occlusive
the profile of adhesion molecules expressed by LSECs changes [99, disease, VOD). SOS denotes a change of the sinusoid that may
104]. In liver inflammation the majority of leukocytes adhere in lead to hepatocyte hypoxia, with liver dysfunction and disruption
the sinusoids. Different from the general mechanism for leukocyte of the portal circulation [112]. In the two major causes of SOS,
recruitment in postcapillary venules, adhesion to the sinusoidal dietary ingestion of pyrrolizidine alkaloids and chemo‐irradiation‐
endothelium is selectin independent [105] and adherence and trans- induced injury, the LSEC is primarily affected. The chemo-
migration occur via integrin‐dependent and integrin‐independent therapeutic drugs most clearly associated with SOS at conventional
mechanisms, involving VCAM‐1, ICAM‐1, MADCAM, VAP‐1, doses are gemtuzumab ozogamicin, a­ctinomycin D, dacar-
hyaluronan, chemokine ligand, and stabilin‐mediated binding bazine, cytosine arabinoside, mithramycin, 6‐­thioguanine, and
of LSECs to leukocytes [99, 104]. urethane. Another common cause of SOS is the myeloablative
Few studies have been reported on how inflammation affects conditioning therapy used to prepare for hematopoietic stem
the LSEC endocytic function in vivo, or vice versa. Uptake of cell  transplantation for malignancy, particularly when cyclo-
hyaluronan, a ligand for stabilin‐2 [25, 106] was impaired in phosphamide is included. A hallmark of SOS is that circulatory
cirrhotic rat liver [107], whereas a study in mouse showed disruption precedes hepatocyte failure. In vitro studies show
increased capacity of fibrotic LSECs to capture ­antigen (man- that drugs and toxins implicated in SOS can be detoxified by
nosylated albumin) via the MR, compared to LSECs from nor- glutathione and are selectively more toxic to LSECs than
mal liver [64]. In accordance with this, MR expression and to  hepatocytes. Studies with dacarbazine and monocrotaline
function in LSEC was upregulated by cytokines involved in the (a pyrrolizidine alkaloid) have shown that the selective toxicity
progression of inflammation and fibrosis, that is, IL1, IL‐10, or to LSECs is related to metabolic activation of the drug.
IL‐4 in combination with IL‐13 [6, 41–43]. From this it may be Cyclophosphamide, a commonly used drug in preparative
hypothesized that the effect of chronic inflammation on LSEC ­regimens for myeloablative hematopoietic stem cell transplanta-
endocytosis varies depending on the receptor involved. tion, has among the highest incidences of SOS. The drug must be
activated by P450 in hepatocytes to be toxic to LSECs.
Altered LSEC fenestration – a proposed cause The sequence of events from administration of monocrotaline
until fulminant SOS is similar to the hepatotoxic action of aceta-
of atherosclerosis minophen [98]. In both cases initial LSEC injury precedes
Due to the fenestrae the LSEC lining functions like a semiperme- hepatocyte necrosis. The initial histological manifestation of
able membrane between the sinusoidal blood and the hepato- both is centrilobular hemorrhagic necrosis, but in SOS sinusoi-
cytes, filtering blood‐borne material of size less than the diameter dal fibrosis evolves in the late phase, whereas acetaminophen
430 THE LIVER:  ACKNOWLEDGMENT

toxicity does not have chronic histological sequelae. The fol- are cleared in the mouse mainly via the FcγRIIb2 of LSECs
lowing important biochemical features are shared between SOS [44], along with the observation that deletion of same receptor
and acetaminophen toxicity: preserving NO levels is protective, causes spontaneous autoimmunity and SLE‐like disease in mice
and inhibition of MMP‐2 and MMP‐9 reduces circulatory injury [120], point to a pivotal role of LSEC FcγRIIb2 in the disease
and subsequent parenchymal necrosis. Of note, SOS injury lasts mechanism of SLE. Moreover, the finding that scavenging of
much longer and may have higher fatality. blood‐borne DNAs is chiefly by SR‐mediated uptake in LSECs
The mechanism of hepatoxicity caused by high doses of aceta- [35], along with the fact that SLE is associated with generation
minophen has been studied in mice [8]. Soon after administration of anti‐DNA antibody, lends additional support to the hypothe-
of acetaminophen LSECs swell and begin to lose their scavenger sis that LSECs participate in the onset of SLE.
activity, as probed with a ligand for stabilin‐1/2. Already after two
hours gaps through the LSEC are formed by the destruction of LSEC function in old liver
fenestrae. These events take place before any evidence of injury to
Since early 2000, a series of studies have reported distinct age‐
parenchymal cells. The sinusoid then collapses, reducing the
related changes in the morphology and function of the hepatic
blood flow. Ethanol binged animals subsequently treated with
microvasculature of human, non‐human primates, and rodents
acetaminophen show the most severe LSEC damage.
[62]. Aging is associated with increased endothelial thickness and
The observation that serum hyaluronan levels increased in SOS
reduced LSEC porosity, and the gradual formation of an organ-
following bone marrow transplantation [113] or monocrotaline
ized basal lamina in the space of Disse [62, 121–128]. Most stud-
treatment [114] suggests either decreased clearance or increased
ies also report LSEC upregulation of von Willebrand factor [62],
production of hyaluronan. Similarly, severe hepatotoxicity due to
an antigen not normally expressed in LSECs of young livers [15].
acetaminophen ingestion correlated with increased serum hyalu-
A common finding is also the presence of highly fat‐filled stellate
ronan [115]. Moreover, SOS following bone marrow transplanta-
cells [62]. Importantly, in primary liver aging the stellate cells
tion showed significantly increased serum tissue plasminogen
remain quiescent, in contrast to liver fibrosis where these cells are
activator (tPA) [116]. Since circulating hyaluronan and tPA are
highly activated. The specific change in sinusoidal morphology
normally cleared by stabilin‐1/2 and MR mainly in LSECs, it is
due to high age has been termed age‐related pseudocapillarization
not far‐fetched to hypothesize that increased serum levels of these
[123, 124], to distinguish it from capillarization associated with
molecules mainly reflect decreased uptake due to LSEC injuries.
liver fibrosis. Postprandial hypertriglyceridemia linked to
The finding that clearance of intravenously administered marker
decreased LSEC porosity, a common condition in old age which
ligands for the LSEC stabilin‐1/2 was decreased in mice following
predisposes for the development of atherosclerosis, has been dealt
acetaminophen treatment strengthens this notion [8].
with in the section on Altered LSEC fenestration  –  a proposed
Common to many therapeutic IgG immunoglobulins is their ulti-
cause of atherosclerosis [110, 124].
mate clearance mediated by the LSEC FcγRIIb2. This pathway may
Aging also affects the LSEC scavenger function [128]. Most
result in the delivery of high concentrations of the immunoconjugate
studies show impaired endocytic capacity of LSECs at old age
and the cytotoxic agent directly to LSECs, which is why LSEC tox-
[127–130] whereas no changes were reported in [131].
icity may be observed with these compounds. Other members of the
Of note, several of the substances normally cleared by
new generation of drugs include to a large extent other types of large
LSECs, such as AGE‐modified proteins [132, 133] and oxidized
molecule compounds and nano formulations, of which many must
LDLs [31, 134] are proinflammatory, and thus have to be effec-
be administered intravascularly. Since LSECs are geared to active
tively eliminated. Reduced LSEC endocytosis as seen with
uptake of this group of compounds [2] it is not surprising that these
aging may therefore increase the risk of vascular pathologies
cells may be easily intoxicated by off‐target mechanisms, causing
within and outside the liver.
subsequent hepatotoxicity [117]. Hence it is utterly important that
The underlying mechanism of the observed changes in the sinu-
developers of large molecule drugs and nano formulations, as well
soids of old liver is not clear, but long‐term effects of gut‐derived
as regulatory authorities take into consideration the possibility that
toxins, several of them known to have negative effects on LSEC
hepatotoxicity elicited by these compounds may be initiated by inju-
fenestration, have been suggested as a driver of the age‐related
ries to LSEC due to their tendency to accumulate very high concen-
pseudocapillarization [62]. A recent study on the effect of dietary
trations of drugs via their specialized clearance receptors.
macronutrients on the liver microcirculation in ad lib fed mice
suggests that diet influences LSEC fenestration, and both diet and
Role of LSECs in the etiology of autoimmune the gut microbiome may have implications for liver function into
disorders old age [135]. Aging in liver may also be linked to the general
Reduced Fc receptor function, and the resulting formation of increase in age‐related inflammatory markers [136], which affect
soluble immune complexes may be important in the etiology of the vasculature of various organs [137], including the hepatic sinu-
autoimmune diseases such as systemic lupus erythematosus soids [127], whereas caloric restriction, which reduces oxidative
(SLE) and Sjögren’s syndrome [118]. The decreased sequestra- cell stress, prevents age‐related LSEC defenestration [138].
tion of immune complexes increases the probability of their
deposition in the kidney or other vulnerable parenchymal
organs, which may be detrimental to the host. Indeed, abnor- ACKNOWLEDGMENT
malities in the clearance of and inflammatory response to
immune complexes are characteristic features of SLE [119]. Jaione Simon‐Santamaria is highly acknowledged for help with
The finding that small soluble IgG‐antigen immune complexes the figures.
34:  The Liver Sinusoidal Endothelial Cell: Basic Biology and Pathobiology 431

REFERENCES 24. Goldstein, J.L., Ho, Y.K., Basu, S.K., and Brown, M.S. Binding site on mac-
rophages that mediates uptake and degradation of acetylated low density
lipoprotein, producing massive cholesterol deposition. Proc Natl Acad Sci U
1. Wisse, E. An ultrastructural characterization of the endothelial cell in the rat
S A, 1979;76:333–7.
liver sinusoid under normal and various experimental conditions, as a contri-
25. McCourt, P.A., Smedsrød, B.H., Melkko, J., and Johansson, S.
bution to the distinction between endothelial and Kupffer cells. J Ultrastruct
Characterization of a hyaluronan receptor on rat sinusoidal liver endothelial
Res, 1972;38(5):528–62.
cells and its functional relationship to scavenger receptors. Hepatology,
2. Sørensen, K.K., Simon‐Santamaria, J., McCuskey R.S., and Smedsrød, B.
1999;30(5):1276–86.
Liver sinusoidal endothelial cells. Compr Physiol, 2015;5(4):1751–74.
26. Schledzewski, K., Geraud, C., Arnold, B. et al. Deficiency of liver sinusoidal
3. Knolle, P.A. and Wohlleber, D. Immunological functions of liver sinusoidal
scavenger receptors stabilin‐1 and ‐2 in mice causes glomerulofibrotic
endothelial cells. Cell Mol Immunol, 2016;13(3):347–53.
nephropathy via impaired hepatic clearance of noxious blood factors. J Clin
4. Ni Y., Li J.M., Liu, M.K. et  al. Pathological process of liver sinusoidal
Invest, 2011;121(2):703–14.
endothelial cells in liver diseases. World J Gastroenterol, 2017;23(43):
27. Geraud, C., Schledzewski, K., Demory, A. et  al. Liver sinusoidal
7666–77.
­endothelium: a microenvironment‐dependent differentiation program in rat
5. DeLeve, L.D. Liver sinusoidal endothelial cells in hepatic fibrosis.
including the novel junctional protein liver endothelial differentiation‐­
Hepatology, 2015;61(5):1740–6.
associated protein‐1. Hepatology, 2010;52(1):313–26.
6. Arteta, B., Lasuen, N., Lopategi, A., Sveinbjörnsson, B., Smedsrød, B., and
28. Martens, J.H., Kzhyshkowska, J., Falkowski‐Hansen, M. et al. Differential
Vidal‐Vanaclocha, F. Colon carcinoma cell interaction with liver sinusoidal
expression of a gene signature for scavenger/lectin receptors by endothelial
endothelium inhibits organ‐specific antitumor immunity through interleu-
cells and macrophages in human lymph node sinuses, the primary sites of
kin‐1‐induced mannose receptor in mice. Hepatology, 2010;51(6):2172–82.
regional metastasis. J Pathol, 2006;208(4):574–89.
7. Zuo, Y., Ren, S., Wang, M. et al. Novel roles of liver sinusoidal endothelial
29. Campbell, F., Bos, F.L., Sieber, S. et al. Directing nanoparticle biodistribu-
cell lectin in colon carcinoma cell adhesion, migration and in‐vivo metastasis
tion through evasion and exploitation of Stab2‐dependent nanoparticle
to the liver. Gut, 2013;62(8):1169–78.
uptake. ACS Nano, 2018;12(3):2138–50.
8. McCuskey, R.S. Sinusoidal endothelial cells as an early target for hepatic
30. Hansen, B., Longati, P., Elvevold, K. et al. Stabilin‐1 and stabilin‐2 are both
toxicants. Clin Hemorheol Microcirc, 2006;34(1–2):5–10.
directed into the early endocytic pathway in hepatic sinusoidal endothelium
9. Pertoft, H. and Smedsrød, B. Separation and characterization of liver cells, in
via interactions with clathrin/AP‐2, independent of ligand binding. Exp Cell
Cell Separation: Methods and Selected Applications. 4, (eds. T.G. Pretlow
Res, 2005;303(1):160–73.
and T.P. Pretlow), Academic Press, New York, 1987, pp. 1–24.
31. Li, R., Oteiza, A., Sørensen, K.K. et al. Role of liver sinusoidal endothelial
10. Seternes, T., Sørensen, K., and Smedsrød, B. Scavenger endothelial cells
cells and stabilins in elimination of oxidized low‐density lipoproteins. Am J
of vertebrates: a nonperipheral leukocyte system for high‐capacity elimi-
Physiol Gastrointest Liver Physiol, 2011;300(1):G71–81.
nation of waste macromolecules. Proc Natl Acad Sci U S A, 2002;99(11):
32. Malerod, L., Juvet, K., Gjoen, T., and Berg, T. The expression of scavenger
7594–7.
receptor class B type I (SR‐BI) and caveolin‐1 in parenchymal and non-
11. Sørensen, K.K., McCourt, P., Berg, T. et al. The scavenger endothelial cell: a
parenchymal liver cells. Cell Tissue Res, 2002;307(2):173–80.
new player in homeostasis and immunity. Am J Physiol Regul Integr Comp
33. Martin‐Armas, M., Simon‐Santamaria, J., Pettersen, I., Moens, U.,
Physiol, 2012;303(12):R1217–30.
Smedsrød, B., and Sveinbjørnsson, B. Toll‐like receptor 9 (TLR9) is present
12. Cogger, V.C. and Le Couteur, D.G. Fenestrations in the liver sinusoidal
in murine liver sinusoidal endothelial cells (LSECs) and mediates the effect
endothelial cell, in The Liver: Biology and Pathobiology, 5th edn, (ed. I.M.
of CpG‐oligonucleotides. J Hepatol, 2006;44(5):939–46.
Arias) Wiley‐Blackwell; 2009.
34. Bijsterbosch, M.K., Manoharan, M., Rump, E.T. et al. In vivo fate of phos-
13. Blouin, A., Bolender, R.P. and Weibel, E.R. Distribution of organelles and
phorothioate antisense oligodeoxynucleotides: predominant uptake by scav-
membranes between hepatocytes and nonhepatocytes in the rat liver paren-
enger receptors on endothelial liver cells. Nucleic Acids Res,
chyma. A stereological study. J Cell Biol, 1977;72(2):441–55.
1997;25(16):3290–6.
14. Smedsrød, B., Pertoft, H., Eriksson, S., Fraser, J.R., and Laurent, T.C.
35. Hisazumi, J., Kobayashi, N., Nishikawa, M., and Takakura, Y. Significant
Studies in vitro on the uptake and degradation of sodium hyaluronate in rat
role of liver sinusoidal endothelial cells in hepatic uptake and degradation of
liver endothelial cells. Biochem J, 1984;223(3):617–26.
naked plasmid DNA after intravenous injection. Pharm Res, 2004;21(7):
15. Smedsrød, B., Pertoft, H., Gustafson, S., and Laurent, T.C. Scavenger func-
1223–8.
tions of the liver endothelial cell. Biochem J, 1990;266(2):313–27.
36. Smedsrød, B., Melkko, J., Risteli, L., and Risteli, J. Circulating C‐terminal
16. Aschoff, L. Das reticulo‐endotheliale System. Ergebnisse die innere
propeptide of type I procollagen is cleared mainly via the mannose receptor
Medizin, 1924;26:1–118.
in liver endothelial cells. Biochem J, 1990;271(2):345–50.
17. Metchnikoff, E. Uber eine Sprosspilzkrankheit der Daphnien; Beitrag zur
37. Smedsrød, B., Einarsson, M., and Pertoft, H. Tissue plasminogen activator is
Lehre uber den Kampf der Phagocyten gegen Krankheitserreger. Virchows
endocytosed by mannose and galactose receptors of rat liver cells. Thromb
Archiv fur pathologische Anatomie und Physiologie, und fur klinische
Haemost, 1988;59(3):480–4.
Medicin, Berlin. 1884;96:177–95.
38. Napper, C.E., Drickamer, K., and Taylor, M.E. Collagen binding by the man-
18. Kawai, Y., Smedsrød, B., Elvevold, K., and Wake, K. Uptake of lithium car-
nose receptor mediated through the fibronectin type II domain. Biochem J,
mine by sinusoidal endothelial and Kupffer cells of the rat liver: new insights
2006;395(3):579–86.
into the classical vital staining and the reticulo‐endothelial system. Cell
39. Malovic, I., Sørensen, K.K., Elvevold, K.H. et al. The mannose receptor on
Tissue Res, 1998;292(2):395–410.
murine liver sinusoidal endothelial cells is the main denatured collagen
19. Kjeken, R., Mousavi, S.A., Brech, A., Gjøen, T., and Berg, T. Fluid phase
clearance receptor. Hepatology, 2007;45(6):1454–61.
endocytosis of I‐125 iodixanol in rat liver parenchymal, endothelial and
40. Fiete, D.J., Beranek, M.C., and Baenziger, J.U. A cysteine‐rich domain of
Kupffer cells. Cell Tissue Res, 2001;304(2):221–30.
the “mannose” receptor mediates GalNAc‐4‐SO4 binding. Proc Natl Acad
20. Juvet, L.K., Berg, T., and Gjøen, T. The expression of endosomal rab pro-
Sci U S A, 1998;95(5):2089–93.
teins correlates with endocytic rate in rat liver cells. Hepatology,
41. Asumendi, A., Alvarez, A., Martinez, I., Smedsrod, B., and Vidal‐
1997;25(5):1204–12.
Vanaclocha, F. Hepatic sinusoidal endothelium heterogeneity with respect
21. Magnusson, S. and Berg, T. Extremely rapid endocytosis mediated by the
to  mannose receptor activity is interleukin‐1 dependent. Hepatology,
mannose receptor of sinusoidal endothelial rat liver cells. Biochem J,
1996;23(6):1521–9.
1989;257(3):651–6.
42. Martinez, I., Sveinbjørnsson, B., Vidal‐Vanaclocha, F., Asumendi, A., and
22. Elvevold, K., Simon‐Santamaria, J., Hasvold, H., McCourt P., Smedsrød, B.,
Smedsrød, B. Differential cytokine‐mediated modulation of endocytosis in
and Sorensen, K.K. Liver sinusoidal endothelial cells depend on mannose
rat liver endothelial cells. Biochem Biophys Res Commun, 1995;212(1):
receptor‐mediated recruitment of lysosomal enzymes for normal degradation
235–41.
capacity. Hepatology, 2008;48(6):2007–15.
43. Liu, Y., Gardner, C.R., Laskin, J.D., and Laskin, D.L. Classical and alterna-
23. Knook, D.L. and Sleyster, E.C. Isolated parenchymal, Kupffer and endothe-
tive activation of rat hepatic sinusoidal endothelial cells by inflammatory
lial rat liver cells characterized by their lysosomal enzyme content. Biochem
stimuli. Exp Mol Pathol, 2013;94(1):160–7.
Biophys Res Commun, 1980;96(1):250–7.
432 THE LIVER:  REFERENCES

44. Mousavi, S.A., Sporstol, M., Fladeby, C., Kjeken, R., Barois, N., and Berg, autocrine regulation. Am J Physiol Gastrointest Liver Physiol, 2004;287(4):
T. Receptor‐mediated endocytosis of immune complexes in rat liver G757–63.
­sinusoidal endothelial cells is mediated by FcgammaRIIb2. Hepatology, 66. Ding, B.S., Nolan, D.J., Butler, J.M. et al. Inductive angiocrine signals from
2007;46(3):871–84. sinusoidal endothelium are required for liver regeneration. Nature,
45. Pulford, K. and Souhami, R.L. The surface properties and antigen‐presenting 2010;468(7321):310–5.
function of hepatic non‐parenchymal cells. Clin Exp Immunol, 67. Shahani, T., Covens, K., Lavend’homme, R. et al. Human liver sinusoidal
1981;46(3):581–8. endothelial cells but not hepatocytes contain factor VIII. J Thromb Haemost,
46. Smedsrød, B., Eriksson, S., Fraser, J.R.E., Laurent, T.C., and Pertoft, H. 2014;12(1):36–42.
Properties of liver endothelial cells in primary monolayer cultures, in 68. Magnusson, S. and Berg, T. Endocytosis of ricin by rat liver cells in vivo and
Sinusoidal Liver Cells, (eds. D.L. Knook and E. Wisse), Elsevier, Amsterdam, in vitro is mainly mediated by mannose receptors on sinusoidal endothelial
1982, pp. 263–70. cells. Biochem J, 1993;291(3):749–55.
47. Smedsrød, B., Pertoft, H., Eggertsen, G., and Sundstrom, C. Functional and 69. Simon‐Santamaria, J., Rinaldo, C.H., Kardas, P. et  al. Efficient uptake of
morphological characterization of cultures of Kupffer cells and liver blood‐borne BK and JC polyomavirus‐like particles in endothelial cells of
endothelial cells prepared by means of density separation in Percoll, and liver sinusoids and renal vasa recta. PLoS One, 2014;9(11):e111762.
selective substrate adherence. Cell Tissue Res, 1985;241(3):639–49. 70. Strauss, O., Phillips, A., Ruggiero, K., Bartlett, A., and Dunbar, P.R.
48. Skogh, T., Blomhoff, R., Eskild, W., and Berg, T. Hepatic uptake of circulat- Immunofluorescence identifies distinct subsets of endothelial cells in the
ing IgG immune complexes. Immunology, 1985;55(4):585–94. human liver. Sci Rep, 2017;7:44356.
49. Kosugi, I., Muro, H., Shirasawa, H., Hirano, M., Amashita, Y., and 71. Yoshida, M., Nishikawa, Y., Omori, Y. et  al. Involvement of signaling of
Miyakawa, A. Effects of the subcutaneous injection of complete Freund’s VEGF and TGF‐beta in differentiation of sinusoidal endothelial cells during
adjuvant on Fc receptor activity and IgG immune complex uptake in liver culture of fetal rat liver cells. Cell Tissue Res, 2007;329(2):273–82.
sinusoidal endothelial cells, in Cells of the Hepatic Sinusoid, (eds. E. Wisse, 72. March, S., Hui, E.E., Underhill, G.H., Khetani, S., and Bhatia, S.N.
and D.L. Knook,), Kupffer Cell Foundation, Leiden, The Netherlands, 1993, Microenvironmental regulation of the sinusoidal endothelial cell phenotype
pp. 434–7. in vitro. Hepatology, 2009;50(3):920–8.
50. Ganesan, L.P., Kim, J., Wu, Y. et  al. FcgammaRIIb on liver sinusoidal 73. Ohmura, T., Enomoto, K., Satoh, H., Sawada, N., and Mori, M. Establishment
endothelium clears small immune complexes. J Immunol, 2012;189(10): of a novel monoclonal antibody, SE‐1, which specifically reacts with rat
4981–8. hepatic sinusoidal endothelial cells. J Histochem Cytochem, 1993;41(8):
51. Mouta Carreira, C., Nasser, S.M., di Tomaso, E. et  al. LYVE‐1 is not 1253–7.
restricted to the lymph vessels: expression in normal liver blood sinusoids 74. Muro, H., Shirasawa, H., Takahashi, Y., Maeda, M., and Nakamura, S.
and downregulation in human liver cancer and cirrhosis. Cancer Res, Localization of Fc receptors on liver sinusoidal endothelium. A histological
2001;61(22):8079–84. study by electron microscopy. Acta Pathol Jpn, 1988;38(3):291–301.
52. Pohlmann, S., Soilleux, E.J., Baribaud, F. et  al. DC‐SIGNR, a DC‐SIGN 75. Kosugi, I., Muro, H., Shirasawa, H., and Ito, I. Endocytosis of soluble IgG
homologue expressed in endothelial cells, binds to human and simian immu- immune complex and its transport to lysosomes in hepatic sinusoidal
nodeficiency viruses and activates infection in trans. Proc Natl Acad Sci U S endothelial cells. J Hepatol,. 1992;16(1–2):106–14.
A, 2001;98(5):2670–5. 76. Xie, G., Wang, L., Wang, X., and DeLeve, L.D. Isolation of periportal, mid-
53. Zhao, D., Zhang, M., Wang, M. et al. Up‐regulation of Cbl‐b is associated lobular, and centrilobular rat liver sinusoidal endothelial cells enables study
with LSECtin‐mediated inhibition of different CD4+ T‐cell subsets. of zonated drug toxicity. Am J Physiol Gastrointest Liver Physiol,
Immunobiology, 2013;218(4):602–8. 2010;299(5):G1204–10.
54. Gardner, J.P., Durso, R.J., Arrigale, R.R. et al. L‐SIGN (CD 209L) is a liver‐ 77. Harb, R., Xie, G., Lutzko, C. et al. Bone marrow progenitor cells repair rat
specific capture receptor for hepatitis C virus. Proc Natl Acad Sci U S A, hepatic sinusoidal endothelial cells after liver injury. Gastroenterology
2003;100(8):4498–503. 2009;137(2):704–12.
55. Li, Y., Hao, B., Kuai, X. et  al. C‐type lectin LSECtin interacts with DC‐ 78. McCloskey, T.W., Todaro, J.A., and Laskin, D.L. Lipopolysaccharide treat-
SIGNR and is involved in hepatitis C virus binding. Mol Cell Biochem, ment of rats alters antigen expression and oxidative metabolism in hepatic
2009;327(1–2):183–90. macrophages and endothelial cells. Hepatology, 1992;16(1):191–203.
56. Nonaka, H., Tanaka, M., Suzuki, K., and Miyajima, A. Development of 79. Wang, L., Wang, X., Xie, G., Wang, L., Hill, C.K., and DeLeve, L.D. Liver
murine hepatic sinusoidal endothelial cells characterized by the expression sinusoidal endothelial cell progenitor cells promote liver regeneration in rats.
of hyaluronan receptors. Dev Dyn, 2007;236(8):2258–67. J Clin Invest, 2012;122(4):1567–73.
57. Oie, C.I., Monkemoller, V., Hubner, W. et al. New ways of looking at very 80. Wang, L., Wang, X., Wang, L. et  al. Hepatic vascular endothelial growth
small holes – using optical nanoscopy to visualize liver sinusoidal endothe- factor regulates recruitment of rat liver sinusoidal endothelial cell progenitor
lial cell fenestrations. Nanophotonics‐Berlin, 2018;7(3):575–96. cells. Gastroenterology, 2012;143(6):1555–63.
58. Braet, F., Taatjes, D.J., and Wisse, E. Probing the unseen structure and func- 81. Spolarics, Z., Lang, C.H., Bagby, G.J., and Spitzer, J.J. Glutamine and fatty
tion of liver cells through atomic force microscopy. Semin Cell Dev Biol, acid oxidation are the main sources of energy for Kupffer and endothelial
2018;73:13–30. cells. Am J Physiol, 1991;261(2 Pt 1):G185–90.
59. Di Martino, J.D., Mascalchi, P., Legros, P. et al. STED microscopy: A sim- 82. Smedsrød, B. Cellular events in the uptake and degradation of hyaluronan.
plified method for liver sinusoidal endothelial fenestrae analysis. Biol Cell, Adv Drug Deliv Rev, 1991;7:265–78.
2018;110(7):159–68. 83. Martinez, I., Nedredal, G.I., Øie, C.I. et al. The influence of oxygen tension
60. Elvevold, K., Smedsrød, B., and Martinez, I. The liver sinusoidal endothelial on the structure and function of isolated liver sinusoidal endothelial cells.
cell: a cell type of controversial and confusing identity. Am J Physiol Comp Hepatol, 2008;7:4.
Gastrointest Liver Physiol, 2008;294(2):G391–400. 84. Matzinger, P. Tolerance, danger, and the extended family. Annu Rev Immunol,
61. DeLeve, L.D. and Maretti‐Mira, A.C. Liver sinusoidal endothelial cell: an 1994;12:991–1045.
update. Semin Liver Dis, 2017;37(4):377–87. 85. Wu, J., Meng, Z., Jiang, M. et al. Toll‐like receptor‐induced innate immune
62. Le Couteur, D.G., Warren, A., Cogger, V.C. et al. Old age and the hepatic responses in non‐parenchymal liver cells are cell type‐specific. Immunology,
sinusoid. Anat Rec (Hoboken), 2008;291(6):672–83. 2010;129(3):363–74.
63. Lalor, P.F., Lai, W.K., Curbishley, S.M., Shetty, S., and Adams, D.H. Human 86. Uhrig, A., Banafsche, R., Kremer, M. et  al. Development and functional
hepatic sinusoidal endothelial cells can be distinguished by expression of ­consequences of LPS tolerance in sinusoidal endothelial cells of the liver.
phenotypic markers related to their specialised functions in vivo. World J Leukoc Biol, 2005;77(5):626–33.
J Gastroenterol, 2006;12(34):5429–39. 87. Boaru, S.G., Borkham‐Kamphorst, E., Tihaa, L., Haas, U., and Weiskirchen,
64. Connolly, M.K., Bedrosian, A.S., Malhotra, A. et al. In hepatic fibrosis, liver R. Expression analysis of inflammasomes in experimental models of inflam-
sinusoidal endothelial cells acquire enhanced immunogenicity. J Immunol, matory and fibrotic liver disease. J Inflamm (Lond), 2012;9(1):49.
2010;185(4):2200–8. 88. Triger, D.R., Cynamon, M.H., and Wright, R. Studies on hepatic uptake of
65. DeLeve, L.D., Wang, X., Hu, L., McCuskey, M.K., and McCuskey, R.S. Rat antigen. I. Comparison of inferior vena cava and portal vein routes of immu-
liver sinusoidal endothelial cell phenotype is maintained by paracrine and nization. Immunology, 1973;25(6):941–50.
34:  The Liver Sinusoidal Endothelial Cell: Basic Biology and Pathobiology 433

89. Crispe, I.N., Giannandrea, M., Klein, I., John, B., Sampson, B., and 117. Godfrey, C., Desviat, L.R., Smedsrod, B. et  al. Delivery is key: lessons
Wuensch, S. Cellular and molecular mechanisms of liver tolerance. learnt from developing splice‐switching antisense therapies. EMBO Mol
Immunol Rev, 2006;213:101–18. Med, 2017;9(5):545–57.
90. Katz, S.C., Pillarisetty, V.G., Bleier, J.I., Shah, A.B., and DeMatteo, R.P. 118. Frank, M.M., Lawley, T.J., Hamburger, M.I., and Brown, E.J. NIH confer-
Liver sinusoidal endothelial cells are insufficient to activate T cells. ence: immunoglobulin G Fc receptor‐mediated clearance in autoimmune
J Immunol, 2004;173(1):230–5. diseases. Ann Intern Med, 1983;98(2):206–18.
91. Limmer, A., Ohl, J., Kurts, C. et  al. Efficient presentation of exogenous 119. Davies, K.A., Robson, M.G., Peters, A.M., Norsworthy, P., Nash, J.T., and
­antigen by liver endothelial cells to CD8+ T cells results in antigen‐specific Walport, M.J. Defective Fc‐dependent processing of immune complexes in
T‐cell tolerance. Nat Med, 2000;6(12):1348–54. patients with systemic lupus erythematosus. Arthritis Rheum, 2002;46(4):
92. Thomson, A.W. and Knolle, P.A. Antigen‐presenting cell function in the 1028–38.
tolerogenic liver environment. Nat Rev Immunol, 2010;10(11):753–66. 120. Ahmed, S.S., Muro, H., Nishimura, M., Kosugi, I., Tsutsi, Y., and
93. Crispe, I.N. Liver antigen‐presenting cells. J Hepatol, 2011;54(2):357–65. Shirasawa, H. Fc receptors in liver sinusoidal endothelial cells in
94. Virgin, H.W. The virome in mammalian physiology and disease. Cell, NZB/W F1 lupus mice: a histological analysis using soluble immunoglobu-
2014;157(1):142–50. lin G‐immune complexes and a monoclonal antibody (2.4G2). Hepatology,
95. Nguyen, S., Baker, K., Padman, B.S. et  al. Bacteriophage transcytosis 1995;22(1):316–24.
­provides a mechanism to cross epithelial cell layers. MBio, 2017;8(6). 121. Cogger, V.C., Warren, A., Fraser, R., Ngu, M., McLean, A.J., and Le
96. Ganesan, L.P., Mohanty, S., Kim, J., Clark, K.R., Robinson, J.M., and Couteur, D.G. Hepatic sinusoidal pseudocapillarization with aging in the
Anderson, C.L. Rapid and efficient clearance of blood‐borne virus by liver non‐human primate. Exp Gerontol, 2003;38(10):1101–7.
sinusoidal endothelium. PLoS Pathog, 2011;7(9):e1002281. 122. Hilmer, S.N., Cogger, V.C., Fraser, R., McLean, A.J., Sullivan, D., and
97. Breiner, K.M., Schaller, H., and Knolle, P.A. Endothelial cell‐mediated Le Couteur, D.G. Age‐related changes in the hepatic sinusoidal endothelium
uptake of a hepatitis B virus: a new concept of liver targeting of hepato- impede lipoprotein transfer in the rat. Hepatology, 2005;42(6):1349–54.
tropic microorganisms. Hepatology, 2001;34(4 Pt 1):803–8. 123. Le Couteur, D.G., Cogger, V.C., Markus, A.M. et al. Pseudocapillarization
98. DeLeve, L.D. Liver sinusoidal endothelial cells and liver regeneration. and associated energy limitation in the aged rat liver. Hepatology,
J Clin Invest, 2013;123(5):1861–6. 2001;33(3):537–43.
99. Shetty, S., Lalor, P.F., and Adams, D.H. Liver sinusoidal endothelial 124. Le Couteur, D.G., Fraser, R., Cogger, V.C., and McLean, A.J. Hepatic pseu-
cells – gatekeepers of hepatic immunity. Nat Rev Gastroenterol Hepatol, docapillarisation and atherosclerosis in ageing. Lancet, 2002;359(9317):
2018;15(9):555–67. 1612–5.
100. Friedman, S.L. Mechanisms of hepatic fibrogenesis. Gastroenterology, 125. McLean, A.J., Cogger, V.C., Chong, G.C. et al. Age‐related pseudocapil-
2008;134(6):1655–69. larization of the human liver. J Pathol, 2003;200(1):112–7.
101. Jarnagin, W.R., Rockey, D.C., Koteliansky, V.E., Wang, S.S., and Bissell, 126. Warren, A., Bertolino, P., Cogger, V.C., McLean, A.J., Fraser, R., and
D.M. Expression of variant fibronectins in wound healing: cellular source Le  Couteur, D.G. Hepatic pseudocapillarization in aged mice. Exp
and biological activity of the EIIIA segment in rat hepatic fibrogenesis. Gerontol, 2005;40(10):807–12.
J Cell Biol, 1994;127(6 Pt 2):2037–48. 127. Ito, Y., Sørensen, K.K., Bethea, N.W. et  al. Age‐related changes in the
102. Bilzer, M., Roggel, F., and Gerbes, A.L. Role of Kupffer cells in host hepatic microcirculation in mice. Exp Gerontol, 2007;42(8):789–97.
defense and liver disease. Liver Int, 2006;26(10):1175–86. 128. Simon‐Santamaria, J., Malovic, I., Warren, A. et al. Age‐related changes in
103. Xie, G., Wang, X., Wang, L. et al. Role of differentiation of liver sinusoidal scavenger receptor‐mediated endocytosis in rat liver sinusoidal endothelial
endothelial cells in progression and regression of hepatic fibrosis in rats. cells. J Gerontol A Biol Sci Med Sci, 2010;65(9):951–60.
Gastroenterology 2012;142(4):918–27 e6. 129. Caperna, T.J. and Garvey, J.S. Antigen handling in aging. II. The role of
104. Lee, W.Y. and Kubes, P. Leukocyte adhesion in the liver: distinct adhesion Kupffer and endothelial cells in antigen processing in Fischer 344 rats.
paradigm from other organs. J Hepatol, 2008;48(3):504–12. Mech Ageing Dev, 1982;20(3):205–21.
105. Wong, J., Johnston, B., Lee, S.S. et al. A minimal role for selectins in the 130. Heil, M.F., Dingman, A.D., and Garvey, J.S. Antigen handling in ageing.
recruitment of leukocytes into the inflamed liver microvasculature. J Clin III. Age‐related changes in antigen handling by liver parenchymal and non-
Invest, 1997;99(11):2782–90. parenchymal cells. Mech Ageing Dev, 1984;26(2–3):327–40.
106. Zhou, B., Weigel, J.A., Fauss, L., and Weigel, P.H. Identification of the 131. Brouwer, A., Barelds, R.J., and Knook, D.L. Age‐related changes in the
hyaluronan receptor for endocytosis (HARE). J Biol Chem, 2000;275(48): endocytic capacity of rat liver Kupffer and endothelial cells. Hepatology,
37733–41. 1985;5(3):362–6.
107. Tamaki, S., Ueno, T., Torimura, T., Sata, M., and Tanikawa, K. Evaluation 132. Smedsrød, B., Melkko, J., Araki, N., Sano, H., and Horiuchi, S. Advanced
of hyaluronic acid binding ability of hepatic sinusoidal endothelial cells in glycation end products are eliminated by scavenger‐receptor‐mediated
rats with liver cirrhosis. Gastroenterology, 1996;111(4):1049–57. endocytosis in hepatic sinusoidal Kupffer and endothelial cells. Biochem J,
108. Ohtani, O. and Ohtani, Y. Lymph circulation in the liver. Anat Rec, 1997;322(2):567–73.
2008;291(6):643–52. 133. Svistounov, D., Oteiza, A., Zykova, S.N. et al. Hepatic disposal of advanced
109. Naito, M. and Wisse, E. Filtration effect of endothelial fenestrations on glycation end products during maturation and aging. Exp Gerontol,
chylomicron transport in neonatal rat liver sinusoids. Cell Tissue Res, 2013;48(6):549–56.
1978;190(3):371–82. 134. Oteiza, A., Li, R., McCuskey, R.S., Smedsrød, B., and Sørensen, K.K.
110. Fraser, R., Cogger, V.C., Dobbs, B. et al. The liver sieve and atherosclero- Effects of oxidized low‐density lipoproteins on the hepatic microvascula-
sis. Pathology, 2012;44(3):181–6. ture. Am J Physiol Gastrointest Liver Physiol, 2011.
111. Braet, F. and Wisse, E. Structural and functional aspects of liver sinusoidal 135. Cogger, V.C., Mohamad, M., Solon‐Biet, S.M. et al. Dietary macronutri-
endothelial cell fenestrae: a review. Comp Hepatol, 2002;1(1):1. ents and the aging liver sinusoidal endothelial cell. Am J Physiol Heart Circ
112. DeLeve, L.D. Hepatic microvasculature in liver injury. Semin Liver Dis, Physiol, 2016;310(9):H1064–70.
2007;27(4):390–400. 136. Franceschi, C., Bonafe, M., Valensin, S. et  al. Inflamm‐aging. An
113. Fried, M.W., Duncan, A., Soroka, S. et al. Serum hyaluronic acid in patients evolutionary perspective on immunosenescence. Ann N Y Acad Sci,
­
with veno‐occlusive disease following bone marrow transplantation. Bone 2000;908:244–54.
Marrow Transplant, 2001;27(6):635–9. 137. Muller, A.M., Skrzynski, C., Nesslinger, M., Skipka, G., and Muller, K.M.
114. Copple, B.L., Roth, R.A., and Ganey, P.E. Anticoagulation and inhibition of Correlation of age with in vivo expression of endothelial markers. Exp
nitric oxide synthase influence hepatic hypoxia after monocrotaline expo- Gerontol, 2002;37(5):713–9.
sure. Toxicology, 2006;225(2–3):128–37. 138. Jamieson, H.A., Hilmer, S.N., Cogger, V.C. et al. Caloric restriction reduces
115. Williams, A.M., Langley, P.G., Osei‐Hwediah, J., Wendon, J.A., and age‐related pseudocapillarization of the hepatic sinusoid. Exp Gerontol,
Hughes, R.D. Hyaluronic acid and endothelial damage due to paracetamol‐ 2007;42(4):374–8.
induced hepatotoxicity. Liver Int, 2003;23(2):110–5. 139. Harris, E.N. and Weigel, P.H. The ligand‐binding profile of HARE: hyalu-
116. Park, Y.D., Yasui, M., Yoshimoto, T. et al. Changes in hemostatic parame- ronan and chondroitin sulfates A, C, and D bind to overlapping sites distinct
ters in hepatic veno‐occlusive disease following bone marrow transplanta- from the sites for heparin, acetylated low‐density lipoprotein, dermatan
tion. Bone Marrow Transplant, 1997;19(9):915–20. sulfate, and CS‐E. Glycobiology, 2008;18(8):638–48.
434 THE LIVER:  REFERENCES

140. Smedsrød, B., Paulsson, M., and Johansson, S. Uptake and degradation in 148. Eskild, W., Smedsrød, B., and Berg, T. Receptor mediated endocytosis of
vivo and in vitro of laminin and nidogen by rat liver cells. Biochem J, formaldehyde treated albumin, yeast invertase and chondroitin sulfate in
1989;261(1):37–42. suspensions of rat liver endothelial cells. Int J Biochem, 1986;18(7):
141. Oie, C.I., Olsen, R., Smedsrød, B., and Hansen, J.B. Liver sinusoidal 647–51.
endothelial cells are the principal site for elimination of unfractionated 149. Praaning‐van Dalen, D.P., de Leeuw, A.M., Brouwer, A., and Knook, D.L.
heparin from the circulation. Am J Physiol Gastrointest Liver Physiol, Rat liver endothelial cells have a greater capacity than Kupffer cells to
2008;294(2):G520–8. endocytose N‐acetylglucosamine‐ and mannose‐terminated glycoproteins.
142. Oynebraten, I., Hansen, B., Smedsrød, B., and Uhlin‐Hansen, L. Serglycin Hepatology, 1987;7(4):672–9.
secreted by leukocytes is efficiently eliminated from the circulation by sinusoi- 150. Smedsrød, B. and Seljelid, R. Fate of intravenously injected aminated
dal scavenger endothelial cells in the liver. J Leukoc Biol, 2000;67(2):183–8. beta(1‐‐‐‐3) polyglucose derivatized with 125I‐tyraminyl cellobiose.
143. Melkko, J., Hellevik, T., Risteli, L., Risteli, J., and Smedsrød, B. Clearance Immunopharmacol, 1991;21(3):149–58.
of NH2‐terminal propeptides of types I and III procollagen is a physiologi- 151. Blomhoff, R., Eskild, W., and Berg, T. Endocytosis of formaldehyde‐
cal function of the scavenger receptor in liver endothelial cells. J Exp Med, treated serum albumin via scavenger pathway in liver endothelial cells.
1994;179(2):405–12. Biochem J, 1984;218(1):81–6.
144. Smedsrød, B., Johansson, S., and Pertoft, H. Studies in vivo and in vitro on 152. Nagelkerke, J.F., Barto, K.P., and van Berkel, T.J. In vivo and in vitro
the uptake and degradation of soluble collagen alpha 1(I) chains in rat liver uptake and degradation of acetylated low density lipoprotein by rat liver
endothelial and Kupffer cells. Biochem J, 1985;228(2):415–24. endothelial, Kupffer, and parenchymal cells. J Biol Chem, 1983;258(20):
145. Niesen, T.E., Alpers, D.H., Stahl, P.D., and Rosenblum, J.L. Metabolism of 12221–7.
glycosylated human salivary amylase: in vivo plasma clearance by rat 153. Adachi, H. and Tsujimoto, M. FEEL‐1, a novel scavenger receptor with in
hepatic endothelial cells and in vitro receptor mediated pinocytosis by rat vitro bacteria‐binding and angiogenesis‐modulating activities. J Biol Chem,
macrophages. J Leukoc Biol, 1984;36(3):307–20. 2002;277(37):34264–70.
146. Van Berkel, T.J., De Rijke, Y.B., and Kruijt, J.K. Different fate in vivo of 154. Irjala, H., Alanen, K., Grenman, R., Heikkila, P., Joensuu, H., and Jalkanen,
oxidatively modified low density lipoprotein and acetylated low density S. Mannose receptor (MR) and common lymphatic endothelial and vascu-
lipoprotein in rats. Recognition by various scavenger receptors on Kupffer lar endothelial receptor (CLEVER)‐1 direct the binding of cancer cells to
and endothelial liver cells. J Biol Chem, 1991;266(4):2282–9. the lymph vessel endothelium. Cancer Res, 2003;63(15):4671–6.
147. Hansen, B., Svistounov, D., Olsen, R., Nagai, R., Horiuchi, S., and Smedsrød, 155. Tamura, Y., Adachi, H., Osuga, J. et al. FEEL‐1 and FEEL‐2 are endocytic
B. Advanced glycation end products impair the scavenger function of rat receptors for advanced glycation end products. J Biol Chem,. 2003;278(15):
hepatic sinusoidal endothelial cells. Diabetologia, 2002;45(10):1379–88. 12613–7.
Fenestrations in the Liver
35 Sinusoidal Endothelial Cell
Victoria C. Cogger, Nicholas J. Hunt, and David G. Le Couteur
Centre for Education and Research on Ageing, University of Sydney and Concord RG Hospital,
Sydney, Australia

INTRODUCTION neuropeptide Y [9] (Figure 35.1). Recently the development of


new imaging technologies (super‐resolution microscopy and
Liver sinusoidal endothelial cells (LSECs) occupy a strategic atomic force microscopy) has enhanced the study of fenestra-
position in the liver, separating blood in the sinusoid from the tions, particularly by revealing their structure and behavior in
extracellular space of Disse and surrounding sheets of hepato- live cells.
cytes. The cytoplasmic extensions of LSECs are very thin and
perforated with patent pores called fenestrations, which lack Structure of fenestrations
diaphragms or underlying basal lamina. Fenestrations are a
portal for the direct transfer of substrates between sinusoidal Due to the predominantly venous blood supply hepatic sinu-
blood and extracellular fluid, and also permit circulating soids bridge afferent portal venules to exiting central hepatic
immune cells, particularly T lymphocytes to interact with venules. Arterial blood, which is delivered by the hepatic
hepatocytes [1]. ­arterioles, is mixed within the sinusoid providing the limited
oxygen required by the liver. The sinusoids have an average
diameter of approximately 5–10 μm and occupy between
10–30% of the total liver volume [10]. LSECs, which repre-
HISTORICAL BACKGROUND sent 2.5% of total liver volume and 15–20% of all liver cells
[7], are a highly specialized type of endothelial cell that lines
In the late nineteenth century, it was concluded from studies the wall of the hepatic sinusoid. The cytoplasmic extensions of
investigating uptake in the liver after the injection of dyes into LSECs are very thin and perforated with fenestrations, which
the liver blood vessels, that there must be small channels linking are circular or polygonal pores approximately 50–200 nm in
the hepatic capillaries and perivascular lymphatics [2]. In 1906, diameter. The fenestrations are complete pores that connect
Herring and Simpson performed a light microscopic examina- the sinusoidal and abluminal surfaces of the LSEC with no
tion of the cat liver and noted the “sinusoidal character of the associated diaphragm. There are approximately 3–20 fenestra-
blood‐vessels and the incomplete nature of their endothelial tions per μm2 of endothelial surface and between 2–10% of the
­lining” [3]. In the 1950s, Fawcett and others used electron surface of the LSEC is covered by fenestrations, which is
microscopy to detect pores or fenestrations in the LSEC [4]. termed “porosity”. Fenestrations are scattered individually
Later, Wisse established the ultrastructure of fenestrations and across the endothelial surface or clustered into groups of ten or
their arrangement in sieve plates [5] and contemporarily, Fraser more called liver sieve plates with between 60–75% of fenes-
showed that fenestrations filter particulate substrates such as trations to be found within sieve plates. There may be a zonal
lipoproteins on the basis of size [6]. Through calcium imaging gradient of fenestration diameter within the sinusoids with
studies Arias [7] and Oda [8] went on to report that fenestrations smaller fenestrations in the periportal (zone 1) sinusoids and
are dynamic structures that can be regulated by endogenous greater porosity being found in pericentral (zone 3) sinusoids
and  exogenous agents such as serotonin, noradrenalin, and [11, 12] (Figure 35.2).

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
436 THE LIVER:  HISTORICAL BACKGROUND

(a) (b)

(c)

Kupffer cell

Hepatic sinusoid Space of Disse

Lipoprotein
Agranular reticulum

Golgi complex

Figure 35.1  Early transmission electron micrographs of the rat liver sinusoid demonstrating fenestrations in the LSEC. (a) Bennett et al. (1959)
[91]. Reproduced with permission of the American Physiological Society. J, fenestrations; G, endothelial cell; S, space of Disse; L, sinusoidal
lumen. (b) Wisse (1970) [5]. Reproduced with permission of Elsevier. →, fenestrations; end, endothelial cell; DS, space of Disse; L, sinusoidal
lumen. (c) By 1968 the concept of fenestrations had entered texts (Bloom and Fawcett 1968) [92]. Reproduced with permission of Elsevier.
The diagram shows a fenestration and lipoproteins in the space of Disse, but also the incorrect historical assumption that Kupffer cells also line
the sinusoid.

The diameter of fenestrations has a Gaussian distribution, has been huge variability in specimen preparation and statistical
which is skewed to the right by the presence of some larger methods, which has led to published conclusions about fenestra-
pores. Very small non‐perforating pores are called pits and tion morphology and response to interventions that can be
larger pores (greater than about 300 nm diameter, depending ­difficult to interpret. Standardized methods for electron micros-
upon the fixation and microscopy method) are called gaps. Gaps copy and reporting of fenestrations have been proposed but not
are known to occur as a result of cellular injury, technical issues yet widely adopted [15]. At a minimum the following informa-
related to high pressure perfusion either physiologically or tion should be provided: the methods used to measure fenestra-
experimentally, hypoxia, fixation, or pathological states. tions; the values for the number, frequency, diameter, and
Fenestrations are also found in mesh‐like, labyrinthine struc- porosity of fenestrations that were measured; and the diameter
tures reminiscent of vesiculo‐vacuolar organelles and pored limits used to define fenestrations.
domes [13]. They have been observed in numerous vertebrate Recent developments in microscopic techniques have been
species (man, rat, mouse, guinea pig, sheep, goat, rabbit, fowl, applied to fenestrations [16, 17]. The first super‐resolution light
monkey, baboon, bat, kitten, dog, turtle, goldfish) where they microscopy of fenestrations in fixed LSECs was undertaken
have similar appearances, including sieve plate formation [12]. using structured illumination microscopy (SIM) in 2010 [18],
The diameter of fenestrations is smaller than the resolution of which yielded a fenestration diameter of 123 nm, and showed
standard light microscopy therefore most morphological studies the three‐dimensional structure of sieve plates including the
in the past have utilized transmission and scanning electron observation that the width of LSEC traversed by fenestrations is
microscopy. However, preparation of tissue for electron micros- about 165 nm. Following from this time numerous other tech-
copy, particularly scanning electron microscopy generates arte- niques have emerged that have provided significant advances in
facts including tissue shrinkage, which has led to underestimation our understanding of fenestration biology (Figure  35.3a).
of the diameter of fenestrations by as much as a third. Another Atomic force microscopy (AFM), is a technique that relies on
important technical issue is that LSECs lose their fenestrations the scanning of a probe along the membrane of a fixed or live
within hours of isolation and culture, limiting the duration of cell allowing the observation of membrane topology. It was first
experiments that can be validly performed [14]. Moreover there applied to LSECs in 2001 but at the time was unable to
(a) 50 μm (b) 10 μm

(c) 1μm (d) 1μm

5 μm 1 μm
(e) (f)

Figure 35.2  Electron micrographs of LSECs and fenestrations. (a) Scanning electron microscope (EM) of vascular cast showing portal vein (PV)
branching into sinusoids (S) (b) Scanning EM of sinusoids (S) showing intercalated plates of hepatocytes (H). Fenestrations (→) are apparent in the
endothelial wall. (c) Scanning EM of sinusoidal endothelial lumen showing fenestrations clustered into sieve plates (SP). (d) Transmission EM
showing sinusoidal lumen, LSEC (E), stellate cell process (SC), and extravascular space of Disse containing microvilli. (e) Scanning EM of isolated
LSEC showing fenestrations distributed in the cytoplasm away from nucleus (Nuc) (f) Scanning EM of isolated LSEC showing fenestrations
­clustered in sieve plates and a network of fenestrations resembling a vesiculo‐vacuolar organelles (VVO). Magnification bars are shown.

1.0

0.8 Fenestration diameter

25 nm
50 nm
Sieving coefficient

0.6 75 nm
100 nm
150 nm

0.4

0.2

0.0
0 20 40 60 80 100 120 140
Lipoprotein diameter (nm)

Figure 35.3  Super resolution microscopy of the LSEC and their fenestrations. (a) Structured illumination microscopy (OMX Deltavision) of an
isolated rat LSEC. The membrane is stained with cell mask red. Fenestrations are readily visible with this method of light microscopy. (b) dSTORM
of an isolated rat LSEC, again fenestrations are clearly apparent using this method of observation. (c) and (d) 3‐dimensional structured illumination
microscopy (OMX Deltavision) reconstructed by rendering the surface of an isolated rat LSEC. The reduced thickness of the cell in regions of
fenestrations and sieves plates is evident. Acknowledgment to Mr. Hong Mao for undertaking the microscopy for images (a) and (b).
438 THE LIVER:  HISTORICAL BACKGROUND

distinguish fenestrations [19]; it has since been used success- fR 2 P


fully to study fixed and cultured living LSECs and has indicated Hagen–Poiseuille formula: J where f is the porosity of
8 l
that fenestrations have a diameter of 140–220 nm and porosity the membrane, R diameter of the pores, ΔP the pressure gradient
of about 4% [16, 20, 21]. Further, application of super‐resolu- across the membrane, η viscosity, and l the thickness of the
tion microscopy to LSECs has revealed important biophysical membrane. In conditions associated with loss of fenestrations
properties of the LSEC membrane that facilitates the formation such as liver disease and aging, there is typically a reduction
of fenestrations. SIM analysis revealed that the cell height is of porosity of 30–50%, a reduction of fenestration diameter of
“pinched” in regions of fenestrations and sieve plates, this work 10%, and an increase in endothelial thickness of 50%. Changes
led to the understanding that fenestrations are present in cell of this magnitude will be associated with a substantial reduction
membrane regions that are devoid of classic lipid raft bound in the flux of fluid and dissolved substrates across the LSEC (of
lipids and proteins [22] and the formulation of the sieve‐raft more than 50%). In addition, flow is greatest across the largest
hypothesis for the regulation of fenestrations [23]. Another pores, so permeability is sensitive to the fraction of large pores,
super‐resolution method, direct stochastic optical reconstruc- that is, the distribution of pore diameter that is present in the
tion microscopy (dSTORM) has been used in isolated LSECs LSEC (Figure 35.4).
and shown a fenestration diameter of 120 nm and identified the Fenestrations influence the hepatic uptake of lipoproteins,
close relationship between fenestrations and the cytoskeleton particularly chylomicron remnants [30, 31]. The first stage in
(Figure  35.3b) [24]. The observations derived from the new the metabolism of lipoproteins is the production of chylomi-
super resolution methods all confirm the presence of fenestra- crons which are triglyceride‐rich lipoproteins formed in the
tions in living LSECs free of fixation artefacts, and have also intestine from dietary lipids. The diameter of chylomicrons
shown that fenestrations are dynamic structures responding over range from 100–1000 nm and they are unable to pass through
minutes to various interventions such as antimycin A [20]. fenestrations; moreover most chylomicrons bypass the liver via
Additional super‐resolution techniques that involve single mol- the thoracic duct. Chylomicrons are metabolized to chylomi-
ecule localization such as PALM (photoactivated localization cron remnants by lipoprotein lipase present on the endothelium
microscopy) [25] and its related technique fPALM (fluorescent of systemic capillaries. Chylomicron remnants are smaller
photoactivated localization microscopy) [26] as well as STED particles (30–80 nm) that have acquired apolipoprotein E
­
(stimulated emission depletion) [27] offer further promise in (apoE). Remnants pass through fenestrations and are seques-
uncovering unknown aspects of LSEC fenestration biology but tered within the space of Disse for receptor‐mediated uptake by
have yet to be applied to the task [17]. This is most likely due to the low density lipoprotein receptor (LDL‐R) and lipoprotein‐
the limitations in the field caused by the absence of known receptor related protein (LRP) into hepatocytes [31]. Electron
markers or proteins that define, label, or visualize fenestrations microscopy has demonstrated that chylomicrons are only found
directly. in the sinusoid whereas remnants are also observed in the space
of Disse. There is d­ ifferential trapping by the liver of radiola-
beled lipoproteins of different sizes such that smaller particles
Physiological role of fenestrations are trapped to a greater extent than those larger than 100 nm [6].
Fenestrations facilitate the highly efficient bidirectional transfer Similar results indicating exclusion on the basis of size have
of substrates between the blood and hepatocytes, while prevent- been reported for large and small liposomes and colloidal gold
ing the access of blood cells and large particulate substrates particles of different diameters [32, 33]. Reduction in the num-
(such as platelets and chylomicrons) to the extracellular space. ber or diameter of ­ fenestrations (“defenestration”) leads to
The fenestrated endothelium acts as a filter, hence, is sometimes impaired clearance of  chylomicron remnants after meals and,
referred to as “the liver sieve” [6, 11]. Fenestrations permit the because remnants are still relatively rich in triglycerides, this is
unimpeded flow of a wide range of substrates (plasma and manifested as postprandial hypertriglyceridemia [30, 31].
­substrates within plasma, plasma proteins including albumin, Defenestration associated with old age [34]; knockout of the
smaller lipoproteins) into the extracellular space of Disse. The vascular endothelial growth factor (VEGF) receptor [35]; alco-
narrowness of hepatic endothelial cells and the lack of basal holic cirrhosis [30], poloxamer 407 [36], and knockout of the
lamina and collagen in the space of Disse ensure that any other plasmalemma vesicle‐associated protein (PLVAP) [37], is asso-
permeability barriers to the diffusion of substrates between ciated with impaired lipoprotein uptake, hypertriglyceridemia
blood and hepatocytes are minimized. Extensive investigations and/or increased ­circulating chylomicron remnants.
of substrate transfer using multiple indicator dilution methods Fenestrations have been shown to be a portal for the transfer
have established that under normal conditions, there is no bar- of several other substrates. Defenestration associated with
rier to the transfer of soluble substrates, albumin, and albumin‐ poloxamer 407 and old age has been found to reduce the hepatic
bound substrates at the LSEC; instead transfer into the space of uptake of some drugs (paracetamol, diazepam [38, 39]) and
Disse is flow‐limited [28]. insulin [40]. The effect of defenestration on insulin uptake in the
Both diameter and frequency of fenestrations determine liver is particularly significant because it provides a new mecha-
diffusive and convective transfer across the LSEC whereas
­ nism for hepatic insulin resistance and hyperinsulinemia.
permselectivity is determined only by fenestration diameter Conversely, increased fenestrations generated by partial loss of
[29]. The LSEC can be considered to be a typical low pressure platelet‐derived growth factor‐beta (PDGF‐β) is associated with
ultrafiltration system because ultrafiltration membranes are increased hepatic insulin sensitivity and lower circulating levels
­usually considered to have pores between 2–100 nm in diame- of insulin [41]. Fenestrations may be important in gene therapy
ter. In ultrafiltration, the volume flux (J) is described by the because they permit the transfecting virus, or other gene carrier,
35:  Fenestrations in the Liver Sinusoidal Endothelial Cell 439

(a) (b)

(c) (d)

Figure 35.4  The effect of fenestration diameter on the sieving coefficient (the fraction of particles able to pass through ultrafiltration pores)
according to lipoprotein diameter [80]. Reproduced with permission of Elsevier.

to gain access to the liver cells, and interestingly many of the space of Disse, finally emptying into the lymphatic vessels
strategies that improve uptake of gene therapy vectors such as around the portal vein [28]. A hydrodynamic analysis of flow in
partial hepatectomy, liver ischemia, and cyclophosphamide the hepatic sinusoids is consistent with the concept of retrograde
cause increased porosity and/or gap formation in the liver sinu- plasma flow along the space of Disse [48].
soidal endothelium [42].
As well as their role in substrate transfer, fenestrations permit
interactions between cells in the sinusoidal lumen and the cell Regulation of fenestrations
membrane of underlying hepatocytes. LSECs express many Fenestrations are dynamic structures that change in frequency
antigens important for interactions with leukocytes and lympho- and diameter in response to numerous stimuli in vitro. In vivo it
cytes and have a possible role in antigen presentation [43]. is likely that fenestrations open and close in response to various
Naïve T cells interact directly with hepatocytes by inserting stimuli such as inflammation, nutrition and fasting, circulating
filopodia through fenestrations (transendothelial hepatocyte
­ vasoactive cytokines and hormones, and local paracrine and
lymphocyte interactions, TEHLI) and this appears to be the first autocrine factors [7] (Table 35.1). The embryonic development
step in the development of immunotolerance [1]. Activated lym- of fenestrations depends upon PLVAP which is linked to
phocytes and other leukocytes accessed the liver tissue in an ­diaphragm formation [37], and VEGF [35], while in adults, the
experimental model of autoimmune hepatitis via passage maintenance of fenestrations seems to depend upon pathways
through the fenestrations and conversely defenestration almost regulating the actin cytoskeleton and its effects on lipid rafts
fully abolished any hepatitis [44]. Cytotoxic T lymphocytes rec- [22, 23]. There is often an inverse relationship between the
ognize hepatitis B viral antigens on hepatocytes by extending effects of an agent on fenestration diameter and frequency.
cytoplasmic protrusions though fenestrations and then initiate
hepatocyte killing, a process that is impaired by defenestration
Lipid rafts
associated with hepatic fibrosis [45].
Fenestrations have several other functions. Contraction of Recently a “sieve‐raft” hypothesis was proposed whereby
fenestrations might increase vascular resistance thereby influ- ­fenestrations form in non‐raft membranes when the membrane
encing hepatic blood flow and pressure [46]. For example, an stabilizing effects of the cytoskeleton and membrane rafts are
endothelin antagonist was found to dilate fenestrations and diminished. Cell membranes are heterogeneous structures com-
cause a reduction in portal perfusion pressure of about 2.5 cm posed of lipid rafts (liquid‐ordered microdomains, 10–200 nm)
H2O [47]. Fenestrations are also involved in the formation of and non‐raft, liquid‐disordered regions. Lipid rafts are regions
hepatic lymph. The space of Disse is continuous with lymphatic of cell membrane enriched in cholesterol, sphingolipids, and
vessels found in the portal triads. This morphology suggests that proteins that provide a platform for many membrane proteins.
plasma flows through the fenestrations, upstream along the They are tethered to the cytoskeleton by a variety of proteins
440 THE LIVER:  HISTORICAL BACKGROUND

Table 35.1  Fenestrations have been reported in all species studied, is the primary stimulus for VEGF production, whereas
and a very wide range of species. Some of the many reports of VEGFR‐2 is found along the entire sinusoidal endothelium
different species are shown in Table 35.1 in order to show that [55]. Differentiation of LSECs and fenestrations require both
fenestrations are widespread, and quite similar, in animals and humans
VEGF and nitric oxide (NO) while VEGF acts via both NO‐
Porosity Diameter Frequency dependent and independent pathways [56]. Genetic reduction of
Species (area %) (nm) (per μm2) Citation
VEFG activity in the liver led to loss of fenestrations associated
Rat (zone 1) 9.6 73 ± 0.13 5.7 ± 0.1 [78] with impaired uptake of chylomicron remnants [35]. VEGF is
Rat (zone 3) 28.5 94 ± 0.11 10.2 ± 0.01 [78]
Rat (zone 1) 6.0 ± 0.2 111 ± 1 9.1 ± 0.3 [10] associated with increased intracellular calcium, and has marked
Rat (zone 3) 7.9 ± 0.3 105 ± 0.2 13.3 ± 0.5 [10] effects on the cytoskeleton [57] but unlike the agents studied by
Rat 4.1 ± 2.3 73 ± 1 2.7 ± 1.1 [66] Gatmaitan and Arias [58], these intracellular changes are associ-
Rat 12.0 ± 2.1 110 ± 7 12.4 ± 3.6 [79] ated with increased – rather than reduced – fenestrations.
Mice 4.1 ± 2.2 74 ± 4 — [80]
Rabbit 5.2 ± 0.9 60 ± 5 17.3 ± 3.8 [81] The Rho family of GTPases include Rho which regulates
Rabbit 4.0 ± 1.5 69 ± 8 12.7 ± 2.5 [79] F‐actin and the cytoskeleton, Rac which regulates lamellipodia,
Chicken 3.6 ± 1.6 99 ± 15 3.9 ± 0.9 [81] and Cdc42 which controls development of filopodia. The Rho
Chicken 2.2 ± 0.6 90 ± 18 2.9 ± 0.3 [79]
Rainbow trout — 123 — [82] pathway, via its effects on the cytoskeleton, is also involved in
Goldfish — 50–200 — [83] regulation fenestrations. Inhibition of the Rho pathway by C3‐
Dog 6.7 118 ± 2 7.2 [84] transferase caused reduction of myosin light chain phosphoryla-
Sheep — 60 ± 2 — [85] tion, loss and retraction of actin filaments, and led to increased
Baboon 2.6 ± 0.2 50 ± 1 12.1 ± 0.8 [86]
Baboon 4.2 ± 0.5 58 ± 1 9.4 ± 0.9 [87] porosity and the formation of large gaps. Similarly, the Rho‐
Baboon 1.8 82 3.3 [88] associated protein kinase (ROCK) inhibitor Y‐2342 increases
Human (zone 1) 7.6 — 19 [89] fenestrations [59]. On the other hand activating Rho with
Human (zone 3) 9.1 — 23.5 [89]
Human (zone 1) 3.4 ± 0.2 170 ± 12 9.8 ± 1.8 [90]
lysophosphatidic acid increased myosin light chain phospho-
Human (zone 3) 4.0 ± 0.4 160 ± 10 11.2 ± 2.6 [90] rylation and actin filaments and led to defenestration [60].
Endothelin‐1 (ET‐1) is a potent vasoconstrictor produced by
endothelial cells and an upstream factor of Rho kinase. ET‐1
increased intracellular calcium and decreased fenestration
including ezrin, radixin, moesin, and stabilin. Between the
diameter [61]. Antagonism of the endothelin receptor, ETA‐R
rafts  lie non‐raft, lipid‐disordered regions of cell membrane.
caused a marked increase in fenestration diameter associated
Super‐resolution microscopy has shown that liver sieve plates
with gap formation [47]. ET‐1 is increased in cirrhosis which is
are distributed between lipid rafts in the non‐raft regions.
associated with increased portal pressure and defenestration [62].
Depletion of lipid rafts (with 7‐ketocholesterol) increases fenes-
trations while depletion of non‐raft membrane (with Triton X)
reduces fenestrations [22, 23]. Nutritional factors
Acute fasting increases the diameter of fenestrations while
Cytoskeleton reducing their frequency [63]. Chronic reduction of food intake
Liver sieve plates are supported by the actin cytoskeleton [49, 50] over a lifetime (caloric restriction) reversed age‐related loss of
and a variety of actin‐based structures involved with the mainte- fenestrations [64]. In long‐term feeding experiments where
nance of fenestrations have been identified such as the fenestrae‐ mice were fed different ratios of macronutrients, it was found
associated cytoskeleton ring, sieve plate associated cytoskeleton, that both fenestration porosity and frequency were inversely
fenestrae forming center, and defenestration‐associated center. related to fat intake, while diameter was inversely associated
Agents that disrupt actin such as cytochalasin B, cytochalasin with protein or carbohydrate intake [65]. Given the role of the
D, misokinolide, and latrunculin increase the numbers of fenes- liver in metabolizing nutrients it is not surprising that fenestra-
trations, usually in the order of two‐fold [51, 52]. The effect of tions respond to diet. Overall it seems that lower amounts of
cytochalasin D on fenestrations is blocked by Triton X which food intake are associated with increased fenestrations.
depletes non‐raft membrane indicating that actin cytoskeleton
acts on fenestrations via its effect on lipid rafts [22].
Pathophysiology of fenestrations
Most other agents that have been shown to influence fenestra-
tions have direct or indirect action on the cytoskeleton. The role There are numerous reports of diseases and pathological
of calcium in regulating fenestrations through effects on the ­processes that influence fenestrations, including: primary liver
cytoskeleton was first reported by Gatmaitan and Arias [53]. disease (cirrhosis, fibrosis, steatosis, hepatitis, hepatic vascular
Several agents (serotonin, metoclopramide, propranolol, indo- diseases, and the sinusoidal obstruction syndrome, vena caval
methacin, calcium ionophore) were identified that reduced the obstruction), liver toxins (acetaminophen, oxidants, bacterial
diameter of fenestrations by about 20% in rat LSECs and were toxins), systemic disease (diabetes mellitus), and other liver
associated with an increase of intracellular calcium. processes (aging, partial hepatectomy, hypoxia, high pressure,
VEGF is a potent inducer of fenestrations. In the liver, hepat- ischemia reperfusion, and transplantation) (reviewed previously
ocytes produce VEGF which acts on liver endothelial cells via [12]). These changes have not usually been diagnostic but the
the receptors: VEGFR‐1 (Flt‐1) and VEGFR‐2 (KDR/Flk‐1) of overall trends are that: (i) acute toxic injury and acute medical
which VEGFR‐2 is the most important [54]. VEGF is expressed conditions are associated with loss of endothelial integrity
more highly in the pericentral regions reflecting hypoxia, which ­characterized by gap formation and (ii) subacute and chronic
35:  Fenestrations in the Liver Sinusoidal Endothelial Cell 441

conditions have been associated with defenestration and reduced the integrity of the LSEC such as MMP inhibition and preserva-
porosity. Three important conditions are described below. tion of NO levels, totally abrogate any hepatocellular injury
[74]. The loss of LSEC morphology appears to be an initiating
Aging and pseudocapillarization step for some hepatotoxic agents.

Old age is associated with thickening and defenestration of the


LSEC, perisinusoidal fibrosis, and increased numbers of fat
engorged, non‐activated stellate cells [66, 67]. There also are CONCLUSIONS
several immunohistochemical age‐related changes including
endothelial upregulation of von Willebrand’s factor and Fenestrations are essential components of a healthy liver acting
ICAM‐1, and reduced expression of caveolin‐1. These changes as a conduit for many endogenous and exogenous substrates
occur in the absence of light microscopic evidence of liver dis- between the blood and hepatocytes. As such presence of fenes-
ease and have been termed age‐related pseudocapillarization. trations in the liver sinusoid serve as an important biological
Pseudocapillarization has been documented in mice, rats, bellwether of overall liver health. Conditions such as aging,
humans, baboons, and genetic mouse models of premature fibrosis, and cirrhosis which are all associated with diminished
aging. The age‐related loss of fenestrations is associated with or total loss of liver function are all known to be associated with
impaired transfer of lipoproteins [34], insulin [40], and some loss of fenestrations and LSEC changes. The advent of exciting
drugs [38, 39], and therefore is a potential mechanism for age‐ new microscopic techniques such as super‐resolution micros-
related changes in circulating lipoproteins, insulin sensitivity, copy has generated new knowledge about the biology of
and drug metabolism. ­fenestrations. Further understanding of proteins and processes
associated with fenestration formation, maintenance, and loss
Cirrhosis and capillarization will provide essential insights into the physiology and patho-
physiology of the liver, as well as offering potential targets
The term “capillarization” was first used in 1963 by Schaffner for  therapies directed at treating liver conditions and diseases
and Popper to describe the ultrastructural changes seen in the associated with aging such as insulin resistance and diabetes.
sinusoidal endothelium in cirrhosis, including a thickened
endothelium with underlying basement membrane and loss of
fenestrations [68]. These findings in human cirrhosis and alco-
holic liver disease have been frequently observed [69] as well as REFERENCES
animal models of cirrhosis [62, 70, 71]. Because of the associa-
tion between alcohol consumption and cirrhosis, there have 1. Warren, A., Le Couteur, D.G., Fraser, R., Bowen, D.G., McCaughan, G.W.,
and Bertolino, P. T lymphocytes interact with hepatocytes through
been several studies of the acute and subacute effects on fenes- fenestrations in murine liver sinusoidal endothelial cells. Hepatology,
­
trations. Results are inconsistent but overall, acute ethanol 2006;44(5):1182–90.
causes dilatation of fenestrations [72]. Impaired transfer of 2. Fraser, J.W. and Fraser, E.H. Preliminary note on intra and intercellular
numerous substrates (albumin, lidocaine, propranolol, prazosin, ­passages in the liver of the frog. J Anat Physiol, 1895;29:240–3.
labetalol, diltiazem, IgM) across the capillarized sinusoidal 3. Herring, P.T. and Simpson, S. On the relation of the liver cells to the blood‐
vessels and lymphatics. Proc Royal Soc Lond, 1906;78:455–97.
endothelium has been documented in cirrhotic livers [28]. 4. Fawcett, D.W. Observations on the cytology and electron microscopy of
hepatic cells. J Natl Cancer Inst, 1955;15 (5):1475–503.
Sinusoidal obstruction syndrome 5. Wisse, E. An electron microscopic study of the fenestrated endothelial lining
of rat liver sinusoids. J Ultrastruct Res, 1970;31:125–50.
The sinusoidal obstruction syndrome is the prototypic disease 6. Fraser, R., Bosanquet, A.G., and Day, W.A. Filtration of chylomicrons
of LSECs and probably the only recognized primary disease of by  the  liver may influence cholesterol metabolism and atherosclerosis.
Atherosclerosis, 1978;29:113–23.
the LSEC [73]. The two main causes of this syndrome are dietary
7. Arias, I.M. The biology of hepatic endothelial cell fenestrae. Prog Liver Dis,
pyrrolizidine alkaloids and chemo‐irradiation, especially asso- 1990;9:11–26.
ciated with bone marrow transplantation [74, 75]. The major 8. Oda, M., Tsukada, N., Komatsu, H., Kaneko, K., Nakamura, M., and
experimental model is induced by monocrotaline, a pyrrolizidine Tsuchiya, M. Electron microscopic localizations of actin, calmodulin and
alkaloid that is activated by cytochrome P450 and acts as an calcium in the hepatic sinusoidal endothelium in the rat, in Cells of the
Hepatic Sinusoid, (eds. A. Kirn, D.L. Knook, E. Wisse), Kupffer Cell
actin disruptor. After administration of monocrotaline, there is
Foundation, Leiden, 1986. pp. 511–2.
gap formation, endothelial swelling, and defenestration. This is 9. Oda, M., Han, J.Y., and Yokomori, H. Local regulators of hepatic sinusoidal
followed by extensive sinusoidal endothelial cell injury associ- microcirculation: recent advances. Clin Hemorheol Microcirc, 2000;23
ated with dissection into the space of Disse and eventually (2–4):85–94.
embolization of the sinusoidal endothelial fragments [73, 76]. 10. Wisse, E., De Zanger, R.B., Jacobs, R., and McCuskey, R.S. Scanning
­electron microscope observations on the structure of portal veins, sinusoids
Centrilobular necrosis ensues as a result of impaired perfusion and central veins in rat liver. Scan Electron Microsc, 1983(3):1441–52.
and clinically is associated with jaundice, hepatomegaly, and 11. Wisse, W., De Zanger, R.B., Charels, K., Van Der Smissen, P., and McCuskey,
ascites and a high mortality. Matrix metalloproteinases, particu- R.S. The liver sieve: considerations concerning the structure and function of
larly MMP‐9 and MMP‐2 are key mediators of the syndrome, endothelial fenestrae, the sinusoidal wall and the space of Disse. Hepatology,
but appear to influence the dehiscence of LSECs rather than 1985;5:683–92.
12. Cogger, V.C. and Le Couteur, D.G. Fenestrations in the liver sinusoidal
­fenestrations. On the other hand, the decrease in NO levels that endothelial cell, in The Liver: Biology and Pathobiology, 5th edn, (eds. I.M.
occurs early [74, 77] could directly influence fenestrations. Of Arias, A. Wolkoff, J.L. Boyer et  al.), John Wiley & Sons, Hoboken, NJ,
great significance is the observation that strategies that maintain 2009, pp. 387–404.
442 THE LIVER:  REFERENCES

13. Braet, F., Riches, J., Geerts, W., Jahn, K.A., Wisse, E., and Frederik, P. 37. Herrnberger, L., Hennig, R., Kremer, W. et  al. Formation of fenestrae in
Three‐dimensional organization of fenestrae labyrinths in liver sinusoidal murine liver sinusoids depends on plasmalemma vesicle‐associated protein
endothelial cells. Liver Int, 2009;29(4):603–13. and is required for lipoprotein passage. PLoS One, 2014;9(12):e115005.
14. Martinez, I., Nedredal, G.I., Oie, C.I. et al. The influence of oxygen tension 38. Mitchell, S.J., Huizer‐Pajkos, A., Cogger, V.C. et al. Age‐related pseudocap-
on the structure and function of isolated liver sinusoidal endothelial cells. illarization of the liver sinusoidal endothelium impairs the hepatic clearance
Comp Hepatol, 2008;7:4. of acetaminophen in rats. J Gerontol A Biol Sci Med Sci, 2011;66:400–8.
15. Cogger, V.C., O’Reilly, J.N., Warren, A., and Le Couteur, D.G. A standard- 39. Mitchell, S.J., Huizer‐Pajkos, A., Cogger, V.C. et al. The influence of old age
ized method for the analysis of liver sinusoidal endothelial cells and their and poloxamer‐407 on the hepatic disposition of diazepam in the isolated
fenestrations by scanning electron microscopy. J Viz Exp, 2015(98):e52698. perfused rat liver. Pharmacol, 2012;90(5–6):233–41.
16. Braet, F., Taatjes, D.J., and Wisse, E. Probing the unseen structure and 40. Mohamad, M., Mitchell, S.J., Wu, L.E. et  al. Ultrastructure of the liver
­function of liver cells through atomic force microscopy. Semin Cell Dev Biol, microcirculation influences hepatic and systemic insulin activity and
2018;73:13–30. provides a mechanism for age‐related insulin resistance. Aging Cell,
­
17. Øie Cristina, I., Mönkemöller, V., Hübner, W. et al. New ways of looking at 2016;15(4):706–15.
very small holes  –  using optical nanoscopy to visualize liver sinusoidal 41. Raines, S.M., Richards, O.C., Schneider, L.R. et al. Loss of PDGF‐B activity
endothelial cell fenestrations. Nanophotonics, 2018:575. increases hepatic vascular permeability and enhances insulin sensitivity.
18. Cogger, V.C., McNerney, G.P., Nyunt, T. et al. Three‐dimensional structured Am J Physiol Endocrinaol Metab, 2011;301(3):E517–26.
illumination microscopy of liver sinusoidal endothelial cell fenestrations. 42. Jacobs, F., Wisse, E., and De Geest, B. The role of liver sinusoidal cells in
J Struct Biol, 2010;171(3):382–8. hepatocyte‐directed gene transfer. Am J Pathol, 2010;176(1):14–21.
19. Braet, F., de Zanger, R., Seynaeve, C., Baekeland, M., and Wisse, E. A com- 43. Holz, L.E., Warren, A., Le Couteur, D.G., Bowen, D.G., and Bertolino, P.
parative atomic force microscopy study on living skin fibroblasts and liver CD8+ T cell tolerance following antigen recognition on hepatocytes.
endothelial cells. J Electron Microsc (Tokyo), 2001;50(4):283–90. J Autoimmun, 2010;34(1):15–22.
20. Zapotoczny, B., Szafranska, K., Owczarczyk, K., Kus, E., Chlopicki, S., and 44. Warren, A., Bertolino, P., Benseler, V., Fraser, R., McCaughan G.W., and Le
Szymonski, M. Atomic force microscopy reveals the dynamic morphology Couteur, D.G. Marked changes of the hepatic sinusoid in a transgenic mouse
of fenestrations in live liver sinusoidal endothelial cells. Sci Rep, 2017; model of acute immune‐mediated hepatitis. J Hepatol, 2007;46:239–46.
7(1):7994. 45. Guidotti, L.G., Inverso, D., Sironi, L. et al. Immunosurveillance of the liver
21. Zapotoczny, B., Owczarczyk, K., Szafranska, K., Kus, E., Chlopicki, S., and by intravascular effector CD8(+) T cells. Cell, 2015;161(3):486–500.
Szymonski, M. Morphology and force probing of primary murine liver sinu- 46. McCuskey, R.S. Morphological mechanisms for regulating blood flow
soidal endothelial cells. J Mol Recognit, 2017;30(7). through hepatic sinusoids. Liver, 2000;20(1):3–7.
22. Svistounov, D., Warren, A., McNerney, G.P. et al. The relationship between 47. Watanabe, N., Takashimizu, S., Nishizaki, Y., Kojima, S., Kagawa, T., and
fenestrations, sieve plates and rafts in liver sinusoidal endothelial cells. PLoS Matsuzaki, S. An endothelin A receptor antagonist induces dilatation of
One, 2012;7(9):e46134. ­sinusoidal endothelial fenestrae: implications for endothelin‐1 in hepatic
23. Cogger, V.C., Roessner, U., Warren, A., Fraser, R., and Le Couteur, D.G. A microcirculation. J Gastroenterol, 2007;42(9):775–82.
sieve‐raft hypothesis for the regulation of endothelial fenestrations. Comput 48. Popescu, D., Movileanu, L., Ion, S., and Flonta, M.L. Hydrodynamic effects
Struct Biotechnol J. 2013;8:e201308003. on the solute transport across endothelial pores and hepatocyte membranes.
24. Monkemoller, V., Schuttpelz, M., McCourt, P., Sorensen, K., Smedsrod, B., Phys Med Biol, 2000;45(11):N157–65.
and Huser, T. Imaging fenestrations in liver sinusoidal endothelial cells by 49. Van Der Smissen, P., Van Bossuyt, H., Charels, K., and Wisse, E. The struc-
optical localization microscopy. Phys Chem Chem Phys, 2014;16(24):12576–81. ture and function of the cytoskeleton in sinusoidal endothelial cells, in Cells
25. Betzig, E. Single molecules, cells, and super‐resolution optics (Nobel lecture). of the Hepatic Sinusoid, (eds. A. Kirn, D.L. Knook, E. Wisse),The Kupffer
Angewandte Chemie (International edn in English). 2015;54(28):8034–53. Cell Foundation, Rijswijk, 1986, pp. 517–22.
26. Hess, S.T., Girirajan, T.P., and Mason, M.D. Ultra‐high resolution imaging 50. Braet, F., De Zanger, R., Baekeland, M., Crabbe, E., Van Der Smissen, P.,
by fluorescence photoactivation localization microscopy. Biophys J, and Wisse, E. Structure and dynamics of the fenestrae‐associated cytoskele-
2006;91(11):4258–72. ton of rat liver sinusoidal endothelial cells. Hepatology, 1995;21(1):180–9.
27. Willig, K.I., Kellner, R.R., Medda, R., Hein, B., Jakobs, S., and Hell, S.W. 51. Steffan, A.M., Gendrault, J.L., and Kirn, A. Increase in the number of
Nanoscale resolution in GFP‐based microscopy. Nat Methods, fenestrae in mouse endothelial liver cells by altering the cytoskeleton with
2006;3(9):721–3. cytochalasin B. Hepatology, 1987;7(6):1230–8.
28. Le Couteur, D.G., Fraser, R., Hilmer, S., Rivory, L.P., and McLean, A.J. The 52. Braet, F., De Zanger, R., Jans, D., Spector, I., and Wisse, E. Microfilament‐
hepatic sinusoid in aging and cirrhosis ‐ Effects on hepatic substrate disposi- disrupting agent latrunculin A induces and increased number of fenestrae in
tion and drug clearance. Clin Pharmacokinet, 2005;44(2):187–200. rat liver sinusoidal endothelial cells: comparison with cytochalasin B.
29. Henriksen, J.H., Horn, T., and Christoffersen, P. The blood‐lymph barrier in Hepatology, 1996;24(3):627–35.
the liver. A review based on morphological and functional concepts of nor- 53. Gatmaitan, Z. and Arias, I.M. Hepatic endothelial cell fenestrae, in Cells of
mal and cirrhotic liver. Liver, 1984;4(4):221–32. the Hepatic Sinusoid, 4th edn, (eds. D.L. Knook and E. Wisse), KupfferCell
30. Fraser, R., Dobbs, B.R., and Rogers, G.W.T. Lipoproteins and the liver sieve: Foundation, Leiden, 1993. pp. 3–7.
the role of the fenestrated sinusoidal endothelium in lipoprotein metabolism, 54. Funyu, J., Mochida, S., Inao, M., Matsui, A., and Fujiwara, K. VEGF can act
atherosclerosis, and cirrhosis. Hepatology, 1995;21:863–74. as vascular permeability factor in the hepatic sinusoids through upregulation
31. Fraser, R., Cogger, V.C., Dobbs, B. et al. The liver sieve and atherosclerosis. of porosity of endothelial cells. Biochem Biophys Res Commun,
Pathology, 2012;44:181–6. 2001;280(2):481–5.
32. Hardonk, M.J., Harms, G., and Koudstaal, J. Zonal heterogeneity of rat 55. Cheluvappa, R., Hilmer, S.N., Kwun, S.Y., et al. The effect of old age on
hepatocytes in the in vivo uptake of 17 nm colloidal gold granules. liver oxygenation and the hepatic expression of VEGF and VEGFR2.
Histochemistry, 1985;83(5):473–7. Exp Gerontol, 2007;42(10):1012–9.
33. Romero, E.L., Morilla, M.J., Regts, J., Koning, G.A., and Scherphof, G.L. 56. Xie, G., Wang, X., Wang, L. et al. Role of differentiation of liver sinusoidal
On the mechanism of hepatic transendothelial passage of large liposomes. endothelial cells in progression and regression of hepatic fibrosis in rats.
FEBS Lett, 1999;448(1):193–6. Gastroenterology, 2012;142(4):918–27.
34. Hilmer, S.N., Cogger, V.C., Fraser, R., McLean, A.J., Sullivan, D., and Le 57. Cogger, V.C., Arias, I.M., Warren, A. et  al. The response of fenestrations,
Couteur, D.G. Age‐related changes in the hepatic sinusoidal endothelium actin, and caveolin‐1 to vascular endothelial growth factor in SK Hep1 cells.
impede lipoprotein transfer in the rat. Hepatology, 2005;42(6):1349–54. Am J Physiol Gastrointest Liver Physiol, 2008;295(1):G137–45.
35. Carpenter, B., Lin, Y., Stoll, S. et  al. VEGF is crucial for the hepatic 58. Gatmaitan, Z., Varticovski, L., Ling, L., Mikkelsen, R., Steffan, A.M., and
vascular  development required for lipoprotein uptake. Development,
­ Arias, I.M. Studies on fenestral contraction in rat liver endothelial cells in
2005;132(14):3293–303. culture. Am J Pathol, 1996;148(6):2027–41.
36. Cogger, V.C., Hilmer, S.N., Sullivan, D., Muller, M., Fraser, R., and Le 59. Venkatraman, L. and Tucker‐Kellogg, L. The CD47‐binding peptide of
Couteur, D.G. Hyperlipidemia and surfactants: the liver sieve is a link. thrombospondin‐1 induces defenestration of liver sinusoidal endothelial
Atherosclerosis, 2006;189(2):273–81. cells. Liver Int, 2013;33(9):1386–97.
35:  Fenestrations in the Liver Sinusoidal Endothelial Cell 443

60. Yokomori, H., Yoshimura, K., Funakoshi, S. et  al. Rho modulates hepatic microcirculation in rat sinusoidal obstruction syndrome. Am J Physiol,
­
sinusoidal endothelial fenestrae via regulation of the actin cytoskeleton in rat 2003;284(6):G1045–52.
endothelial cells. Lab Invest, 2004;84(7):857–64. 77. DeLeve, L.D., Wang, X., Kanel, G.C. et al. Decreased hepatic nitric oxide
61. Kamegaya, Y., Oda, M., Yokomori, H., and Ishii, H. Role of endothelin production contributes to the development of rat sinusoidal obstruction
receptors in endothelin‐1‐induced morphological changes of hepatic sinusoi- ­syndrome. Hepatology, 2003;38(4):900–8.
dal endothelial fenestrae: morphometric evaluation with scanning electron 78. Vidal‐Vanaclocha, F. and Barbera‐Guillem, E. Fenestration patterns
microscopy. Hepatol Res, 2002;22(2):89–101. in  endothelial cells of rat liver sinusoids. J Ultrastruct Res,
62. Takashimizu, S., Kojima, S., Nishizaki, Y. et  al. Effect of endothelin A 1985;90(2):115–23.
­receptor antagonist on hepatic hemodynamics in cirrhotic rats. Implications 79. Fraser, R., Heslop, V.R., Murray, F.E., and Day, W.A. Ultrastructural studies
for  endothelin‐1 in portal hypertension. Tokai J Exp Clin Med, of the portal transport of fat in chickens. Br J Exp Pathol, 1986;67:783–91.
2011;36(2):37–43. 80. Warren, A., Bertolino, P., Cogger, V.C., McLean, A.J., Fraser, R., and Le
63. O’Reilly, J.N., Cogger, V.C., Fraser, R., and Le Couteur, D.G. The effect of Couteur, D.G. Hepatic pseudocapillarization in aged mice. Exp Gerontol,
feeding and fasting on fenestrations in the liver sinusoidal endothelial cell. 2005;40(10):807–12.
Pathology, 2010;42(3):255–8. 81. Fraser, R., Day, W.A., and Fernando, N.S. Atherosclerosis and the liver sieve,
64. Jamieson, H.A., Hilmer, S.N., Cogger, V.C., Warren, A., Cheluvappa, R., and in Cells of the Hepatic Sinusoid, 1st edn, (eds. A. Kirn, D.L. Knook, and
Abernethy, D.R., et al. Caloric restriction reduces age‐related pseudocapil- E. Wisse), The Kupffer Cell Foundation, Rijswijk, 1986, pp. 317–22.
larization of the hepatic sinusoid. Exp Gerontol, 2007;42(4):374–8. 82. McCuskey, P.A., McCuskey, R.S., and Hinton, D.E. Electron microscopy of
65. Cogger, V.C., Mohamad, M., Solon‐Biet, S.M. et al. Dietary macronutrients the cells of the hepatic sinsusoids in rainbow trout, in Cells of the Hepatic
and the aging liver sinusoidal endothelial cell. Am J Physiol Heart Circ Sinusoid, 1st edn, (eds. A. Kirn, D.L. Knook, and E. Wisse), The Kupffer
Physiol, 2016;310(9):H1064–70. Cell Foundation, Rijswijk, 1986, pp. 489–94.
66. Le Couteur, D.G., Cogger, V.C., Markus, A.M.A. et al. Pseudocapillarization 83. Nopanitaya, W., Carson, J.L., Grisham, J.W., and Aghajanian, J.G. New
and associated energy limitation in the aged rat liver. Hepatology, observations on the fine structure of the liver in goldfish (Carassius auratus).
2001;33(3):537–43. Cell Tissue Res, 1979;196(2):249–61.
67. Le Couteur, D.G., Warren, A., Cogger, V.C. et al. Old age and the hepatic 84. McCuskey, R.S., McCuskey, P.A., Mitchell, D.B., Dezanger, R.B., and
sinusoid. Anat Rec, 2008;291(6):672–83. Wisse, E. Ultrastructure of the canine hepatic sinusoid, in Cells of the
68. Schaffner, F. and Popper, H. Capillarisation of hepatic sinusoids in man. Hepatic Sinusoid, 1st edn, (eds. A. Kirn, D.L. Knook, and E. Wisse),
Gastroenterology, 1963;44:239–42. The Kupffer Cell Foundation, Rijswijk, 1986, pp. 509–10.
69. Xu, B., Broome, U., Uzunel, M. et al. Capillarization of hepatic sinusoid by 85. Wright, P.L., Smith, K.F., Day, W.A., and Fraser, R. Hepatic sinusoidal
liver endothelial cell‐reactive autoantibodies in patients with cirrhosis and endothelium in sheep: an ultrastructural reinvestigation. Anat Rec,
chronic hepatitis. Am J Pathol, 2003;163(4):1275–89. 1983;206(4):385–90.
70. Fraser, R., Rogers, G.W., Bowler, L.M., Day, W.A., Dobbs, B.R., and Baxter, 86. Jamieson, H.A., Cogger, V.C., Twigg, S.M. et al. Alterations in liver sinusoi-
J.N. Defenestration and vitamin A status in a rat model of cirrhosis, in Cells dal endothelium in a baboon model of type 1 diabetes. Diabetologia,
of the Hepatic Sinusoid, 2nd edn, (eds. , E. Wisse, D.L. Knook, and R.S. 2007;50(9):1969–76.
McCuskey),Kupffer Cell Foundation, Leiden, 1991, pp. 195–8. 87. Cogger, V.C., Warren, A., Fraser, R., Ngu, M., McLean, A.J., and Le Couteur,
71. Mori, T., Okanoue, T., Sawa, Y., Hori, N., Ohta, M., and Kagawa, K. D.G. Hepatic sinusoidal pseudocapillarization with aging in the non‐human
Defenestration of the sinusoidal endothelial cell in a rat model of cirrhosis. primate. Exp Gerontol, 2003;38(10):1101–7.
Hepatology, 1993;17(5):891–7. 88. Mak, K.M. and Lieber, C.S. Alterations in endothelial fenestrations in liver
72. Jacobs, F., Wisse, E., and De Geest, B. Early effect of a single intravenous sinusoids of baboons fed alcohol: a scanning electron microscopic study.
injection of ethanol on hepatic sinusoidal endothelial fenestrae in rabbits. HeHepatologypatol, 1984;4(3):386–91.
Comp Hepatol, 2009;8:4. 89. Horn, T., Henriksen, J.H., and Christoffersen, P. The sinusoidal lining cells
73. Vreuls, C.P., Driessen, A., Olde Damink, S.W. et al. Sinusoidal obstruction in “normal” human liver. A scanning electron microscopic investigation.
syndrome (SOS): A light and electron microscopy study in human liver. Liver, 1986;6(2):98–110.
Micron, 2016;84:17–22. 90. Madarame, T., Nagaoka, T., Inomata, M. et  al. Basement membrane‐like
74. DeLeve, L.D. Hepatic microvasculature in liver injury. Semin Liver Dis, structure may play a role in the alteration of endothelial fenestration, in Cells
2007;27(4):390–400. of the Hepatic Sinusoid, 3rd edn, (eds. E. Wisse, D.L. Knook, and R.S.
75. DeLeve, L.D., Shulman, H.M., and McDonald, G.B. Toxic injury to hepatic McCuskey), Kupffer Cell Foundation, Leiden, 1991, pp. 98–101.
sinusoids: sinusoidal obstruction syndrome (veno‐occlusive disease). Semin 91. Bennett, H.S., Luft, J.H., and Hampton, J.C. Morphological classification of
Liver Dis, 2002;22(1):27–42. vertebrate blood capillaries. Am J Physiol, 1959;196:381–90.
76. DeLeve, L.D., Ito, Y., Bethea, N.W., McCuskey, M.K., Wang, X., and 92. Bloom, W. and Fawcett, D.W. A Textbook of Histology. 9th edn, Saunders,
McCuskey, R.S. Embolization by sinusoidal lining cells obstructs the Philadelphia, 1968.
Stellate Cells and Fibrosis
36 Youngmin A. Lee1 and Scott L. Friedman2
1

2
Department of Surgery, Vanderbilt University Medical Center, Nashville, TN, USA
Division of Liver Diseases, Icahn School of Medicine at Mount Sinai, New York, NY, USA

INTRODUCTION common diseases that can lead to cirrhosis include autoimmune


hepatitis, hemochromatosis, Wilson’s disease, and primary
The identification and isolation of hepatic stellate cells (HSCs) and  secondary biliary cholangitis (see Table  36.1) [1, 2].
as the main fibrogenic cells in liver, as well as techniques to Traditionally, the development of antifibrotic therapies has been
establish cultured HSC lines have enabled significant progress focused on treatment of the underlying liver disease. With the
over the past two decades in understanding key events in the arrival of curative direct acting antivirals for chronic hepatitis C
transdifferentiation process of quiescent HSCs to fibrogenic and improved therapies for chronic hepatitis B virus (HBV)
myofibroblasts (MFBs). Together with the implementation of infection, combined with preventive measures such as vaccina-
potent direct acting antivirals for the treatment of chronic viral tion for HBV, the prevalences of these diseases are expected to
hepatitis, there is now clear evidence that fibrosis and even cir- drop (Table  36.2). In contrast, lifestyle changes necessary for
rhosis can regress. Advances in clarifying HSC biology and improvement of obesity and metabolic syndrome‐associated
responses to injury have led to the promise of effective antifi- NAFLD and NASH are hard to implement; thus, development
brotic therapies in the near future. of antifibrotic therapies has significantly intensified over the
past years for NASH, with a multitude of drugs now in the pipe-
line and in clinical trials.

EPIDEMIOLOGY AND DISEASES
UNDERLYING HEPATIC FIBROSIS
HEPATIC FIBROSIS
Hepatic fibrosis due to chronic liver disease (CLD) represents a
major global health burden. It is estimated that over 840 million The liver is a highly regenerative organ; however, repeated and
people are affected by CLD, contributing to two million deaths sustained epithelial injury leads to a wound‐healing response
annually [1, 2]. Population‐based studies estimate the US preva- with excessive deposition of extracellular matrix (ECM), even-
lence of cirrhosis at 630 000 adults, of which 2/3 of patients are tually impairing the liver’s regenerative capacities and leading
unaware of having liver disease [3]. Chronic hepatitis B and C to fibrosis. Fibrillar collagens I and III, which normally make up
are the most frequent underlying liver diseases and are also about 3% of the liver by weight, are deposited by activated
globally endemic, with particularly high prevalence in Asia and HSCs, the main fibrogenic cell population in the liver. In cir-
Sub‐Saharan Africa (up to 8%) [4], while alcoholic liver disease rhotic liver, the content of fibrillar collagens I and III increases
is more prevalent in Europe and the United States (around 12%) three- to ten-fold and sulfated proteoglycans and ECM proteins
[5]. An epidemic rise in obesity in western countries has also led are deposited at the primary site of injury and within the suben-
to a massive increase in the prevalence of non‐alcoholic fatty dothelial space of Disse, at the interface of hepatocytes and
liver disease (NAFLD) (up to 46%) and non‐alcoholic steato- liver  sinusoidal endothelial cells where HSCs reside. While
hepatitis (NASH) (up to 16% in the United States). Other less the  molecular composition of the ECM in fibrotic diseases is

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
36:  Stellate Cells and Fibrosis 445

Table 36.1  Causes of fibrotic liver diseases Table 36.3  Extracellular cell matrix (ECM) components and normal
Disease localization in liver. Adapted from [105]
ECM Normal
Alcoholic liver disease
Chronic infections Collagens  
Viruses Type I Portal tract/hepatic vein
Hepatitis B Type III Portal tract, space of Disse
Hepatitis C Type IV Portal tract basement, space of Disse
Hepatitis delta Type V Portal tract, space of Disse
Bacterial Type VI Portal tract, space of Disse
Brucellosis Glycoproteins  
Parasitic Laminin Portal tract basement membranes, space of Disse
Echinococcus Fibronectin Portal tract, space of Disse
Schistosomiasis Elastin Portal tract matrix
Autoimmune hepatitis Entactin Portal tract BM
Non‐alcoholic steatohepatitis Fibrillin Portal tract matrix
Biliary cirrhosis Proteoglycans  
Primary biliary cirrhosis Heparansulfate Portal tract basement membranes
Primary sclerosing cirrhosis
Autoimmune cholangiopathy
Obstructive cholestasis The space of Disse is usually comprised of delicate strands of
Inherited metabolic diseases
Alpha antitrypsin deficiency collagen III, IV, and VI, however, during injury fibrillar colla-
Cystic fibrosis gens, laminin and fibronectin are deposited within the space of
Hereditary hemochromatosis Disse, thus forming a major solute barrier that inhibits exchange
Wilson disease
Glycogen storage diseases
between hepatocytes and plasma (see Table 36.3). Liver sinusoi-
Fructosemia dal endothelial cells, which have fenestrae allowing for solute
Galactosemia exchange, lose these pores as excess matrix is deposited in a
Lysosomal storage diseases (e.g. Fabry’s, Gaucher’s diseases) process known as “capillarization”. In murine fibrosis models
Porphyrias
Urea cycle defects ECM content is increased to up to 25–40‐fold. All these altera-
Cryptogenic cirrhosis tions eventuate in advanced liver disease with cirrhosis with
Drug induced liver injury (e.g. MTX, α‐methyl DOPA amiodarone, portal hypertension and formation of arteriovenous shunts,
isoniazid, antibiotics)
Pediatric liver disorders
which are associated with impaired synthetic and metabolic
Congenital hepatic fibrosis function of the liver.
Biliary atresia The natural course of most chronic liver disease requires
Congenital biliary cysts ­decades for advanced fibrosis to develop, and its asymptomatic
Vascular diseases
Budd‐Chiari syndrome and painless course through most of that interval allows it to
Congestive hepatopathy (cardiac cirrhosis) remain undetected for up to decades. Once advanced fibrosis
Hereditary hemorrhagic telangiectasia develops, it portends a growing increased risk for hepatocellular
Sinusoidal obstruction syndrome (formerly “veno‐occlusive disease”) carcinoma (HCC), which currently has the fastest rising cancer
incidence in the world. Moreover, HCC is already the third lead-
Table 36.2  Current and estimated future global incidences and  ing cause of cancer‐related death worldwide. Curative therapies
prevalence of liver diseases. Adapted from [2] for HCC are only possible through ablative or surgical tech-
Incidence Prevalence Current estimation Future estimation niques when lesions are smaller and fewer in number, whereas
Disease (millions) (%) (millions) 2030 (millions) advanced HCCs are rarely curable. Although cirrhosis is the
HBV 4.5–6 3.6 240 120 single most important risk factor, apart from regular screening
HCV 3–4 2.5 170 85 and treatment of the underlying liver disease, no specific chemo-
ALD 16.6 4.5 N/A 19.3
NAFLD 13.6 5–8 570 16.2 preventive therapies are yet available (Chapter 16).
NASH 2.5 less than 4 145 3.8
HBV, hepatitis B virus; HCV, hepatitis C virus; ALD, alcoholic liver disease;
NAFLD, non‐alcoholic fatty liver disease; NASH, non‐alcoholic steatohepatitis.
CIRRHOSIS
similar across fibrosis of different organs and etiologies, fibrosis
patterns vary significantly, indicating disease‐specific mecha- The term “cirrhosis” was coined by the French physician Rene
nisms of fibrogenesis. In liver, patterns of fibrosis include: T. H. Lannaëc in 1812 after the Greek word “kirrhos” for the
(i) portal–central fibrotic septae and nodule formation (e.g. auto- tawny, yellow nodules associated with this disease [6] and was
immune hepatitis, hepatitis C), (ii), perisinusoidal/pericellular adapted internationally as a synonym for end‐stage liver dis-
fibrosis (also called “chicken wire” fibrosis) which is typical for ease. Cirrhosis is characterized by fibrous septae that are present
alcoholic liver disease and non‐alcoholic steatohepatitis, throughout the liver, subdividing the organ into parenchymal
(iii) biliary cirrhosis with portal–portal fibrosis and proliferation nodules, which may vary between micronodules (less than
of bile ducts, which is mainly observed in cholestatic liver dis- 3 mm) or macronodules (3 mm to several cm). As noted above,
eases and pediatric biliary atresia, (iv) centrilobular fibrosis the fibrous septae may be present as thin bands or in different
with development of central–central vein fibrotic septa as a con- patterns such as portal‐to‐portal, central‐to‐central, or also por-
sequence of chronic hepatic venous stasis (e.g. heart failure). tal‐to‐central patterns with portal tract‐based fibrosis, frequently
446 THE LIVER:  COMMON MECHANISMS OF LIVER INJURY CONTRIBUTING TO HEPATIC FIBROSIS

producing a more severe and irregular fibrosis pattern. While it fascinating observations suggest that HSCs might retain a
was once thought that cirrhosis is an irreversible, progressive transgenerational memory of insult with decreased susceptibil-
disease, evidence from clinical trials show that regression of ity to hepatotoxins and liver fibrosis in offspring (F1 and F2) of
­cirrhosis is possible upon eradication of the underlying liver dis- rats (F0) that had been subjected to experimental liver fibrosis
ease in a broad range of diseases from viral hepatitis B and C [19]. In this seminal study, the authors detected decreased num-
[7–9], to hemochromatosis [10, 11], autoimmune hepatitis [12], bers of liver MFBs and increased expression of the antifibro-
Wilson disease [13], and also in biliary fibrosis such as biliary genic mediator PPARγ, with decreased expression of
cholangitis [14]. Depending on the underlying liver disease and profibrogenic TGFβ in livers from F1 and F2 rats. Sperm from
the degree of fibrosis it may take years for fibrotic tissue to be rats were enriched in H2A.Z and H3K27me3 indicating epige-
resorbed, with improved liver function and histology, however, netic regulation of fibrosis that requires further elucidation.
the liver’s architecture cannot be completely reconstituted and Cirrhosis is a highly dynamic process with constant meta-
incomplete septal cirrhosis may remain. Irreversible changes bolic changes of the fibrotic tissue. This has been demonstrated
which are evidence of a regressed cirrhosis are vascular aberra- in patients with chronic hepatitis C in which liver function was
tions that include venous obliterative lesions likely resulting assessed by metabolic labeling with “heavy‐water”, a non‐­
from thrombosis and to a lesser extent portal venous outflow radioactive but detectable labeling with deuterated water (2H2O),
obstruction. Loss of hepatocytes leads to collapse of liver paren- which is incorporated into amino acids and proteins. Tandem
chyma which may lead to narrowing of portal tracts and central mass spectrometry measured high turnover rates of collagen
veins, allowing for the formation of fast channels in which I and III and collagen‐associated proteins, indicating a high rate
hepatic arterioles and portal veins directly drain into hepatic of collagen remodeling even in advanced fibrosis [20].
veins, bypassing the perfusion of liver parenchyma. The forma- These findings underscore that cirrhosis is a simplified term
tion of these portal–hepatic and arterio–venous shunts further for a range of stages of advanced liver disease, with varying prog-
contributes to loss of both liver function and regenerative capac- noses and levels of reversibility. Although the term “cirrhosis”
ity due to impaired access to nutrition and key substrates that had been used for more than 200 years, an international liver
support hepatocyte homeostasis. Persistence of vascular changes pathology study group has proposed abandoning this term in
can be observed in fibrosis that has regressed, which may be a favor of a more refined diagnosis that incorporates the grade of
contributing factor to sustained portal hypertension even after activity, features of progression and regression, and risk f­ actors
improvement of liver function. Similarly, it remains unclear if for malignancy, among others [21].
the loss of fenestrae in liver sinusoidal endothelia during fibro-
sis progress is reversible upon its cessation, which is a require-
ment to restore full metabolite exchange and secretory functions
of hepatocytes.
COMMON MECHANISMS OF LIVER
Alterations in liver structure and function that persist after INJURY CONTRIBUTING TO HEPATIC
regression of fibrosis, which are not as easily detectable as vas- FIBROSIS
cular aberrations, also include changes in the key fibrogenic cell
population, the HSCs. Animal models where fibrosis reverses Oxidative and nitrosative stress, inflammation, angiogenesis,
after cessation of hepatocellular injury demonstrate that acti- and apoptosis are processes which are closely interrelated and
vated and fibrogenic HSCs may revert to an inactivated pheno- common to chronic liver diseases of different etiologies.
type, allowing for resorption and fibrosis improvement [15, 16].
However, upon renewed liver injury inactivated stellate cells
may reactivate faster and become more fibrogenic than truly
Oxidative stress
quiescent HSCs, which could explain why patients with ongo- Reactive oxygen species (ROS) and reactive nitrogen species
ing liver injury have acceleration of fibrosis progression as the (RNS) are generated during the consumption of oxygen or
stage of fibrosis advances. Evaluation of liver specimens from are derived from nitric oxide (NO) and superoxide by inducible
patients with hepatitis C‐related fibrosis, who have been fol- nitric oxide (iNOS) and NADPH oxidases, respectively.
lowed up after virological cure (SVR, sustained virological Endothelial nitric oxide, hepatocytic cytochrome P450
response) demonstrate that elimination of the underlying liver monooxygenases (e.g. CYP2E1), and NADPH oxidases (NOX)
disease leads to a regression of fibrosis in the majority of in Kupffer cells and HSCs can contribute to generation of ROS/
patients. However, a small but consistent fraction (1–14% of RNS, which react indiscriminately with all major biologically
patients, depending on the study) experience fibrosis progres- relevant macromolecules (DNA, lipids, proteins, cytoskeletal
sion with worsening of either fibrosis score or collagen propor- proteins, mitochondria) leading to DNA strand breaks, damag-
tionate area based on liver biopsies (see overview in [17]). Liver ing and inactivating metabolic pathways that provoke cell death.
biopsies from patients with chronic HCV who achieved SVR ROS/RNS inhibit parenchymal regeneration and enhance fibro-
show return to normal lobular metabolic zonation as assessed by genesis and inflammation by acting as chemotactic agents
staining for glutamine synthase (GS) and CYP2E1; however, towards HSCs and by directly activating HSCs by inducing
staining for alpha smooth muscle actin (αSMA), a marker of ­proliferation, collagen I synthesis, and TGFβ production [22].
myofibroblasts (MFB), persists in MFB before and after SVR. ROS directly activate Kupffer cells to produce proinflammatory
Moreover, a significant fraction of patients (31%) display a cytokines such as TNFα, and to amplify the inflammatory
worsening of αSMA score after SVR [18], indicating persistent response by stimulating resident and non‐resident liver cells to
remodeling of HSCs despite virologic cure. Enigmatic, but produce proinflammatory and profibrogenic mediators [23].
36:  Stellate Cells and Fibrosis 447

Other reactive oxygen species such as malondialdehyde (MDA) HSCs were first described by Kupffer in 1875 as “star‐
and 4‐hydroxy‐2,3‐nonenal (HNE) can form protein adducts shaped” cells. Their function remained unknown for the next 75
which confer antigenic properties, thus inducing antibodies that years until Ito described lipid‐containing perisinusoidal cells
could promote immune‐mediated hepatocellular injury [24] in (lipocytes or Ito cells), but it was not until Wake confirmed that
patients with NASH and alcoholic steatohepatitis (ASH). Risk these were the same cells that Kupffer had described. Quiescent
factors which amplify ROS/RNS generation include endoge- HSCs only represent 4–8% of cells in healthy liver and are char-
nous conditions such as obesity, insulin resistance, bile, and acterized by their storage of retinyl‐esters in cytoplasmic lipid
exogenous drivers such as alcohol, drugs, pollutants, environ- droplets, which allowed for their isolation and purification by
mental toxins, viruses, and ultraviolet light. An oral NOX1 and gradient centrifugation. This technique was crucial toward iden-
NOX4 inhibitor, GKT137831 attenuates liver fibrosis and apop- tifying HSCs as the main fibrogenic cells and to characterizing
tosis in rodent models [25] and is currently being evaluated for their function. It is now clear that HSCs contribute to hepatic
treatment of chronic liver disease. development, regeneration, immune responses, angiogenesis,
and are the body’s main storage site of vitamin A, of which up
to 90% are stored in the HSCs [35].
Hypoxia/angiogenesis In normal liver, HSCs maintain a quiescent, non‐prolifera-
Liver injury leads to activation of HSCs and subendothelial tive phenotype. Upon liver injury, HSCs transdifferentiate to
ECM deposition, promoting capillarization of liver sinusoids highly secretory MFBs and acquire chemotactic, inflamma-
that increases resistance to blood flow and compromises oxygen tory, and highly proliferative phenotype, losing their retinoid
delivery. Loss in size and numbers of fenestrae in liver sinusoi- droplets in the process. Activated HSCs produce most of the
dal endothelial cells impairs hepatocyte and mitochondrial matrix proteins of fibrotic liver, including fibrillar and non‐
function and enhances inflammation. Hypoxia inducible factors fibrillar collagens, components of the organized sinusoidal
HIF1α and HIF2α increase transcription and synthesis of angio- basement membrane (collagen IV, laminin and perlecan) and
genic factors including VEGF, Ang‐1, and their receptors proteoglycans [36–38]. Portal fibroblasts, which are located
VEGFR2 and Tie2, as well as PAI‐1, and adrenomedullin‐1 around the portal vein at the sinusoidal inlet, may also contrib-
(ADM1) and ADM2. In addition to VEGF’s profibrogenic func- ute to the pool of MFBs in biliary fibrosis, although HSCs are
tion toward HSCs it also promotes endothelial cell proliferation still the main source of MFBs [39–42]. Portal fibroblasts
[26–28]. PDGF, leptin (which may upregulate VEGF and likely maintain integrity of small bile ducts and interact with
Ang‐1) and HGF also have proangiogenic functions [29, 30]. cholangiocytes.
HSCs can be morphologically distinguished from portal
fibroblasts by their inclusion of vitamin A droplets and by the
Apoptosis expression of the following genes: desmin, glial fibrillary acidic
Apoptosis, or programmed cell death has been implicated in many protein (GFAP) (see Figure  36.1), lecithin retinol acyltrans-
chronic liver diseases. Apoptotic bodies are a strong profibrogenic ferase (Lrat), Hand2, vimentin, PDGFRβ, cytoglobin, and ree-
stimulus for HSCs in vivo and in vitro and is mediated in part by the lin, though there are some species‐dependent differences. For
death receptor Fas and TRAIL [31–33]. The relationship between example, human HSCs do not express desmin at all, while it is a
fibrosis and apoptosis may be bidirectional, as fibrosis may very reliable marker of mouse HSCs. Portal fibroblasts may be
stimulate apoptosis in parenchymal cells due to induction of proa- differentiated from HSCs by their expression of ecto‐ATPase
poptotic genes while stimulating survival in activated HSCs. Thus, nucleoside triphosphate diphosphohydrolase‐2 (NTPDase 2)
some therapeutic approaches that focus on inhibiting hepatocyte
apoptosis could also enhance HSC s­ urvival. Newer forms of pro-
grammed cell death may also stimulate fibrosis, for example,
necroptosis, which is activated by the necrosome consisting of
receptor‐interacting serine/threonine‐protein kinase1 (RIPK1) and
RIPK3 and mixed lineage kinase domain‐like protein (MLKL).
This pathway might be s­ pecifically important in NAFLD more
than in other types of liver injury (for review see [34]).

CELLULAR BASIS OF FIBROGENESIS


IN LIVER
Hepatic MFBs are a heterogenous population and mediate the
excessive deposition of ECM in fibrosis. They are contractile
and highly secretory cells that can be identified in part by their
expression of αSMA. Hepatic MFBs transdifferentiate from
Figure 36.1  Hepatic stellate cells in normal liver. Immunohisto­
quiescent resident mesenchymal cells, primarily stellate cells, chemistry for glial fibrillary acidic protein (GFAP, dark red) in normal
but also from periportal fibroblasts in biliary liver injuries asso- mouse liver. Note their long processes and relatively small size com-
ciated with portal fibrosis. pared to hepatocytes.
448 THE LIVER:  PATHWAYS OF HSC ACTIVATION

Table 36.4  Markers of hepatic stellate cells (HSCs) and portal factor signaling leading to perpetuation of myofibroblast phe-
fibroblasts notypes. While HSCs are evenly and ubiquitously distributed
Hepatic stellate cells Portal myofibroblasts throughout the liver lobules in normal liver, during liver injury
GFAP Ecto‐ATPase nucleoside triphosphate HSCs become most activated in zones of the most severe liver
Synaptophysin diphosphohydrolase‐2 (NTPDase2) injury [54] and inflammation. Perpetuation of HSC activation is
CRBP1 Fibulin 2 P100 characterized by events that amplify their activated phenotype
Desmin a2‐macroglobulin
Hand2 Synaptophysin
with distinctive features such as proliferation, contractility,
Reelin NCAM fibrogenesis, matrix degradation, chemotaxis, inflammatory
signaling, and loss of retinoids. These features are predomi-
nantly driven by specific sets of cytokines (Figure 36.2).
fibulin‐2, elastin, thymocyte differentiation, Gremlin 1, meso-
thelin (Msln), and cellular retinol‐binding protein‐1 (Creb1) Transforming growth factor‐β (TGFβ)
[43–45]. Other differences from HSCs might include the com- TGFβ is a key mediator in hepatic fibrogenesis [55]. TGFβ is
position of ECM each cell type secretes (see Table 36.4). ECM synthesized in a pro‐form and remains bound to latent TGFβ‐
deposited from HSC‐derived myofibroblasts is fibrillin‐1‐posi- binding protein (LTBP) and latency associated peptide (LAP) in
tive and elastin‐negative and thus differs from ECM from portal a biologically inactive state within the ECM. Latent TGFβ may
fibroblasts‐derived MFBs, which is fibrillin‐1‐ positive and be activated by matrix metalloproteinases (MMPs), plasmin,
elastin‐positive [46] though further studies are warranted but plasminogen activators, avβ6 integrins, thrombospondins,
currently limited due to fewer available tools for portal fibro- Kupffer cells, and other cell populations [56]. Mechanisms of
blasts compared to HSCs. Single cell sequencing of mesenchy- activating TGFβ may differ across different tissues and even in
mal cell populations in liver will be required to definitively different types of liver disease. In HSCs, TGFβ is profibrogenic
characterize fibrogenic cell types in liver. Immortalized rat por- through its cell surface receptors, leading to downstream signal-
tal MFB have been recently described [47], but there are no ing via Smad3 that induces collagen I, collagen, III and fibronec-
mouse reporter lines developed to study portal fibroblasts in tin expression and inducing chemotaxis. TGFβ also induces
vivo. In addition, isolation techniques and markers have species‐ mitogen‐activated proteinase (MAPK) pathways (ERK, p38
specific differences for portal MFBs between rat and mouse MAPK, and JNK) independently of Smads [57, 58].
(personal observation). Because systemic inhibition of TGFβ promotes tumorigene-
The development of tools and murine models to study HSCs in sis and inflammation, liver‐ or cell type‐specific inhibition of
vivo has greatly enhanced the characterization of HSC biology, TGFβ would be ideal, since systemic neutralization of TGFβ
including the availability of immortalized human (LX2, TWNT‐4) might decrease fibrosis but could also elevate the risk of hepato-
and murine HSC lines [48, 49], transgenic reporters driving Cre cellular or biliary cancers [59, 60].
recombinase for stellate cell‐specific conditional deletion in liver
(lecithin retinol acetyltransferase, Lrat) [40], and murine models
to deplete HSCs in vivo [50, 51]. The growing capacity to com- Platelet‐derived growth factor B (PDGF‐B)
pare large datasets has enabled comprehensive comparisons of PDGF‐B is the most potent mitogen and chemoattractant for
gene expression between quiescent and activated HSCs, and HSCs. PDGF signals through the tyrosine kinase receptor β‐
between HSCs and other liver cell populations, which have PDGFR, which is rapidly upregulated during initiation in HSCs
helped identify HSC specific transcripts and potential therapeutic in humans and in rodent studies [61, 62], thereby amplifying
targets [52]. Fluorescent‐­tagging of HSCs by bHLH transcription PDGF-B signaling in HSCs. Major sources of PDGF-B are
factor (Tg:Hand2‐EGFP) in zebrafish allows for the characteriza- HSCs themselves, macrophages, and platelets [61, 63]. In murine
tion of HSCs during liver development in this species, and has models of liver fibrosis, conditional deletion of β-PDGFR on HSCs
further identified LSECs as pivotal to enable the migration and led to decreased liver injury and fibrosis [64]. In preclinical
localization of HSCs into the liver [53]. ­studies, HSC specific delivery of erlotinib via nanoparticles that
bind to β‐PDGFR [65] has been proposed as a novel approach to
cancer chemoprevention based on animal studies. Additional
HSC mitogens include VEGF, basic fibrocyte growth factor
PATHWAYS OF HSC ACTIVATION (bFGF), thrombin, TGFα, and keratinocyte growth factor [66].

The process of HSC activation from quiescent vitamin A‐rich


Connective tissue growth factor (CTGF/CCN2)
cells into fibrogenic MFBs has been well characterized and has
yielded a plethora of antifibrotic targets being pursued in clini- CTGF is a promising antifibrotic target. CTGF is a profibro-
cal trials. HSC activation conceptually consists of two phases, genic “CCN” protein which is mainly expressed by HSCs in
initiation and perpetuation which could be followed by either injured liver, thereby promoting fibrogenesis, adhesion, migra-
apoptosis, senescence, or reversion to quiescence of HSCs if the tion, and cell survival [67]. Inhibition of CTGF by human anti‐
underlying insult has been resolved, thus contributing to resolu- CTGF antibodies (FG‐3019) are in clinical trials for pulmonary
tion of fibrosis. fibrosis but the trial patients with hepatitis B was halted
During initiation, quiescent HSCs become responsive to because patients on antiviral therapies had significant fibrosis
stimuli of the injured liver and triggers of their environment by improvement without the antibody therapy (ClinicalTrials.gov
upregulating growth factor receptors and modulation of growth ID #NCT01217632).
36:  Stellate Cells and Fibrosis 449

Figure 36.2  Functions and features of hepatic stellate cells (HSCs) in healthy and diseased liver. Quiescent hepatic stellate cells activate upon
liver injury, entering from an initiation phase into perpetuation. HSCs transdifferentiate to myofibroblasts with specific phenotype changes includ-
ing proliferation, contractility, fibrogenesis, altered matrix degradation, chemotaxis, and inflammatory signaling. From [106]. Reproduced with
permission of BMJ Publishing Group Ltd.

Endothelin‐1 (ET1) preclinical models of liver injury (NASH and TAA‐induced


liver fibrosis) [73]. Phase two clinical trials in patients with
ET1 is a major regulator of HSC contractility. HSCs are located
NASH and fibrosis have demonstrated significant antifibrotic
in collagenous bands between nodules of fibrotic liver and may
activity compared to placebo treated patients after one year of
constrict portal blood flow in individual sinusoids thus contrib-
therapy [74], however the benefit did not extend to two years.
uting to portal hypertension [68].
HSCs may also function as antigen presenting cells [75].
Vascular endothelial growth factor (VEGF) Tissue inhibitor of metalloproteinases (TIMP)
VEGF is released from liver sinusoidal endothelial cells and and metalloproteinases (MMPs)
HSCs in injured liver. It is a potent mitogen and stimulates HSC Activated HSCs produce tissue inhibitors of metalloproteinases
migration and collagen production. While angiogenesis is a (TIMPs) TIMP1 and TIMP2, of which TIMP1 is antiapoptotic
pathogenic process in advanced liver disease, it may also be a toward HSCs in part through induction of bcl‐2, thus promoting the
requirement for liver regeneration [69, 70]. Thus, efforts to survival of fibrogenic cells [76]. MMP2 and MMP9, which are also
block angiogenesis in liver disease should ideally be restricted produced by HSCs, allow for disruption of normal hepatic matrix
to tumor related angiogenesis and not angiogenesis that accom- that could hasten its replacement by fibrotic matrix [77–79].
panies and supports regeneration.

Immunoregulation
METABOLIC REGULATION OF HSCS
HSCs produce several immunoregulatory chemokines including FUNCTION AND ACTIVATION
macrophage inflammatory protein‐2 (MIP2) [71, 72] MCP‐1,
CCl2, RANTES, and CCR5 which activate and recruit mac-
rophages and other immune cells. The CC chemokines RANTES
Adipokines
(CCL5), MPC‐1, and CCL21 may also directly promote prolif- Adipokine signaling has emerged as an important regulator of
eration and migration by HSCs. The dual CCR2/CCR5 HSC activation in the context of metabolic syndrome. Adipokines
chemokine receptor antagonist Cenicriviroc is antifibrotic in including leptin and resistin secreted from adipose tissue
450 THE LIVER:  EXTRACELLULAR EVENTS THAT LEAD TO ACTIVATION OF HSCS

promote profibrogenic effects on liver. Leptin has been clearly Hepatocytes


associated with HSC fibrogenesis [80–82] increasing expression
of αSMA, collagen 1α1, and TGFβ in HSCs as well as proin- Disease‐specific mechanisms may lead to hepatocyte injury and
flammatory and proangiogenic cytokines [83–85]. In contrast, release of ROS, Hedgehog ligands, nucleotides, lipid peroxides,
adiponectin may have protective effects by decreasing fatty acid and cytokines which contribute to HSC activation. Dead or
oxidation and inhibiting hepatic gluconeogenesis. In human dying hepatocytes release damage‐associated molecular pat-
and murine models, lower adiponectin levels have been associ- terns (DAMPS) which amplify an inflammatory response. IL‐33
ated with increased severity of liver inflammation [83, 84] [87] activates resident innate lymphoid cells (ILC2) to release
IL13, acting through type II IL‐4 receptor dependent signaling
via signal transducer and activator of transcription 6 (STAT6) to
Autophagy activate HSCs [88] and promote fibrosis.
HSCs are highly dependent on autophagy to generate energy sub-
strates that fuel cellular activation. This occurs by autophagy‐ Kupffer cells and macrophages
mediated cleavage of retinyl esters within cytoplasmic droplets to
generate free fatty acids [86]. HSC‐specific genetic deletion of Tissue resident macrophages, Kupffer cells, produce ROS and
autophagy related protein 7 (Atg7) in murine models of liver fibro- TGFβ which not only promote transdifferentiation of HSCs into
sis and in cultured HSCs both lead to reduced fibrogenesis [86]. MFBs, but TGFβ is also a potent profibrogenic cytokine, stimu-
lating collagen I production in HSCs via the Smad transcrip-
tional regulators and C/EBPβ‐dependent pathways [89]. TGFβ
also induces synthesis of TIMP1, which inhibits apoptosis in
EXTRACELLULAR EVENTS THAT LEAD HSCs. Kupffer cells release large amounts of TNFα, which acti-
TO ACTIVATION OF HSCS vates HSCs to release other cytokines including NOX, PDGF,
MCP1, and PDGF. In mouse, monocytes can be differentiated
HSCs have complex interactions with all liver cell types and into at least two subgroups: M2A/Gr1hi likely promote profi-
may be activated or inhibited by them in an etiology‐specific or brotic effects via TGFβ, PDGF, FGF2, Galectin 3, CCL2,
generalized manner (see Figure 36.3). CCL18, and IGFBP5, whereas Ly6‐Clo macrophages (CD11bhi/

Figure 36.3  Extracellular stimuli of hepatic stellate cell (HSC) activation. HSCs may be inhibited or activated by hepatic cells including hepato-
cytes, liver sinusoidal endothelial cells, macrophages, and natural killer (NK) and natural killer T cells (NKT) cells through the release of cytokines,
growth factors, or signaling molecules. Green lines indicate promotion and red lines inhibition of HSC activation. From [106]. Reproduced with
permission of BMJ Publishing Group Ltd.
36:  Stellate Cells and Fibrosis 451

F4/80int/Ly6Clo) promote fibrolytic effects via release of MMP9 Platelets


and MMP12; the human counterparts to these murine cell sub-
sets have not been definitively characterized. Platelets release growth factors such as PDGF and TGFβ, which
have potent mitogenic and profibrogenic properties, respec-
tively, toward HSCs. Platelet‐derived chemokine ligand 4
Liver sinusoidal endothelial cells (LSEC) (CXCL4) stimulates proliferation, chemotaxis, and chemokine
LSECs are a key source of intrahepatic fibronectin A isoform, expression of HSCs in vitro. CXCL4 knockout mice have sig-
which is crucial for initiation of HSCs and mitogenic and fibro- nificantly reduced liver damage upon injury. Procoagulant fac-
genic stimulation by TGFβ [90, 91] enhancing HSC survival tors (e.g. thrombin, factor V) are strongly positively associated
[92]. A splice variant of fibronectin, fibronectin EDA produced with fibrosis progression [95, 96].
by LSECs, is primarily expressed during development and in
response to injury. While hepatocytes and HSCs may also Natural killer cells (NK)
express fibronectin, LSECs are early responders with fibronectin
release [91, 93]. Depending on the context, LSECs have diver- NK cells are an important part of the liver’s innate immune sys-
gent functions and may promote fibrosis by fibroblast growth tem. NK cells have context‐specific, divergent profibrotic and
factor receptor 1 (FGFR1) and CXCR4 or proregenerative path- antifibrogenic properties. NK cells may have inhibitory effects
ways which are regulated by CXCR7. These observations have on liver fibrosis by killing activated HSCs in a TRAIL−,
been tested in mouse models: deletion of CXCR7 in LSECS NKG2D− and granzyme‐dependent manner. Following release
impairs liver regeneration while deletion of either FGFR1 in of IFNγ, NK cells can induce cell arrest and apoptosis in HSC
LSECS or CXCR4 promotes proregenerative pathways [94]. via STAT1‐signaling and help clear senescent activated HSCs [97].

Figure 36.4  Strategies for antifibrotic approaches. (1) Control or cure of the underlying disease is still the most effective antifibrotic therapy. (2)
Inhibition of hepatic stellate cell (HSC) activation by modulation of receptor‐ligand interactions by either established or experimental drugs can
attenuate fibrosis development. (3) Inhibition of the most potent profibrogenic pathways in HSCs, for example, by blocking CTGF or inhibiting
activation of latent TGFβ. (4) Promoting resolution of fibrosis by enhancing apoptosis of HSCs or by enhancing matrix degradation by TIMP1
antagonists and preventing cross‐linking of collagen by LOX2 inhibition. FXR, farnesoid‐X‐receptor; PPAR, peroxisome proliferator‐activated
receptor; UDCA, ursodeoxycholic acid; SVR, sustained virological response; CB1, cannabinoid receptor type 1; ARB, angiotensin II receptor
blocker; ET‐1, endothelin 1; TGFβ, transforming growth factor β; CTGF, connective tissue growth factor; mAb, monoclonal antibody; NF‐κB,
nuclear factor kappa‐light‐chain‐enhancer of activated B cells; NK, natural killer; TIMP, tissue inhibitor of metalloproteinase; LOXL2, lysyl
oxidase 2. From [106]. Reproduced with permission of BMJ Publishing Group Ltd.
452 THE LIVER: REFERENCES

Natural killer T cells (NKT cells), a subpopulation of NK 12. Dufour, J.F., DeLellis, R., and Kaplan, M.M. Reversibility of hepatic fibrosis
cells also expressing the T‐cell receptor, likely have a dual role in autoimmune hepatitis. Ann Intern Med, 1997;127:981–5.
13. Falkmer, S., Samuelson, G., and Sjölin, S. Penicillamine‐induced normaliza-
in hepatic fibrogenesis by expressing both profibrogenic tion of clinical signs, and liver morphology and histochemistry in a case of
cytokines (IL4, IL13) and the CXCL16 receptor CXCR6. Wilson’s disease. Pediatrics, 1970;45:260–8.
LSECs and macrophages, which express CXCL16, can influ- 14. Kaplan, M.M., DeLellis, R.A., and Wolfe, H.J. Sustained biochemical and
ence NKT cell migration and promote NKT cell‐mediated histologic remission of primary biliary cirrhosis in response to medical treat-
ment. Ann Intern Med, 1997;126:682–8.
liver fibrogenesis [98, 99].
15. Kisseleva, T., Cong, M., Paik, Y. et al. Myofibroblasts revert to an inactive
phenotype during regression of liver fibrosis. Proc Natl Acad Sci USA,
2012;109:9448–53.
16. Troeger, J.S., Mederacke, I., Gwak, G.Y. et al. Deactivation of hepatic stel-
ANTIFIBROTIC STRATEGIES late cells during liver fibrosis resolution in mice. Gastroenterology, 2012;
143:1073–83.e22.
The elucidation of fibrogenic pathways and determinants for 17. Lee, Y.A. and Friedman, S.L. Reversal, maintenance or progression: What
happens to the liver after a virologic cure of hepatitis C? Antiviral Res, 2014;
resolution has made immense strides within the past 15 years 107C:23–30.
changing our perception of cirrhosis and advanced liver disease. 18. D’Ambrosio, R., Aghemo, A., Rumi, M.G. et al. A morphometric and immu-
Our understanding of underlying molecular mechanisms and key nohistochemical study to assess the benefit of a sustained virological response
cellular determinants for fibrosis regression have informed the in hepatitis C virus patients with cirrhosis. Hepatology, 2012;56:532–43.
19. Zeybel, M., Hardy, T., Wong, Y.K. et al. Multigenerational epigenetic adap-
transition of many antifibrotic agents into clinical trials [100–
tation of the hepatic wound‐healing response. Nat Med, 2012;18:1369–77.
102] with the following strategies: (i) controlling and reducing 20. Decaris, M.L., Emson, C.L., Li, K. et al. Turnover rates of hepatic collagen
the underlying disease, (ii) inhibiting pathways and cytokines and circulating collagen‐associated proteins in humans with chronic liver
critical for HSC activation, (iii) inhibiting most potent profibro- disease. PLoS One, 2015;10:e0123311.
genic pathways, and (iv) promoting resolution of fibrosis by 21. Hytiroglou, P., Snover, D.C., Alves, V. et al. Beyond “cirrhosis”: a proposal
from the International Liver Pathology Study Group. Am J Clin Pathol,
HSC clearance (apoptosis) or enhancing resolution of fibrosis
2012;137:5–9.
(see Figure 36.4) with a selection of drugs in clinical trials. 22. Browning, J.D. and Horton, J.D. Molecular mediators of hepatic steatosis
The identification of core fibrogenic pathways, which might be and liver injury. J Clin Invest, 2004;114:147–52.
shared across organ systems, for example, pulmonary fibrosis and 23. Urtasun, R., Conde de la Rosa, L., and Nieto, N. Oxidative and nitrosative
renal fibrosis has widened our arsenal of substances that might be stress and fibrogenic response. Clin Liver Dis, 2008;12:769–90.
24. Pessayre, D., Berson, A., Fromenty, B., and Mansouri, A. Mitochondria in
suitable for advanced liver diseases as well [103]. Big data analy- steatohepatitis. Semin Liver Dis, 2001;21:57–69.
ses with interrogation of large transcriptomic and genomic data- 25. Aoyama, T., Paik, Y.H., Watanabe, S. et al. Nicotinamide adenine dinucleo-
sets also offer potential for discovering diagnostics and therapeutics tide phosphate oxidase in experimental liver fibrosis: GKT137831 as a novel
for liver disease [52, 104]. All these developments should pave the potential therapeutic agent. Hepatology, 2012;56:2316–27.
way for a range of antifibrotic therapies in the coming years. 26. Yoshiji, H., Kuriyama, S., Yoshii, J. et al. Vascular endothelial growth factor
and receptor interaction is a prerequisite for murine hepatic fibrogenesis.
Gut, 2003;52:1347–54.
27. Ankoma‐Sey, V., Matli, M., Chang, K.B. et  al. Coordinated induction of
VEGF receptors in mesenchymal cell types during rat hepatic wound heal-
REFERENCES ing. Oncogene, 1998;17:115–21.
28. Olaso, E., Salado, C., Egilegor, E. et al. Proangiogenic role of tumor‐acti-
1. Byass, P. The global burden of liver disease: a challenge for methods and for vated hepatic stellate cells in experimental melanoma metastasis. Hepatology,
public health. BMC Med, 2014;12:159. 2003;37:674–85.
2. Marcellin, P. and Kutala, B.K. Liver diseases: a major, neglected global pub- 29. Semela, D., Das, A., Langer, D., Kang, N., Leof, E., and Shah, V. Platelet‐
lic health problem requiring urgent actions and large‐scale screening. Liver derived growth factor signaling through ephrin‐b2 regulates hepatic vascular
Int, 2018;38(1):2–6. structure and function. Gastroenterology, 2008;135:671–9.
3. Scaglione, S., Kliethermes, S., Cao, G. et al. The epidemiology of cirrhosis 30. Fernandez, M., Semela, D., Bruix, J., Colle, I., Pinzani, M., and Bosch, J.
in the United States: a population‐based study. J Clin Gastroenterol, 2015; Angiogenesis in liver disease. J Hepatol, 2009;50:604–20.
49:690–6. 31. Canbay, A., Taimr, P., Torok, N., Higuchi, H., Friedman, S., and Gores, G.J.
4. Hope, V.D., Eramova, I., Capurro, D., and Donoghoe, M.C. Prevalence and Apoptotic body engulfment by a human stellate cell line is profibrogenic.
estimation of hepatitis B and C infections in the WHO European region: a Lab Invest, 2003;83:655–63.
review of data focusing on the countries outside the European Union and the 32. Canbay, A., Higuchi, H., Bronk, S.F., Taniai, M., Sebo, T.J., and Gores, G.J.
European Free Trade Association. Epidemiol Infect, 2014;142:270–86. Fas enhances fibrogenesis in the bile duct ligated mouse: a link between
5. European AFTSOL. EASL clinical practical guidelines: management of apoptosis and fibrosis. Gastroenterology, 2002;123:1323–30.
alcoholic liver disease. J Hepatol, 2012;57:399–420. 33. Canbay, A., Friedman, S., and Gores, G.J. Apoptosis: the nexus of liver
6. Roguin, A. Rene Theophile Hyacinthe Laënnec (1781–1826): the man injury and fibrosis. Hepatology, 2004;39:273–8.
behind the stethoscope. Clin Med Res, 2006;4:230–5. 34. Schwabe, R.F. and Luedde, T. Apoptosis and necroptosis in the liver: a mat-
7. Yuen, M.F. and Lai, C.L. Treatment of chronic hepatitis B: evolution over ter of life and death. Nat Rev Gastroenterol Hepatol, 2018;15:738–52.
two decades. J Gastroenterol Hepatol, 2011;26(1):138–43. 35. Hendriks, H.F., Verhoofstad, W.A., Brouwer, A., de Leeuw, A.M., and
8. Ohkoshi, S., Hirono, H., Watanabe, K., Hasegawa, K., Kamimura, K., and Knook, D.L. Perisinusoidal fat‐storing cells are the main vitamin A storage
Yano, M. Natural regression of fibrosis in chronic hepatitis B. J Gastroenterol, sites in rat liver. Exp Cell Res, 1985;160:138–49.
2016;22:5459–66. 36. Friedman, S.L. Molecular regulation of hepatic fibrosis, an integrated cellu-
9. Gonzalez, H.C. and Duarte‐Rojo, A. Virologic cure of hepatitis c: impact on lar response to tissue injury. J Biol Chem, 2000;275:2247–50.
hepatic fibrosis and patient outcomes. Curr Gastroenterol Rep, 2016;18:32. 37. Bataller, R. and Brenner, D.A. Liver fibrosis. J Clin Invest, 2005;115:209–18.
10. Powell, L.W. and Kerr, J.F. Reversal of “cirrhosis” in idiopathic haemochro- 38. Schulze‐Krebs, A., Preimel, D., Popov, Y. et al. Hepatitis C virus‐replicating
matosis following long‐term intensive venesection therapy. Australas Ann hepatocytes induce fibrogenic activation of hepatic stellate cells. Gastroen­
Med, 1970;19:54–7. terology, 2005;129:246–58.
11. Blumberg, R.S., Chopra, S., Ibrahim, R. et al. Primary hepatocellular carcinoma 39. Lua, I., Li, Y., Zagory, J.A. et al. Characterization of hepatic stellate cells,
in idiopathic hemochromatosis after reversal of cirrhosis. Gastroenterology, portal fibroblasts, and mesothelial cells in normal and fibrotic livers.
1988;95:1399–402. J Hepatol, 2016;64:1137–46.
36:  Stellate Cells and Fibrosis 453

40. Mederacke, I., Hsu, C.C., Troeger, J.S. et al. Fate tracing reveals hepatic stel- 66. Friedman, S.L. Hepatic stellate cells: protean, multifunctional, and enig-
late cells as dominant contributors to liver fibrosis independent of its aetiol- matic cells of the liver. Physiol Rev, 2008;88:125–72.
ogy. Nat Commun, 2013;4:2823. 67. Huang, G. and Brigstock, D.R. Regulation of hepatic stellate cells by con-
41. Iwaisako, K., Jiang, C., Zhang, M. et  al. Origin of myofibroblasts in the nective tissue growth factor. Front Biosci, 2012;17:2495–507.
fibrotic liver in mice. Proc Natl Acad Sci USA, 2014;111:E3297–305. 68. Reynaert, H., Thompson, M.G., Thomas, T., and Geerts, A. Hepatic stellate
42. Lemoinne, S., Cadoret, A., El Mourabit, H., Thabut, D., and Housset, C. cells: role in microcirculation and pathophysiology of portal hypertension.
Origins and functions of liver myofibroblasts. Biochim Biophys Acta, 2013; Gut, 2002;50:571–81.
1832:948–54. 69. Yang, L., Kwon, J., Popov, Y. et al. Vascular endothelial growth factor pro-
43. Uchio, K., Tuchweber, B., Manabe, N., Gabbiani, G., Rosenbaum, J., and motes fibrosis resolution and repair in mice. Gastroenterology, 2014;146:
Desmoulière, A. Cellular retinol‐binding protein‐1 expression and modula- 1339–50.
tion during in vivo and in vitro myofibroblastic differentiation of rat hepatic 70. Kantari‐Mimoun, C., Castells, M., Klose, R. et al. Resolution of liver fibrosis
stellate cells and portal fibroblasts. Lab Invest, 2002;82:619–28. requires myeloid cell‐driven sinusoidal angiogenesis. Hepatology, 2015;61:
44. Dranoff, J.A. and Wells, R.G. Portal fibroblasts: underappreciated mediators 2042–55.
of biliary fibrosis. Hepatology, 2010;51:1438–44. 71. Sprenger, H., Kaufmann, A., Garn, H., Lahme, B., Gemsa, D., and Gressner,
45. Karin, D., Koyama, Y., Brenner, D., and Kisseleva, T. The characteristics of A.M. Induction of neutrophil‐attracting chemokines in transforming rat
activated portal fibroblasts/myofibroblasts in liver fibrosis. Differentiation, hepatic stellate cells. Gastroenterology, 1997;113:277–85.
2016;92:84–92. 72. Marra, F., Valente, A.J., Pinzani, M., and Abboud, H.E. Cultured human liver
46. Lamireau, T., Dubuisson, L., Lepreux, S. et al. Abnormal hepatic expression of fat‐storing cells produce monocyte chemotactic protein‐1. Regulation by
fibrillin‐1 in children with cholestasis. Am J Surg Pathol, 2002;26:637–46. proinflammatory cytokines. J Clin Invest, 1993;92:1674–80.
47. Fausther, M., Goree, J.R., Lavoie, É.G., Graham, A.L., Sévigny, J., and 73. Lefebvre, E., Moyle, G., Reshef, R. et  al. Antifibrotic effects of the dual
Dranoff, J.A. Establishment and characterization of rat portal myofibroblast CCR2/CCR5 antagonist cenicriviroc in animal models of liver and kidney
cell lines. PLoS One, 2015;10:e0121161. fibrosis. PLoS One, 2016;11:e0158156.
48. Xu, L., Hui, A.Y., Albanis, E. et al. Human hepatic stellate cell lines, LX‐1 74. Friedman, S.L., Ratziu, V., Harrison, S.A. et al. A randomized, placebo‐con-
and LX‐2: new tools for analysis of hepatic fibrosis. Gut, 2005;54:142–51. trolled trial of cenicriviroc for treatment of nonalcoholic steatohepatitis with
49. Shibata, N., Watanabe, T., Okitsu, T. et al. Establishment of an immortalized fibrosis. Hepatology, 2018;67:1754–67.
human hepatic stellate cell line to develop antifibrotic therapies. Cell 75. Winau, F., Hegasy, G., Weiskirchen, R. et al. Ito cells are liver‐resident anti-
Transplant, 2003;12:499–507. gen‐presenting cells for activating T cell responses. Immunity, 2007;26:
50. Shen, K., Chang, W., Gao, X. et al. Depletion of activated hepatic stellate 117–29.
cell correlates with severe liver damage and abnormal liver regeneration in 76. Yoshiji, H., Kuriyama, S., Miyamoto, Y. et al. Tissue inhibitor of metallopro-
acetaminophen‐induced liver injury. Acta Biochim Biophys Sin (Shanghai), teinases‐1 promotes liver fibrosis development in a transgenic mouse model.
2011;43:307–15. Hepatology, 2000;32:1248–54.
51. Puche, J.E., Lee, Y.A., Jiao, J. et al. A novel murine model to deplete hepatic 77. Benyon, R.C. and Arthur, M.J. Extracellular matrix degradation and the role
stellate cells uncovers their role in amplifying liver damage in mice. of hepatic stellate cells. Semin Liver Dis, 2001;21:373–84.
Hepatology, 2013;57:339–50. 78. Iredale, J.P. Hepatic stellate cell behavior during resolution of liver injury.
52. Zhang, D.Y., Goossens, N., Guo, J. et al. A hepatic stellate cell gene expres- Semin Liver Dis, 2001;21:427–36.
sion signature associated with outcomes in hepatitis C cirrhosis and hepato- 79. Arthur, M.J. Collagenases and liver fibrosis. J Hepatol, 1995;22:43–8.
cellular carcinoma after curative resection. Gut, 2016;65(10):1754–64. 80. Ikejima, K., Takei, Y., Honda, H. et al. Leptin receptor‐mediated signaling
53. Yin, C., Evason, K.J., Maher, J.J., and Stainier, D.Y. The bHLH transcription regulates hepatic fibrogenesis and remodeling of extracellular matrix in the
factor Hand2 marks hepatic stellate cells in zebrafish: analysis of stellate cell rat. Gastroenterology, 2002;122:1399–410.
entry into the developing liver. Hepatology, 2012; 56(5):1958–70. 81. Leclercq, I.A., Farrell, G.C., Schriemer, R., and Robertson, G.R. Leptin is
54. Enzan, H., Himeno, H., Iwamura, S. et al. Sequential changes in human Ito essential for the hepatic fibrogenic response to chronic liver injury. J
cells and their relation to postnecrotic liver fibrosis in massive and submas- Hepatol, 2002;37:206–13.
sive hepatic necrosis. Virchows Arch, 1995;426:95–101. 82. Saxena, N.K., Saliba, G., Floyd, J.J., and Anania, F.A. Leptin induces
55. Inagaki, Y. and Okazaki, I. Emerging insights into transforming growth fac- increased alpha2(I) collagen gene expression in cultured rat hepatic stellate
tor beta Smad signal in hepatic fibrogenesis. Gut, 2007;56:284–92. cells. J Cell Biochem, 2003;89:311–20.
56. Hellerbrand, C., Stefanovic, B., Giordano, F., Burchardt, E.R., and Brenner, 83. Coombes, J.D., Choi, S.S., Swiderska‐Syn, M. et al. Osteopontin is a proxi-
D.A. The role of TGFbeta1 in initiating hepatic stellate cell activation in mal effector of leptin‐mediated non‐alcoholic steatohepatitis (NASH) fibro-
vivo. J Hepatol, 1999;30:77–87. sis. Biochim Biophys Acta, 2016;1862:135–44.
57. Hanafusa, H., Ninomiya‐Tsuji, J., Masuyama, N. et al. Involvement of the 84. Sahai, A., Malladi, P., Melin‐Aldana, H., Green, R.M., and Whitington, P.F.
p38 mitogen‐activated protein kinase pathway in transforming growth fac- Upregulation of osteopontin expression is involved in the development of
tor‐beta‐induced gene expression. J Biol Chem, 1999;274:27161–7. nonalcoholic steatohepatitis in a dietary murine model. Am J Physiol
58. Engel, M.E., McDonnell, M.A., Law, B.K., and Moses, H.L. Interdependent Gastrointest Liver Physiol, 2004;287:G264–73.
SMAD and JNK signaling in transforming growth factor‐beta‐mediated tran- 85. Aleffi, S., Petrai, I., Bertolani, C. et al. Upregulation of proinflammatory and
scription. J Biol Chem, 1999;274:37413–20. proangiogenic cytokines by leptin in human hepatic stellate cells. Hepatology,
59. Fan, X., Zhang, Q., Li, S. et al. Attenuation of CCl4‐induced hepatic fibrosis 2005;42:1339–48.
in mice by vaccinating against TGF‐β1. PLoS One, 2013;8:e82190. 86. Hernandez‐Gea, V., Ghiassi‐Nejad, Z., Rozenfeld, R. et  al. Autophagy
60. Ling, H., Roux, E., Hempel, D. et al. Transforming growth factor β neutrali- releases lipid that promotes fibrogenesis by activated hepatic stellate cells in
zation ameliorates pre‐existing hepatic fibrosis and reduces cholangiocarci- mice and in human tissues. Gastroenterology, 2012;142:938–46.
noma in thioacetamide‐treated rats. PLoS One, 2013;8:e54499. 87. Arshad, M.I., Piquet‐Pellorce, C., and Samson, M. IL‐33 and HMGB1
61. Wong, L., Yamasaki, G., Johnson, R.J., and Friedman, S.L. Induction of alarmins: sensors of cellular death and their involvement in liver pathology.
beta‐platelet‐derived growth factor receptor in rat hepatic lipocytes during Liver Int, 2012;32:1200–10.
cellular activation in vivo and in culture. J Clin Invest, 1994;94:1563–9. 88. McHedlidze, T., Waldner, M., Zopf, S. et  al. Interleukin‐33‐dependent
62. Pinzani, M. PDGF and signal transduction in hepatic stellate cells. Front innate lymphoid cells mediate hepatic fibrosis. Immunity,
Biosci, 2002;7:d1720–6. 2013;39:357–71.
63. Pinzani, M., Milani, S., Grappone, C., Weber, F.L.J., Gentilini, P., and 89. Garcia‐Trevijano, E.R., Iraburu, M.J., Fontana, L. et al. Transforming growth
Abboud, H.E. Expression of platelet‐derived growth factor in a model of factor beta1 induces the expression of alpha1(I) procollagen mRNA by a
acute liver injury. Hepatology, 1994;19:701–7. hydrogen peroxide‐C/EBPbeta‐dependent mechanism in rat hepatic stellate
64. Kocabayoglu, P., Lade, A., Lee, Y.A. et al. beta‐PDGF receptor expressed by cells. Hepatology, 1999;29:960–70.
hepatic stellate cells regulates fibrosis in murine liver injury, but not carcino- 90. Serini, G., Bochaton‐Piallat, M.L., Ropraz, P. et al. The fibronectin domain
genesis. J Hepatol, 2015;63(1):141–7. ED‐A is crucial for myofibroblastic phenotype induction by transforming
65. Deshmukh, M., Nakagawa, S., Higashi, T. et al. Cell type‐specific pharma- growth factor‐beta1. J Cell Biol, 1998;142:873–81.
cological kinase inhibition for cancer chemoprevention. Nanomedicine, 91. Jarnagin, W.R., Rockey, D.C., Koteliansky, V.E., Wang, S.S., and Bissell,
2018;14:317–25. D.M. Expression of variant fibronectins in wound healing: cellular source
454 THE LIVER: REFERENCES

and biological activity of the EIIIA segment in rat hepatic fibrogenesis.  99. Wehr, A., Baeck, C., Heymann, F. et  al. Chemokine receptor CXCR6‐
J Cell Biol, 1994;127:2037–48. dependent hepatic NK T Cell accumulation promotes inflammation and
92. Thiele, G.M., Duryee, M.J., Freeman, T.L. et al. Rat sinusoidal liver endothe- liver fibrosis. J Immunol, 2013;190:5226–36.
lial cells (SECs) produce pro‐fibrotic factors in response to adducts formed 100. Friedman, S.L. Hepatic fibrosis: emerging therapies. Dig Dis, 2015;33:
from the metabolites of ethanol. Biochem Pharmacol, 2005;70:1593–600. 504–7.
93. George, J., Wang, S.S., Sevcsik, A.M. et al. Transforming growth factor‐beta 101. Fallowfield, J.A. Future mechanistic strategies for tackling fibrosis—an
initiates wound repair in rat liver through induction of the EIIIA‐fibronectin unmet need in liver disease. Clin Med (Lond), 2015;15(6):s83–7.
splice isoform. Am J Pathol, 2000;156:115–24. 102. Friedman, S.L., Neuschwander‐Tetri, B.A., Rinella, M., and Sanyal, A.J.
94. Ding, B.S., Cao, Z., Lis, R. et al. Divergent angiocrine signals from vascular Mechanisms of NAFLD development and therapeutic strategies. Nat Med,
niche balance liver regeneration and fibrosis. Nature, 2014;505:97–102. 2018;24:908–22.
95. Wright, M., Goldin, R., Hellier, S. et al. Factor V Leiden polymorphism and 103. Friedman, S.L., Sheppard, D., Duffield, J.S., and Violette, S. Therapy for
the rate of fibrosis development in chronic hepatitis C virus infection. Gut, fibrotic diseases: nearing the starting line. Sci Transl Med, 2013;5:167sr1.
2003;52:1206–10. 104. Wooden, B., Goossens, N., Hoshida, Y., and Friedman, S.L. Using big data
96. Papatheodoridis, G.V., Papakonstantinou, E., Andrioti, E. et al. Thrombotic to discover diagnostics and therapeutics for gastrointestinal and liver dis-
risk factors and extent of liver fibrosis in chronic viral hepatitis. Gut, eases. Gastroenterology, 2017;152:53–67.
2003;52:404–9. 105. Martinez‐Hernandez, A. and Amenta, P.S. The hepatic extracellular matrix.
97. Krizhanovsky, V., Yon, M., Dickins, R.A. et al. Senescence of activated stel- I. Components and distribution in normal liver. (editorial). Virchows Arch A
late cells limits liver fibrosis. Cell, 2008;134:657–67. Pathol Anat Histopathol, 1993;423(1):1.
98. Liepelt, A., Wehr, A., Kohlhepp, M. et al. CXCR6 protects from inflamma- 106. Lee, Y.A., Wallace, M.C., and Friedman, S.L. Pathobiology of liver fibro-
tion and fibrosis in NEMOLPC‐KO mice. Biochim Biophys Acta Mol Basis Dis, sis: a translational success story. Gut, 2015;64:830–41.
2018;1865(2):391–402.
PART THREE:
FUNCTIONS OF THE LIVER
SECTION A:
METABOLIC FUNCTIONS
Non‐alcoholic Fatty Liver
37 Disease and Insulin
Resistance
Max C. Petersen1,3, Varman T. Samuel1,2, Kitt Falk Petersen1, and
Gerald I. Shulman1,3
1
Department of Internal Medicine, Section of Endocrinology, Yale University School of Medicine,
New Haven, CT, USA
2
Veterans Affairs Medical Center, West Haven, CT, USA
3
Department of Molecular and Cellular Physiology, Yale University School of Medicine, New Haven, CT, USA

INTRODUCTION myocyte and hepatocyte causes insulin resistance have been


partially elucidated in recent years. This chapter will: (i) review
In the wake of the obesity epidemic, the world has witnessed a the current model for the pathogenesis of insulin resistance in
rising prevalence of associated conditions such as the metabolic the skeletal muscle, (ii) discuss how insulin resistance in s­ keletal
syndrome, type 2 diabetes mellitus (T2D), and non‐alcoholic muscle may promote the development of NAFLD, (iii) explore
fatty liver disease (NAFLD). The International Diabetes the mechanisms whereby NAFLD may lead to hepatic insulin
Federation estimated the 2017 prevalence of T2D at over 400 resistance, and finally, (iv) review the role of insulin‐sensitizing
million adults worldwide [1]. Diabetes is a leading cause of therapies in correcting hepatic insulin resistance.
blindness in working adults, end‐stage renal failure, non‐traumatic
limb loss, and is a major risk factor for ischemic heart disease
[2]. Similarly, NAFLD is increasingly recognized as a major
threat to public health. NAFLD is now considered the most MECHANISM OF LIPID‐INDUCED
common chronic liver disease, with estimated prevalence of SKELETAL MUSCLE INSULIN
around 30% in the United States [3], and is a major risk factor RESISTANCE
for the development of non‐alcoholic steatohepatitis (NASH),
cirrhosis, and liver‐related death [4]. As cirrhosis secondary to A review of the mechanisms responsible for skeletal muscle
viral hepatitis becomes rarer, NASH is poised to become the lead- insulin resistance serves as an important reference point from
ing indication for liver transplantation in the United States [5]. which to proceed to a discussion of hepatic insulin resistance. In
Hepatic insulin resistance frequently accompanies NAFLD skeletal muscle, insulin binds to its receptor, activating insulin
[4]. Simply considered, resistance to insulin develops in the receptor tyrosine kinase activity with subsequent phosphoryla-
peripheral tissues (muscle and fat) and in the liver. Insulin resist- tion and activation of insulin receptor substrates (e.g. IRS1).
ance in the periphery impairs insulin‐stimulated glucose uptake When phosphorylated, IRS1 activates phosphatidylinositol‐3‐
and muscle glycogen synthesis. In the liver, insulin resistance kinase (PI3 kinase) and, through a cascade of downstream
diminishes the ability of insulin to suppress gluconeogenesis ­signaling effectors, ultimately increases the translocation of glu-
and stimulate net hepatic glycogen synthesis. The distinct cose transporter 4 (GLUT4)‐containing vesicles to the plasma
mechanisms by which lipid accumulation within the skeletal membrane. A current concept of muscle insulin resistance links

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
460 THE LIVER:  MECHANISM OF LIPID‐INDUCED SKELETAL MUSCLE INSULIN RESISTANCE

intramyocellular lipid accumulation  –  specifically, diacylglyc- more than three hours, a second mechanism for lipid inhibition
erol (DAG) accumulation  –  with impaired insulin signaling of glucose utilization in the skeletal muscle was apparent.
and  consequent impairments in insulin stimulation of glucose Roden et  al. infused normal insulin‐sensitive subjects with
transport [6]. either glycerol or Intralipid/heparin. As the triglyceride in the
This model has evolved over several decades. In the 1960s, Intralipid was hydrolyzed by lipoprotein lipase and released
Randle et al. first formed a hypothesis linking increased lipids from the endothelium by heparin, the plasma non‐esterified
to decreased glucose metabolism in muscle [7]. Randle ­reasoned fatty acid (NEFA) concentration increased. Using 13C MRS to
that increased delivery of fatty acids (FAs) should promote FA measure muscle glycogen content, insulin‐stimulated muscle
oxidation and inhibit glucose oxidation. The mechanism was as glycogen synthesis was then assessed during a euglycemic–
follows: lipid oxidation increases the intramitochondrial [acetyl‐ hyperinsulinemic clamp [15]. After 3 hours of Intralipid/heparin
CoA]/[CoA] and [NADH]/[NAD+] ratios, which in turn leads to infusion, the rate of insulin‐mediated glycogen synthesis began
the inactivation of pyruvate dehydrogenase (PDH), the enzyme to decrease. This decrease in glycogen synthesis was preceded
responsible for the conversion of pyruvate into acetyl‐CoA. by a drop in the concentration of G6P, as assessed by 31P MRS.
Additionally, accumulation of intracellular citrate inhibits phos- This finding was inconsistent with the glucose–fatty acid cycle,
phofructokinase (PFK), a key glycolytic enzyme. The block in which would have predicted an accumulation of G6P. Instead,
PFK would lead to accumulation of glucose 6‐phosphate (G6P), these data suggested that the lipid infusion caused muscle insu-
which in turn inhibits hexokinase activity. This inhibition of lin resistance by inducing a defect in either glucose transport or
hexokinase activity would result in an increase in intracellular phosphorylation activity. To distinguish between these two pos-
glucose, decreasing the chemical driving force for glucose sibilities, Dresner et al. studied normal subjects with a similar
uptake. In essence, Randle hypothesized that oxidation of fat glycerol vs. Intralipid/heparin infusion protocol and used a 13C
prevents oxidation of glucose by inhibiting key glycolytic MRS technique to measure intramyocellular glucose concentra-
enzymes. This model, known as the glucose–fatty acid cycle, tions [16]. Compared to the glycerol infusion, the Intralipid/
provided a logical biochemical basis for oxidative substrate heparin infusion was associated with lower concentrations of
switching based on local availability and was observed in intramyocellular glucose and a failure to increase G6P concen-
­multiple experimental settings during the 1980s and 1990s, par- trations. These data suggested that the lipid infusion induced
ticularly in humans and rats subjected to acute increases in muscle insulin resistance by impairing insulin‐stimulated glu-
plasma FA concentrations [8, 9]. cose transport activity, consistent with the defect observed in
diabetic humans [14]. Furthermore, in muscle biopsies obtained
Intramyocellular lipid accumulation causes from these subjects, the apparent defect in insulin‐stimulated
glucose uptake was associated with impaired activation of
muscle insulin resistance
IRS1‐associated PI3 kinase activity [16].
The development of non‐invasive magnetic resonance spectros- The generation of mice with constitutively active pyruvate
copy (MRS) methods to interrogate metabolism in vivo sparked dehydrogenase (PDH), produced by deletion of pyruvate
an interest in reassessing the link between intracellular lipids ­dehydrogenase kinases 2 and 4 (Pdk2−/−/Pdk4−/− mice), enabled
and glucose metabolism. Using 1H MRS and the hyperinsuline- a particularly dramatic illustration of the unnecessity of the
mic–euglycemic clamp method in young, non‐diabetic, non‐ ­glucose–fatty acid cycle for lipid‐induced muscle insulin resist-
obese humans, intramyocellular lipid (IMCL) content was ance. With this genetic activation of PDH, the Pdk2−/−/Pdk4−/−
found to be a strong predictor of insulin resistance [10, 11]. mice preferentially oxidized glucose in both fasting and fed
Rothman et al., using 13C and 31P MRS, studied the biochemical states. The Pdk2−/−/Pdk4−/− mice, unable to switch to fatty acid
mechanisms responsible by measuring insulin‐mediated oxidation, accumulated IMCL and displayed profound muscle
changes in muscle glycogen and G6P both in patients with T2D insulin resistance during hyperinsulinemic–euglycemic clamp
[12] and in non‐diabetic, first‐degree relatives of patients with studies [17].
T2D [13]. In contrast to Randle’s hypothesis, the accumulation Taken together, these studies indicated that lipid‐induced
of G6P was diminished in both the diabetic subjects and the muscle insulin resistance occurs by a mechanism independent
first‐degree relatives of diabetic patients. Subsequent MRS of the Randle glucose–fatty acid cycle. In this mechanism,
experiments in diabetic human subjects revealed that neither intramyocellular lipids induce a defect in the insulin signaling
intramyocellular G6P nor intramyocellular glucose rose to lev- pathway that blunts insulin stimulation of myocellular glucose
els remotely high enough for the glucose–fatty acid cycle to transport.
account for the diabetic impairment in insulin‐stimulated
­muscle glycogen synthesis [14]. These data demonstrated that
muscle insulin resistance in human diabetes was due to impair-
How does lipid accumulate in muscle?
ment of insulin stimulated glucose transport, rather than accu- Simply stated, IMCL accumulates when myocellular lipid
mulation of glycolytic intermediates. The data also indicated uptake exceeds its utilization. On the supply side of the equation
that these changes were present in the insulin‐resistant muscle are the increased concentration of plasma lipids often seen in
prior to the development of overt T2D. obesity and the metabolic syndrome. As described above,
Although the Randle glucose–fatty acid cycle had been increasing delivery of fat to the muscle via lipid infusion induces
observed in the setting of acute lipid administration, it could not insulin resistance in normal, insulin‐sensitive subjects. At the
account for the impaired insulin‐mediated glucose metabolism level of the muscle, flux into the myocyte may be increased by
of insulin‐resistant human muscle. When lipid was infused for the action of lipoprotein lipase, which acts at the endothelium to
37:  Non‐alcoholic Fatty Liver Disease and Insulin Resistance 461

release fatty acids from circulating triglycerides. Pollare et al. age‐related decreases in mitochondrial oxidation and phospho-
measured lipoprotein lipase (LPL) activity and insulin sensitiv- rylation were associated with increases in intramyocellular tri-
ity, as assessed by the hyperinsulinemic–euglycemic clamp in glyceride content and decreases in insulin sensitivity [21]. These
insulin‐sensitive controls and various insulin‐resistant popula- observations, in both aging subjects and in the insulin‐resistant
tions (obese/non‐diabetic, obese/diabetic, etc.) [18]. They found offspring of T2D patients, suggest that impaired mitochondrial
that muscle LPL activity was inversely related to the glucose function is a common mechanism which predisposes to myocel-
infusion rate during the clamp and positively correlated to lular lipid accumulation and insulin resistance.
­fasting plasma insulin. That is, the higher the LPL activity, the
more insulin‐resistant the subject. The valine breakdown prod-
uct 3‐hydroxyisobutyrate (3‐HIB) was recently shown to be a How does muscle fat accumulation cause
paracrine positive regulator of myocellular FA uptake; produced muscle insulin resistance?
by muscle cells, it acts at the endothelium [19]. 3‐HIB supple- The above discussion has presented the evidence associating
mentation in mice increased IMCL and decreased muscle insu- IMCL and muscle insulin resistance in humans. Mechanistic
lin sensitivity, and muscle 3‐HIB levels were increased in insights linking accumulation of lipid metabolites to impair-
diabetic db/db mice and diabetic human subjects. The physio- ments in insulin action have been gleaned from rodent studies,
logic significance of 3‐HIB regulation of myocellular FA deliv- and more recently examined in humans. DAG activates protein
ery is unclear, but it is worth noting that increased plasma levels kinase C (PKC) isoforms, and intramyocellular DAG accumula-
of branched‐chain amino acids, including valine, are associated tion has been repeatedly associated with activation of protein
with insulin resistance [20]. These studies support the hypothe- kinase C‐θ (PKCθ) (Figure 37.1a, b) [6]. This isoform belongs
sis that increased delivery and uptake of FA promotes accumu- to the “novel” class of PKCs that exhibit sn 1,2 DAG‐dependent,
lation of IMCL and muscle insulin resistance. Ca2+‐independent activation. Skeletal muscle DAG accumula-
Balancing muscle lipid delivery is muscle lipid utilization, tion has been shown to temporally precede the activation of
specifically mitochondrial fat oxidation. Muscle mitochondrial PKCθ and the development of muscle insulin resistance during
function has been shown to be impaired in two groups of patients an acute lipid infusion [29]. Importantly, PKCθ knockout mice
who are at risk for developing T2D: the offspring of patients were protected from lipid‐induced insulin resistance during
with T2D and the elderly [21, 22]. Mitochondrial function can lipid infusion–glucose clamp studies [30], but the PKCθ targets
be assessed in vivo using a combination of 13C MRS with 13C‐ responsible for this phenotype remain incompletely defined.
acetate infusion to quantify tricarboxylic acid (TCA) cycle flux
IRS1 Ser1101 was identified as a PKCθ substrate [31], but mice
and 31P MRS to quantify ATP synthesis. These techniques were
carrying a serine‐to‐alanine mutation at this residue are not pro-
used to study lean offspring of T2D patients, who already had
tected from lipid‐induced insulin resistance [32].
skeletal muscle lipid accumulation and insulin resistance, and
Current data support a role for PKCθ activation in the patho-
were at high risk for developing T2D later in life [22]. These
genesis of lipid‐induced muscle insulin resistance in humans as
patients were found to have around 40% lower rates of basal
well. Increased myocellular DAG and PKCθ activity has been
mitochondrial ATP synthesis [22]. The mechanistic basis for
observed in the muscles of patients with T2D compared to non-
decreased mitochondrial activity in aging is likely multifacto-
diabetic controls [33, 34]. Muscle insulin resistance following
rial, but oxidative free radical damage to mitochondrial DNA
an oral or intravenous lipid challenge has also been associated
(mtDNA) is one potential contributor. In mice, targeted overex-
with intramyocellular DAG accumulation and PKCθ activation
pression of the free radical scavenging enzyme catalase to the
[34, 35]. Overall, a substantial body of correlative data links the
mitochondrion prevented age‐associated mtDNA damage, asso-
DAG–PKCθ axis to lipid‐induced skeletal muscle insulin resist-
ciated with preserved skeletal muscle mitochondrial oxidative
ance, but the mechanisms by which activated PKCθ impairs
capacity and insulin sensitivity [23]. Interestingly, recent stud-
insulin signaling remain incompletely defined.
ies of the human genetics of T2D are consistent with a role for
mitochondrial function in T2D pathogenesis, although the tis-
sues implicated have been liver and adipose rather than skeletal
muscle. Loss of the murine homolog of N‐acetyltransferase 2
(NAT2), a polymorphism of which has been associated with
PERIPHERAL INSULIN RESISTANCE
human insulin resistance independent of BMI, results in reduced CONTRIBUTES TO THE DEVELOPMENT
mitochondrial function and whole‐body energy expenditure, OF HEPATIC STEATOSIS
predisposing to lipid‐induced insulin resistance in mice [24,
25]. Variants in SLC16A11, which account for around 20% of There is a close association between the development of NAFLD
the increase in T2D risk among Mexicans, alter mitochondrial and peripheral insulin resistance. An attractive model posits that
fatty acid oxidation to favor lipid storage in hepatocytes [26]. muscle insulin resistance, by impairing postprandial muscle gly-
The development of insulin resistance is not limited to those cogen synthesis [36], facilitates hepatic de novo lipogenesis and
with an inherent susceptibility toward insulin resistance. Indeed, resultant intrahepatic triglyceride (IHTG) accumulation. This
as we age, most humans will manifest similar changes; T2D hypothesis fits data collected by several investigators. IHTG can
prevalence increases steadily with age [1]. In hyperinsulinemic– arise from either de novo lipogenesis or re‐esterification of FAs
euglycemic clamp studies, elderly subjects have been found to released from adipose or absorbed after a meal. Donnelly et al.
manifest both peripheral and hepatic insulin resistance [27, 28]. measured the relative contributions of these pathways in subjects
Using 13C and 31P MRS techniques, Petersen et al. showed that with NAFLD [37]. Although re‐esterification of circulating
462 THE LIVER:  PERIPHERAL INSULIN RESISTANCE CONTRIBUTES TO THE DEVELOPMENT OF HEPATIC STEATOSIS

(a) Muscle (c) Liver

IRS2
IRS1
Insulin IR P P PI3K
Insulin IR P P PI3K
P P
P P AKT
AKT
Insulin
Sensitive G6PC
PCK1
GCK
GLUT4
storage
vesicle

Glucose
Glucose

Glucose
Glycogen Glucose Glycogen

(b) (d)

sn-1,2 DAG sn-1,2 DAG

PKCθ PKCε

Insulin IR Insulin IR pT1160


Insulin
Resistant
GLUT4 Insulin
storage
vesicle
Glycogen synthesis

EGP suppression

Figure 37.1  Pathogenesis of lipid‐induced insulin resistance. (a) Normal insulin action on glucose metabolism in skeletal muscle. In the myocyte,
insulin acts to stimulate the translocation of GLUT4‐containing vesicles to the plasma membrane, allowing for glucose entry into the myocyte.
(b) Lipid‐induced insulin resistance in the skeletal muscle involves a blockade in proximal insulin signaling by PKCθ, which is activated by the
bioactive lipid DAG. This results in impaired insulin‐stimulated GLUT4 translocation. (c) Normal insulin action on glucose metabolism in liver. In
the hepatocyte, insulin inhibits gluconeogenesis and activates glycogen synthesis, thereby decreasing hepatic glucose production. (d) Lipid‐induced
insulin resistance in the hepatocyte. In the steatotic liver, DAG‐activated PKCε inhibits insulin receptor tyrosine kinase activity by phosphorylation
of Thr1160, impairing all downstream insulin action.

NEFAs was the dominant pathway (accounting for over half of fed and fasted states in control subjects [37]. Diraison et al. also
IHTG synthesis), rates of de novo lipogenesis were severalfold used a stable isotope approach to measure sources of IHTG and
higher than those previously observed in lean control subjects observed an approximately threefold increase in rates of de novo
and accounted for about one‐fourth of IHTG synthesis [37]. lipogenesis in subjects with NAFLD from normal controls [38].
Moreover, in these NAFLD patients, de novo lipogenesis was Finally, in a study of adult subjects with the metabolic syndrome,
persistently elevated without the normal oscillations seen during Lambert et  al. also observed a threefold increase in absolute
37:  Non‐alcoholic Fatty Liver Disease and Insulin Resistance 463

de  novo lipogenic flux in subjects with NAFLD compared to ­resistant subjects commonly display increased rates of de novo
those without [39]. These reports strongly suggest that increased lipogenesis. For this reason, it has been proposed that the liver
de novo lipogenesis is a contributor to NAFLD. exhibits “pathway‐selective hepatic insulin resistance”; resist-
However, is increased lipogenesis present even before ance to insulin’s effects on glucose metabolism but not to those
NAFLD has developed in subjects at risk, such as lean but insu- involving lipid metabolism [41]. However, upon closer exami-
lin‐resistant individuals? Petersen et al. measured lipogenesis in nation, the supposed “paradox” of selective hepatic insulin
lean, healthy, young individuals that were either insulin‐sensi- resistance reveals itself to be most likely the product of con-
tive or insulin‐resistant, as determined by oral glucose tolerance founding. When decoupled from the carbohydrate excess that
test [40]. These subjects were matched for multiple factors, often accompanies insulin resistance, hepatocyte insulin resist-
including age, BMI, adiposity, blood pressure, and activity. ance  –  produced by either brief high‐fat feeding or insulin
Fasting plasma glucose and insulin, while still within normal receptor knockdown  –  results in decreased rates of de novo
limits, were significantly higher in the insulin‐resistant cohort. ­lipogenesis as measured by stable isotope methods in rats [42].
In addition, plasma triglycerides were increased and plasma This finding reveals that de novo lipogenesis can become resist-
HDL decreased in the insulin‐resistant subjects. At baseline, ant to insulin, so why is de novo lipogenesis increased in insulin
hepatic triglyceride content, as assessed by 1H MRS, was nor- resistant humans with NAFLD? Interestingly, insulin is not nec-
mal (less than 1.0%) and similar between the two groups. After essary for induction of the de novo lipogenic program. Indeed,
a carbohydrate challenge, however, the insulin‐resistant group de novo lipogenesis is regulated by multiple nutritional inputs,
had marked elevations in plasma insulin but decreased muscle including glucose and fructose (through the carbohydrate
glycogen synthesis. Interestingly, the insulin‐resistant group response element binding protein [ChREBP]) and amino acids
exhibited a significant increase in liver triglyceride content, (through mTORC1). In the setting of chronic nutritional, and
with a concordant twofold increase in de novo lipogenesis. particularly glucose, excess that accompanies typical human
Thus, in these young insulin‐resistant individuals, the impaired insulin resistance, it is quite conceivable that nutritional inputs
non‐oxidative disposal of glucose as muscle glycogen led to could bypass hepatic insulin resistance to induce the de novo
increased de novo lipogenesis and greater hepatic triglyceride lipogenic program on their own.
content [40]. This suggests that a primary defect in muscle insu- Finally, it is worth emphasizing that the major pathway for
lin sensitivity enables increased hepatic de novo lipogenesis hepatic triglyceride synthesis is re‐esterification of circulating
from ingested carbohydrates, which over time could develop FAs, as described above. Re‐esterification is a substrate‐­
into NAFLD (Figure 37.2). dependent but insulin‐independent process [42]. So while
De novo lipogenesis is an insulin‐regulated process; insulin de novo lipogenesis may be a particularly important pathway for
induces the SREBP‐1c transcriptional program that upregulates the development of NAFLD, the quantitatively dominant source
multiple lipogenic genes. Yet, as described above, insulin of IHTG in NAFLD  –  re‐esterification  –  is not affected by

Insulin sensitive Insulin resistant

Hypertriglyceridemia

Glycogen De novo lipogenesis

Glucose

Glycogen Glycogen

Figure 37.2  Schematic of whole‐body energy distribution after high‐carbohydrate mixed meals in insulin‐sensitive and insulin‐resistant individu-
als. In insulin‐sensitive states, glucose is non‐oxidatively disposed as glycogen in liver and muscle. In insulin‐resistant states, entry of glucose into
the muscle is impaired, leading to increased delivery to the liver, where de novo lipogenesis ensues, leading to increased hepatic triglyceride accu-
mulation and secretion. Adapted from [40]. Reproduced with permission of the National Academy of Sciences.
464 THE LIVER:  MECHANISMS FOR HEPATIC INSULIN RESISTANCE IN NAFLD

hepatic insulin sensitivity or resistance. In light of current signaling. Thus, these results in the fatless mouse model also
knowledge, therefore, it does not seem necessary to invoke par- support the hypothesis that hepatic fat accumulation leads to
adoxical pathway‐selective hepatic insulin resistance to explain insulin resistance through a proximal block in the insulin signal-
the development of NAFLD or increased rates of de novo lipo- ing cascade.
genesis in insulin‐resistant subjects.
Acute high‐fat feeding in normal rats
In normal male Sprague–Dawley (SD) rats, high‐fat feeding for
MECHANISMS FOR HEPATIC INSULIN three days (3d HFF) results in the accumulation of IHTG, with-
RESISTANCE IN NAFLD out any significant change in body weight or muscle lipid con-
tent [47]. In 3d HFF rats, although plasma glucose was
The close association between peripheral insulin resistance and unchanged, plasma insulin concentrations nearly doubled, sug-
NAFLD can confound studies of NAFLD and hepatic insulin gesting insulin resistance. When tissue‐specific insulin action
resistance. Specifically, in studying the action of insulin in the was assessed in 3d HFF rats using the hyperinsulinemic–eugly-
liver, it may become difficult to determine which changes are cemic clamp, there were no differences in insulin‐stimulated,
due to peripheral insulin resistance and which are due solely to whole‐body glucose disposal or muscle‐ and adipose‐specific
hepatic steatosis. Thus, to understand how intrahepatic lipid glucose uptake. In contrast, the liver was completely insulin‐
renders the liver resistant to insulin, it is instructive to consider resistant in the 3d HFF rat, with insulin suppressing hepatic glu-
models in which hepatic steatosis is present without pronounced cose production by only around 10%, compared with nearly
adiposity and peripheral insulin resistance. 80% suppression of hepatic glucose production (HGP) in con-
trol chow‐fed rats. Interestingly, if hepatic fat accumulation was
prevented by promoting mitochondrial lipid oxidation with the
Liver‐specific LPL overexpressing mice
mitochondrial uncoupler 2,4‐dinitrophenol, hepatic insulin sen-
One model in which to study the mechanism for lipid‐induced sitivity was preserved. In the 3d HFF rat model, IHTG accumu-
hepatic insulin resistance is the liver‐specific LPL overexpress- lation was associated with proximal defects in insulin signaling,
ing mouse. LPL is the rate‐limiting enzyme for the hydrolysis with impaired tyrosine phosphorylation of IRS1 and IRS2 by
of triglyceride from triglyceride‐rich lipoproteins, such as chy- the insulin receptor ultimately impairing the ability of insulin to
lomicrons and very low‐density lipoproteins (VLDLs). By activate hepatic glycogen synthesis. Importantly, similar defects
crossing a transgenic mouse expressing human LPL under the in insulin‐stimulated hepatic glycogen synthesis and insulin‐
control of the apoA‐I promoter (A‐I LPL) with a heterozygote mediated suppression of hepatic glucose production have been
LPL knockout (het KO) mouse, it was possible to generate a in patients with T2D when studied under careful conditions with
mouse in which LPL was expressed primarily in the liver [43]. matched insulin concentrations [48].
The resulting liver LPL overexpressing mouse displays IHTG To identify the mechanism by which IHTG accumulation
accumulation. Using the hyperinsulinemic–euglycemic clamp impairs insulin signaling, possible involvement of the DAG–PKC
technique, insulin‐sensitivity of this mouse was assessed [44]. axis was examined. A screen for PKC activation in 3d HFF liver
While whole‐body glucose metabolism and insulin‐stimulated was performed. This revealed that hepatic steatosis and hepatic
muscle glucose uptake were normal, the ability of insulin to insulin resistance were most strongly associated with the activa-
suppress hepatic glucose production was markedly diminished, tion of PKCε, which, like PKCθ, is a DAG‐dependent, Ca2+‐inde-
with impaired hepatic insulin signaling to IRS2 and PI3 kinase. pendent novel PKC isoform (Figure 37.1) [47]. If hepatic steatosis
This model illustrates that hepatic lipid accumulation can lead was prevented with the use of 2,4, dinitrophenol, PKCε activation
specifically to hepatic insulin resistance through a block in the was also prevented. This association between DAG accumula-
insulin signaling cascade. tion, PKCε activation, and hepatic insulin resistance has since
been observed in dozens of rodent models [49].
Four studies to date have examined the potential involvement
Lipoatrophic mice
of the DAG–PKCε axis in hepatic insulin resistance in humans.
Another mouse model of hepatic fat accumulation and hepatic Kumashiro et  al. measured multiple putative mediators of
insulin resistance is the “fatless” lipoatrophic mouse. By hepatic insulin resistance in liver biopsy samples from a cohort
expressing a dominant‐negative protein, A‐ZIP/F, under the of obese, non‐diabetic subjects [50]. Hepatic DAG content was
control of the adipocyte‐specific AP‐1 promoter, Moitra et al. found to be the best predictor of insulin sensitivity as measured
generated mice that are devoid of white adipose tissue [45]. by the homeostatic model assessment (HOMA), and was
Without adipocytes, these mice store lipids in the muscle and strongly correlated with PKCε activation. Magkos et al. studied
liver. When studied using the hyperinsulinemic–euglycemic obese non‐diabetic subjects using suppression of hepatic glu-
clamp, these fatless mice displayed profound peripheral and cose production during hyperinsulinemic–euglycemic clamp
hepatic insulin resistance [46]. Analysis of insulin signaling in studies as the readout of hepatic insulin sensitivity, and observed
the liver revealed that insulin failed to increase IRS2‐associated that hepatic DAG, of several lipid species measured, was the
PI3 kinase activity in the fatless mice. Remarkably, this pheno- most strongly correlated with hepatic insulin resistance [51].
type was rescued by transplantation of fat pads from wild‐type Luukkonen et  al. studied adults undergoing bariatric surgery
littermates [46]. With restoration of adipocytes, tissue lipid and reported that four out of five DAG species measured were
concentrations normalize, as do muscle and hepatic insulin
­ correlated with HOMA [52]. Most recently, Ter Horst et  al.
37:  Non‐alcoholic Fatty Liver Disease and Insulin Resistance 465

measured insulin suppression of hepatic glucose production in insulin receptor Thr1160 as a substrate of PKCε [57]. Structural
obese non‐diabetic subjects and noted that subjects with hepatic studies revealed that owing to its position in the activation loop
insulin resistance displayed increased DAG in the cytosolic/ of the insulin receptor kinase domain, Thr1160 phosphorylation
lipid droplet fraction and increased PKCε activation compared was predicted to destabilize the active configuration of the
to subjects with preserved hepatic insulin sensitivity [53]. kinase, impairing INSR activity. These predictions were borne
The total hepatic DAG content does not always correlate out in studies of insulin receptors carrying the phosphomimetic
with the degree of hepatic insulin resistance in rodent models Thr1160Glu mutation, which were severely kinase‐defective.
[49]. The reason for this dissociation is not always apparent, To evaluate the significance of insulin receptor Thr1160 phos-
but a failure to measure individual DAG stereoisomers is likely phorylation for lipid‐induced hepatic insulin resistance, mice
to explain at least some discrepant models. The placement of carrying an alanine mutation at the homologous residue
two fatty acyl chains along the three‐carbon glycerol backbone (InsrT1150A mice) were studied. Mice were studied after an 8–10d
of DAG yields three possible stereoisomers: sn‐1,2, sn‐2,3, HFF protocol, which like the 3d HFF protocol in rats achieved
and sn‐1,3 DAG. PKC enzymes are only activated by sn‐1,2 liver‐specific lipid accumulation and insulin resistance in wild‐
DAG [54]. Interestingly, adipose triglyceride lipase (ATGL, or type mice. The InsrT1150A mice developed hepatic lipid accumu-
PNPLA2)  –  the major triglyceride lipase in the hepato- lation, including DAG, and exhibited PKCε translocation on
cyte  –  preferentially generates sn‐1,3 DAG when working this diet. However, the InsrT1150A mice were protected from
alone and preferentially generates sn‐1,3 and sn‐2,3 DAG hepatic insulin resistance, displaying improved suppression of
when working with its cofactor CGI‐58 [55]. The sn‐1,2 DAG hepatic glucose production, stimulation of net hepatic glycogen
moieties capable of activating PKC enzymes, including PKCε, synthesis, and activation of INSR activity compared to wild‐
are ­ therefore primarily generated through hydrolysis of type mice [57]. These studies demonstrated the necessity of
­phosphatidylinositol‐4,5‐bisphosphate (PIP2) in the plasma insulin receptor Thr1150 phosphorylation for high‐fat, diet‐
­membrane and through de novo glycerolipid synthesis in the induced, hepatic insulin resistance in mice, and revealed a
endoplasmic reticulum, not through lipolysis. Many rodent molecular mechanism by which hepatic lipid accumulation
models in which total hepatic DAG content has been dissoci- leads to impaired hepatic insulin signaling (Figure 37.1c, d).
ated from hepatic insulin resistance have genetic disruptions in INSR Thr1160 is conserved from fly to man. The identifica-
lipid synthesis or lipolysis, and the stereoisomer composition tion of a conserved molecular mechanism for lipid‐induced cel-
of the DAG pool within those hepatocytes thus cannot be lular insulin resistance raises the question of its evolutionary
assumed to approximate the normal state. This is a subject ripe utility. One attractive possibility is that lipid‐induced hepatic
for future investigation. insulin resistance is an adaptation to starvation. Prolonged fast-
ing in the rat (48 hours) induces hepatic DAG and TAG accumu-
lation and activates PKCε [58]. Induction of hepatic insulin
PKCε links hepatic steatosis to hepatic insulin resistance would impair glycogen storage upon refeeding,
resistance through insulin receptor kinase ­conserving glucose for essential tissues such as brain.
inhibition
The specific role of PKCε in the pathogenesis of hepatic insulin
resistance was assessed using antisense oligonucleotides FASTING HYPERGLYCEMIA AND
(ASOs) against PKCε [56]. ASOs are synthetic, modified GLUCONEOGENESIS
­oligonucleotides that are taken up preferentially in liver, adipose
tissue, and kidney, though not in other key tissues such as mus- Fasting hyperglycemia is a cardinal feature of T2D. The devel-
cle, brain, or β‐cells. They offer the opportunity to specifically opment of fasting hyperglycemia requires increased endoge-
inhibit any gene product of interest, enabling both target valida- nous glucose production (EGP), which is the sum of both
tion and potential therapeutic utility. After treatment with either gluconeogenesis and net glycogenolysis. Resolving the relative
saline, a control (random sequence) ASO, or PKCε ASO, nor- contributions of these two pathways, while not trivial, is critical
mal SD rats were subjected to the 3d HFF protocol and insulin to understanding the pathogenesis of T2D. This section will
action was assessed using the hyperinsulinemic–euglycemic review current knowledge regarding the measurement, regula-
clamp. Though hepatic lipid accumulation, and specifically tion, and significance of gluconeogenesis in the development
DAG accumulation, was equal in all groups, PKCε ASO treated of T2D.
rats displayed improved hepatic insulin sensitivity and insulin
signaling compared to both control groups. Mechanistically, Hepatic gluconeogenesis in fed
this was associated with improved activation of the insulin
receptor itself. Whereas high‐fat feeding impaired liver insulin
and fasted states
receptor (INSR) activation compared to control chow‐fed rats, Gluconeogenesis contributes substantially to EGP in both the
INSR tyrosine kinase activity was preserved in HFF rats treated fed and fasted states in the normal human. This was demon-
with PKCε ASO. These studies suggested that PKCε mediated strated in several studies by employing 13C MRS measurements
lipid‐induced hepatic insulin resistance through inhibition of the of hepatic glycogen [59–61]. After a meal, liver glycogen
insulin receptor. ­concentration increases in a near linear fashion, peaking at
To identify the molecular mechanism of this inhibition, phos- around 5 hours [59]. During this period of hepatic glycogen syn-
phopeptide mass spectrometry was employed and revealed thesis, when ingested glucose is absorbed from the gut, EGP is
466 THE LIVER:  FASTING HYPERGLYCEMIA AND GLUCONEOGENESIS

normally suppressed. The rate of glycogenolysis can be meas- impaired hepatic insulin action, secondary to hepatic insulin
ured using repeated 13C MRS measures of hepatic glycogen dur- resistance and β‐cell dysfunction, glucagon likely promotes glu-
ing this time. By subtracting the rate of glycogenolysis from the coneogenesis and fasting hyperglycemia.
rate of EGP, the rate of gluconeogenesis is calculated. About 5 These two models have in common a central role for direct
hours after a meal, hepatic glycogen is utilized in a linear rate up hepatic effects of insulin and glucagon. Insulin regulates gluco-
to about 22 hours [60, 61]. During this period, hepatic glycogen- neogenesis through chronic transcriptional mechanisms, while
olysis accounts for only 40–50% of total EGP. Gluconeogenesis glucagon utilizes both chronic transcriptional and acute post‐
accounts for the remainder, or 50–60% of EGP during the early translational modes of regulation [72]. Insulin can suppress
part of a fast. Between 22 and 46 hours, net hepatic glycogen- transcription by phosphorylating and inactivating key transcrip-
olysis still accounts for around 20% and gluconeogenesis around tion factors of the FOXO family, especially FOXO1. Insulin‐
80% of EGP [60]. With continued fasting up to 46–64 hours, net stimulated phosphorylation of FOXO transcription factors by
hepatic glycogenolysis diminishes to less than 5%, while the AKT2 excludes them from the nucleus and prevents them from
remaining 95% of EGP results from gluconeogenesis [60]. joining with the transcriptional coactivator PPARγ coactivator
1α (PGC‐1α) to activate transcription of the gluconeogenic
Gluconeogenesis is increased in T2DM enzymes phosphoenolpyruvate carboxykinase (PCK1) and glu-
cose‐6‐phosphatase (G6PC) [73]. The FOXO module appears to
Magnusson et al., using the 13C MRS technique, determined the be particularly critical for upregulating G6PC transcription
relative contributions of gluconeogenesis in patients with T2D ­during long periods of fasting (around 18 hours in mice) [74]. A
[62]. Rates of glycogenolysis and EGP were determined in con- second major mode of transcriptional regulation of gluconeo-
trol and diabetic subjects fasted over 23 hours. The rate of EGP genesis by insulin involves the cyclic AMP‐responsive element
was increased by around 25% in the diabetic subjects, and this binding protein (CREB) and its regulators CREB‐binding pro-
increase was entirely accounted for by an approximate 60% tein (CBP) and CREB‐regulated transcriptional coactivator 2
increase in the contribution of gluconeogenesis to EGP [62]. (CRTC2) [75]. The CREB‐CBP‐CRTC2 module upregulates
These findings have been confirmed by other techniques as PCK1, G6PC, and PGC‐1α, and appears to be operative in the
well. Wajngot et al. used the 2H2O method to measure rates of earlier stages of a fast (around 0–6 hours in mice) [74]. Insulin
gluconeogenesis in normal and diabetic subjects between 15 phosphorylates CRTC2, leading to its nuclear exclusion [76].
and 22 hours of fasting [63]. They found that the percentage Thus, the overall direct effect of insulin on hepatic gluconeo-
contribution of gluconeogenesis increased from 57.6 ± 2.5% to genesis centers on deactivating transcription of gluconeogenic
70.6 ± 2.6% in normal subjects, and from 62.5 ± 2.8% to 75.6 ± genes, a slow process.
3.1% in subjects with T2DM [63]. Glucagon also acts through transcriptional regulation, but in
addition acts quickly through phosphorylation of several key
From insulin resistance to fasting enzymes of glucose metabolism. Glucagon increases adenylate
cyclase activity and activates protein kinase A (PKA). PKA then
hyperglycemia phosphorylates liver‐type pyruvate kinase (L‐PK) and the
Hepatic steatosis alone will not increase hepatic gluconeogene- bifunctional enzyme phosphofructokinase‐2/fructose‐2,6‐bis-
sis. In rodents, hepatic steatosis, while associated with hepatic phosphatase (PFK‐2/FBPase‐2). The net effect is to favor gluco-
insulin resistance, is insufficient for increased fasting EGP or neogenesis over glycolysis [72]. PKA also phosphorylates
hyperglycemia [47, 56, 57]. Similarly, in humans the presence CREB and leads to the dephosphorylation of CRTC2, activating
of hepatic steatosis is associated with hepatic insulin resistance the CREB‐CBP‐CRTC2 transcriptional module to promote
[50, 51, 53], but not with increased rates of fasting EGP, as ­gluconeogenic gene expression [75].
is  seen in T2D [64]. Additional changes beyond the develop- As described above, insulin and glucagon exert powerful
ment of NAFLD are clearly required to permit fasting effects on the expression of the key gluconeogenic genes PCK1
hyperglycemia. and G6PC. But what is the pathophysiological significance of
Several models have been proposed to describe the progres- these effects? The hypothesis that increased transcription of glu-
sion from hepatic insulin resistance to overt T2D. A popular, coneogenic enzymes accounts for the increased gluconeogene-
straightforward model holds that hyperglycemia develops only sis was tested in two rodent models of T2D and in human
after pancreatic β‐cell dysfunction leads to waning plasma subjects with T2D [77]. The first T2D rat model was induced by
­insulin concentrations [65]. An alternative model, the Unger 3d HFF in combination with nicotinamide and streptozotocin to
“bi‐hormonal hypothesis” for the pathogenesis of hyperglyce- induce β‐cell dysfunction (STZ/HFF). In this model of T2D, the
mia, holds glucagon equally culpable as insulin [66]. Glucagon, rats had modest fasting hyperglycemia and increased rates of
even at basal levels, potently sustains hepatic glucose EGP. However, there was no increase in the expression of PCK1
­production. In canines, with glucose and insulin held constant, or G6PC [77]. The second T2D rat model sought to decouple 3d
suppression of basal glucagon reverses net hepatic glucose pro- HFF‐induced hepatic insulin resistance from compensatory
duction, instead causing net hepatic glucose uptake [67]. In hyperinsulinemia by employing somatostatin to pause endo-
humans, basal glucagon activates gluconeogenesis and blunts crine pancreatic secretion. Under these conditions, infusion of
insulin‐stimulated glycogen synthesis [68]. Moreover, patients insulin and glucagon into the portal vein at basal replacement
with diabetes have an inappropriately high glucagon level rela- rates resulted in hyperglycemia in the HFF rats compared to
tive to their degree of glycemia [69–71]. Thus, in the setting of chow‐fed controls. Though there were clear increases in EGP
37:  Non‐alcoholic Fatty Liver Disease and Insulin Resistance 467

and gluconeogenesis in HFF rats, PCK1 and G6PC expression than nondiabetic controls [84]. Similarly, subjects with T2D dis-
were equivalent in HFF liver compared to control liver [77]. play higher rates of glycerol turnover and gluconeogenesis from
Finally, PCK1 and G6PC mRNAs were quantitated from liver glycerol than nondiabetic controls [85]. Importantly, increased
biopsy samples obtained from normoglycemic (control) and plasma FA concentrations are a risk factor for incident T2D
hyperglycemic untreated T2D patients undergoing bariatric sur- [86]. It is tempting to speculate that a key mechanism whereby
gery. The expression of PCK1 and G6PC was not increased in obesity‐induced adipose tissue inflammation and insulin resist-
the T2D subjects [77]. Other studies have also observed no ance promote T2D is by driving EGP. But studies directly
increase in PCK1 and G6PC transcription in insulin resistant rat addressing the role of adipose lipolysis in the transition to overt
liver [78, 79]. Interestingly, in G6PC knockout mice, re‐expression T2D are currently lacking. With regard to the second question,
of G6PC to only around 20% of wild‐type levels is sufficient to what role hepatocellular insulin resistance and the DAG–PKCε–
rescue hypoglycemia, suggesting that modest oscillations in INSR axis might play in fasting hyperglycemia, it is prudent to
gluconeogenic gene transcription may exert limited effects on note the limitations of the rodent studies that have emphasized
gluconeogenic flux [80]. In summary, these studies provide the importance of indirect insulin action for EGP suppression.
­evidence against the hypothesis that increased transcription of Specifically, the use of overnight‐fasted rodents in these studies
gluconeogenic enzymes is responsible for the increased gluco- minimized the contribution of glycogenolysis to EGP and maxi-
neogenesis and EGP observed in patients with T2D. Instead, mized the contribution of gluconeogenesis. However, as dis-
these data suggest that other mechanisms are responsible for cussed above, glycogenolysis is a physiologically significant
these abnormalities. contributor to EGP in humans during the first 24 hours of a fast
If transcriptional effects are insufficient to account for the [60]. Hepatocellular insulin resistance potently affects hepatic
transition from hepatic insulin resistance to fasting hyperglyce- glycogen metabolism [57], and insulin‐stimulated hepatic gly-
mia, what other mechanisms might be at work? In this context, cogen deposition is impaired in T2D [48]. Thus, for the two
it is useful to consider other regulators of EGP, especially sub- major components of EGP, a reasonable simplification may be
strate availability and allosteric modulators of gluconeogenesis to state that direct hepatocellular effects dominate insulin regu-
[72]. The potently insulin‐regulated process of adipose tissue lation of hepatic glycogen metabolism, while indirect effects
lipolysis yields metabolites that act through both mecha- dominate insulin regulation of gluconeogenesis (Figure 37.3).
nisms  –  substrate push and allosteric activation  –  to promote
gluconeogenesis [78]. Lipolysis yields glycerol, which is con-
verted to glucose at rates proportional to its availability [81]. It
also yields FAs, which upon β‐oxidation in the hepatocellular INSULIN SENSITIZING AGENTS
mitochondrion generate acetyl‐CoA, a potent allosteric activa- AND HEPATIC INSULIN RESISTANCE
tor of the key gluconeogenic enzyme pyruvate carboxylase [78].
In fasted rats, insulin infusion suppressed lipolysis and therefore
Metformin
decreased turnover of glycerol and FAs; this was accompanied
by reductions in hepatic acetyl‐CoA and EGP [78]. Remarkably, Metformin, a biguanide, has been used to treat patients with T2D
this insulin suppression of EGP could be totally prevented by for over 30 years and improves liver function test abnormalities,
co‐infusion of glycerol and acetate at rates calibrated to prevent at least temporarily, in some patients with NAFLD. Since hepatic
insulin‐induced suppression of glycerol turnover and hepatic glucose production is the sum of gluconeogenesis and glycogen-
acetyl‐CoA concentration, respectively [78]. These studies were olysis, it is important to know which pathway is affected by met-
consistent with multiple dog and human studies carried out in formin. Using [6,6−2H] glucose measures of EGP combined with
the 1990s that pointed to the insulin concentration in the periph- 13
C MRS measures of net hepatic glycogenolysis and 2H2O
ery, rather than in the portal vein, as the major controller of EGP measures of gluconeogenesis, Hundal et al. studied nine diabetic
[72]. Indeed, they indicated that, at least in the fasted rodent subjects before and after three months of metformin therapy and
dependent on gluconeogenesis for nearly all of EGP, hepatocel- compared the results with seven age‐, sex‐, and weight‐matched
lular insulin signaling was dispensable for insulin suppression controls [87]. At baseline, the diabetic subjects had increased
of EGP. Multiple rodent models of severely disrupted hepatocel- rates of EGP compared to non‐diabetic control subjects. Using
lular insulin signaling that retain the ability to suppress EGP 13
C MRS measures of net hepatic glycogenolysis and 2H2O
upon insulin administration have lent dramatic credence to this measures of gluconeogenesis, it was clear that increased hepatic
hypothesis [78, 82, 83]. gluconeogenesis accounted for the increase in EGP, consistent
The realization that control of EGP has a large indirect com- with previous studies [62]. After three months of metformin
ponent raises two interesting questions: one, whether increased ­therapy, EGP was lowered in the diabetic subjects, due to a 36%
lipolysis mediates the transition to fasting hyperglycemia that reduction in hepatic gluconeogenesis [87].
marks overt T2D; and two, whether hepatocellular insulin resist- Though metformin clearly reduces fasting glycemia and it is
ance even contributes to the increased EGP of T2D. These ques- widely accepted that this occurs through decreased gluconeo-
tions remain incompletely answered but merit some comment. genesis, the cellular mechanisms for the effect of metformin
With regard to the first question, some evidence supports a role remain debated. A recent study by Madiraju et al. is unique in
of lipolysis in T2D pathophysiology. Poorly controlled diabetic identifying both a specific molecular interaction for metformin
subjects (increased EGP with fasting glucose >250 mg dl−1) and a physiological basis for the rare metformin adverse effect
exhibit higher plasma FA concentrations throughout the day of lactic acidosis [88]. Madiraju et al. observed non‐competitive
468 THE LIVER:  INSULIN SENSITIZING AGENTS AND HEPATIC INSULIN RESISTANCE

Adipose insulin resistance


inflammation Hepatic insulin resistance
Insulin secretion

Hepatocellular insulin signaling


Adipose lipolysis

Acetyl CoA
Glycerol turnover

Gluconeogenesis
Net glycogenolysis

Hepatic glucose production

Figure 37.3  Effect of adipose and hepatic insulin resistance on hepatic glucose production. Hepatic glucose production is controlled through both
direct and indirect mechanisms. Direct effects are mediated through the hepatocellular insulin receptor and, in the acute setting, are mediated largely
through promotion of glycogen storage. Indirect effects include, but are not limited to, inhibition of adipocyte lipolysis which decreases whole‐body
glycerol and fatty acid turnover. This in turn decreases allosteric activation of gluconeogenesis by hepatic mitochondrial acetyl‐CoA and substrate‐
driven gluconeogenesis from glycerol. A prediction that follows from this model is that hepatocellular insulin resistance mainly affects the glycog-
enolytic component of hepatic glucose production, while adipocyte insulin resistance (often associated with inflammation) mainly  affects the
gluconeogenic component of hepatic glucose production.

inhibition of the redox shuttle enzyme mitochondrial glycer- Thiazolidinediones


ophosphate dehydrogenase (mGPD, or GPD2) by metformin
at physiologically relevant concentrations, an effect that is pre- The thiazolidinediones (TZDs) are PPARγ agonists that improve
dicted to cause an increase in the cytoplasmic redox state and insulin sensitivity and have been used in the treatment of T2D.
a  reciprocal decrease in the mitochondrial redox state of the Both TZDs currently approved for use in the United States,
hepatocyte. These redox effects were indeed observed with rosiglitazone and pioglitazone, have been shown to improve
­metformin treatment in rats, and resulted in inhibition of gluco- hepatic insulin sensitivity in human studies [93, 94].
neogenesis from redox‐dependent substrates (i.e. lactate and The ability of the TZDs to improve hepatic insulin action is
glycerol) but not from redox‐independent substrates (i.e. pyru- primarily due to their extrahepatic effects. PPARγ is mainly
vate and alanine) in cultured hepatocytes. Rats treated with expressed in adipocytes with lower expression in the liver and
­antisense oligonucleotides targeting mGPD, as well as mGPD skeletal muscle, though there are species differences in expres-
knockout mice, phenocopied metformin‐treated animals and did sion. Given this discordance between the site of PPARγ expres-
not further suppress EGP upon metformin treatment. The role of sion and the site of drug effects, it was hypothesized that TZDs
mGPD inhibition in metformin‐treated humans remains to be redistribute fat from the liver and muscle into the adipocyte
investigated. [95]. Mayerson et al. tested this hypothesis, giving rosiglitazone
Other hypotheses for metformin action continue to be inves- to nine patients with T2D in order to assess changes in insulin
tigated as well. Metformin has been found to inhibit mitochon- sensitivity (as assessed by the hyperinsulinemic–euglycemic
drial complex I [89], though this may occur only at clamp) and tissue fat content (as measured by 1H MRS) [96].
supraphysiologic concentrations [88]. Inhibition of the respira- Insulin‐mediated suppression of lipolysis in subcutaneous fat
tory chain would be predicted to increase the cellular [AMP] : was also assessed using microdialysis to measure glycerol
[ATP] ratio, and indeed, metformin stimulates the AMP‐acti- release. After three months of therapy, rosiglitazone therapy was
vated kinase (AMPK) [88]. However, AMPK knockout mice associated with around 40% decrease in hepatic triglyceride
remain responsive to metformin treatment [90] and pharmaco- content, around 40% increase in extramyocellular triglyceride
logical activation of AMPK to the same extent as metformin concentration, and improved suppression of adipocyte lipolysis.
was insufficient to phenocopy metformin’s suppression of EGP Though there were no detectable decreases in intramyocellular
[88]. Metformin‐induced increases in the [AMP] : [ATP] ratio triglyceride in this study, rosiglitazone did improve insulin‐
have also been hypothesized to suppress EGP by antagonizing mediated whole‐body glucose disposal. This apparent discon-
the cAMP‐mediated effects of glucagon [91]; however, the nect between intramyocellular triglyceride and peripheral
physiological relevance of this mechanism was challenged by a insulin action underscores the fact that intramyocellular triglyc-
human study in which metformin did not inhibit glucagon‐­ eride is only a crude marker for the active metabolite (putatively
stimulated increases in EGP [92]. DAG) responsible for lipid‐induced insulin resistance [97].
37:  Non‐alcoholic Fatty Liver Disease and Insulin Resistance 469

In  summary, TZDs exert their beneficial effects by shifting 2. Gregg, E.W., Li, Y., Wang, J. et al. Changes in diabetes‐related complications
intracellular lipid from the liver and muscle into the adipose in the United States, 1990–2010. N Engl J Med, 2014;370(16):1514–23.
3. Le, M.H., Devaki, P., Ha, N.B. et al. Prevalence of non‐alcoholic fatty liver
­tissue, via PPARγ‐mediated improvements in adipose insulin disease and risk factors for advanced fibrosis and mortality in the United
sensitivity. States. PloS One, 2017;12(3):e0173499.
4. Tilg, H., Moschen, A.R., and Roden, M. NAFLD and diabetes mellitus.
Nat Rev Gastroenterol Hepatol, 2016;14(1):147.
Weight loss 5. Pais, R., Barritt, A.S., Calmus, Y. et  al. NAFLD and liver transplantation:
Current burden and expected challenges. J Hepatol, 2016;65(6):1245–57.
Weight loss reduces intrahepatic fat content and improves 6. Shulman, G.I. Ectopic fat in insulin resistance, dyslipidemia, and cardio-
hepatic insulin sensitivity in humans. Petersen et  al. studied metabolic disease. N Engl J Med, 2014;371(12):1131–41.
eight obese subjects before and after weight loss induced by a 7. Randle, P.J., Garland, P.B., Hales, C.N. et al. The glucose fatty‐acid cycle. Its
hypocaloric, low‐fat diet [98]. Patients were kept on a diet until role in insulin sensitivity and the metabolic disturbances of diabetes mellitus.
Lancet, 1963;1(7285); 785–9.
their fasting plasma glucose levels had stabilized. This required
8. Jucker, B.M., Rennings, A.J., Cline, G.W. et al. 13C and 31P NMR studies
3–12 weeks, with an average weight loss of around 8 kg. This on the effects of increased plasma free fatty acids on intramuscular glucose
only led to a modest change in BMI, with subjects dropping metabolism in the awake rat. J Biol Chem, 1997;272(16):10464–73.
from 30 to 28 kg m−2. Thus, even after weight loss, subjects 9. Randle, P.J. Regulatory interactions between lipids and carbohydrates: the glu-
remained overweight. A comparison was made to lean, seden- cose fatty acid cycle after 35 years. Diabetes Metab Rev, 1998;14(4):263–83.
10. Krssak, M., Falk Petersen, K., Dresner, A. et al. Intramyocellular lipid con-
tary control subjects without diabetes, who had an average BMI centrations are correlated with insulin sensitivity in humans: a 1H NMR
of 24 kg m−2. Weight loss did not alter IMCL content, which was spectroscopy study. Diabetologia, 1999;42(1):113–6.
still elevated compared to the lean controls. However, intrahe- 11. Perseghin, G., Scifo, P., De Cobelli, F. et  al. Intramyocellular triglyceride
patic triglyceride content dropped substantially from around content is a determinant of in vivo insulin resistance in humans: a 1H‐13C
12% to 2%, approaching the less than 1% values seen in the lean nuclear magnetic resonance spectroscopy assessment in offspring of type 2
diabetic parents. Diabetes, 1999;48(8):1600–6.
controls. Weight loss and the reduction in intrahepatic triglycer- 12. Rothman, D.L., Shulman, R.G., and Shulman, G.I. 31P nuclear magnetic
ide decreased the rate of fasting EGP and improved hepatic resonance measurements of muscle glucose‐6‐phosphate. Evidence for
insulin sensitivity (as assessed by hyperinsulinemic–euglyce- reduced insulin‐dependent muscle glucose transport or phosphorylation
mic clamp) to near normal. Thus, in this and other studies [99– activity in non‐insulin‐dependent diabetes mellitus. J Clin Invest,
1992;89(4):1069–75.
101], modest weight loss can potently correct hepatic steatosis
13. Rothman, D.L., Magnusson, I., Cline, G. et  al. Decreased muscle glucose
and restore hepatic insulin sensitivity. Finally, the DiRECT trial, transport/phosphorylation is an early defect in the pathogenesis of non‐­
which compared an intensive weight loss intervention using insulin‐dependent diabetes mellitus. Proc Natl Acad Sci USA,
hypocaloric meal replacement shakes to control usual care, 1995;92(4):983–7.
extended these findings in overweight adults with type 2 diabe- 14. Cline, G.W., Petersen, K.F., Krssak, M. et al. Impaired glucose transport as a
cause of decreased insulin‐stimulated muscle glycogen synthesis in type 2
tes not on insulin [102]. A subset of this trial cohort underwent
diabetes. N Engl J Med, 1999;341(4):240–6.
measurements of liver fat content, and a large decrease (from 15. Roden, M., Price, T.B., Perseghin, G. et al. Mechanism of free fatty acid‐
around 16% to 3%) in liver fat content was observed in the induced insulin resistance in humans. J Clin Invest, 1996;97(12):2859–65.
weight loss cohort [103]. These findings underscore the revers- 16. Dresner, A., Laurent, D., Marcucci, M. et  al. Effects of free fatty acids
ibility of NAFLD with weight loss. on  glucose transport and IRS‐1‐associated phosphatidylinositol 3‐kinase
­activity. J Clin Invest, 1999;103(2):253–9.
As of this writing, no reports have determined whether exer- 17. Rahimi, Y., Camporez, J.‐P.G., Petersen, M.C. et  al. Genetic activation of
cise, independent of weight loss, can improve hepatic steatosis pyruvate dehydrogenase alters oxidative substrate selection to induce skeletal
and hepatic insulin sensitivity. Several trials have tested exercise muscle insulin resistance. Proc Natl Acad Sci USA, 2014;111(46):16508–13.
regimens of varying intensity and duration and demonstrated 18. Pollare, T., Vessby, B., and Lithell, H. Lipoprotein lipase activity in skeletal
beneficial effects on intrahepatic triglyceride and metabolic muscle is related to insulin sensitivity. Arterioscler Thromb J Vasc Biol,
1991;11(5):1192–203.
parameters, but also on body weight [104–106]. However, in 19. Jang, C., Oh, S.F., Wada, S. et al. A branched‐chain amino acid metabolite
one cross‐sectional study, physical activity was inversely corre- drives vascular fatty acid transport and causes insulin resistance. Nat Med,
lated with intrahepatic triglyceride content, even after control- 2016;22(4):421–6.
ling for other key variables such as age, sex, BMI, HOMA, and 20. Lynch, C.J., and Adams, S.H. Branched‐chain amino acids in metabolic
­signalling and insulin resistance. Nat Rev Endocrinol, 2014;10(12):723–36.
adiponectin [107]. It is tantalizing to hypothesize that exercise,
21. Petersen, K.F., Befroy, D., Dufour, S. et al. Mitochondrial dysfunction in the
by reversing muscle insulin resistance [108], would allow more elderly: possible role in insulin resistance. Science, 2003;300 5622;, 1140–2.
glucose to be taken up into the muscle, with resultant decreases 22. Petersen, K.F., Dufour, S., Befroy, D. et al. Impaired mitochondrial activity
in de novo lipogenesis and hepatic fat content. However, more in the insulin‐resistant offspring of patients with type 2 diabetes. N Engl J
needs to be done to determine whether this hypothesis is valid, Med, 2004;350(7):664–71.
23. Lee, H.‐Y., Choi, C.S., Birkenfeld, A.L. et al. Targeted expression of catalase
and more importantly, whether such lifestyle interventions can
to mitochondria prevents age‐associated reductions in mitochondrial
be implemented and adopted by an ever‐increasing at‐risk ­function and insulin resistance. Cell Metab, 2010;12(6):668–74.
population. 24. Camporez, J.P., Wang, Y., Faarkrog, K. et al. Mechanism by which arylamine
N‐acetyltransferase 1 ablation causes insulin resistance in mice. Proc Natl
Acad Sci USA, 2017;114(52):E11285–92.
25. Knowles, J.W., Xie, W., Zhang, Z. et  al. Identification and validation of
N‐acetyltransferase 2 as an insulin sensitivity gene. J Clin Invest,
REFERENCES 2015;125(4):1739–51.
26. Rusu, V., Hoch, E., Mercader, J.M. et al. Type 2 Diabetes Variants Disrupt
1. International Diabetes Federation. IDF Diabetes Atlas, 2017, International Function of SLC16A11 through Two Distinct Mechanisms. Cell,
Diabetes Federation. 2017;170(1):199–212.
470 THE LIVER:  REFERENCES

27. Fink, R.I., Kolterman, O.G., Griffin, J. et  al. Mechanisms of insulin 52. Luukkonen, P.K., Zhou, Y., Sädevirta, S. et al. Hepatic ceramides dissociate
­resistance in aging. J Clin Invest, 1983;71(6):1523–35. steatosis and insulin resistance in patients with non‐alcoholic fatty liver
28. Rowe, J.W., Minaker, K.L., Pallotta, J.A. et  al. Characterization of the ­disease. J Hepatol, 2016;64(5):1167–75.
­insulin resistance of aging. J Clin Invest, 1983;71(6):1581–7. 53. Ter Horst, K.W., Gilijamse, P.W., Versteeg, R.I. et al. Hepatic Diacylglycerol‐
29. Yu, C., Chen, Y., Cline, G.W. et al. Mechanism by which fatty acids inhibit Associated Protein Kinase Cε Translocation Links Hepatic Steatosis to
insulin activation of insulin receptor substrate‐1(IRS‐1)‐associated Hepatic Insulin Resistance in Humans. Cell Rep, 2017;19(10):1997–2004.
phosphatidylinositol 3‐kinase activity in muscle. J Biol Chem,
­ 54. Rando, R.R. and Young, N. The stereospecific activation of protein kinase C.
2002;277(52):50230–6. Biochem Biophys Res Commun, 1984;122(2):818–23.
30. Kim, J.K., Fillmore, J.J., Sunshine, M.J. et  al. PKC‐theta knockout mice 55. Eichmann, T.O., Kumari, M., Haas, J.T. et al. Studies on the substrate and
are  protected from fat‐induced insulin resistance. J Clin Invest, stereo/regioselectivity of adipose triglyceride lipase, hormone‐sensitive
2004;114(6):823–7. lipase, and diacylglycerol‐O‐acyltransferases. J Biol Chem, 2012;
31. Li, Y., Soos, T.J., Li, X. et al. Protein kinase C θ inhibits insulin signaling 287(49):41446–57.
by phosphorylating IRS1 at Ser1101. J Biol Chem, 2004;279(44):45304–7. 56. Samuel, V.T., Liu, Z.‐X., Wang, A. et al. Inhibition of protein kinase Cepsilon
32. Petersen, M.C., Camporez, J.P.G., and Shulman, G.I. IRS1 Ser1101 prevents hepatic insulin resistance in nonalcoholic fatty liver disease. J Clin
phosphorylation impairs insulin‐stimulated muscle glucose metabolism.
­ Invest, 2007;117(3):739–45.
Diabetes, 2017;66(1):A37. 57. Petersen, M.C., Madiraju, A.K., Gassaway, B.M. et  al. Insulin receptor
33. Itani, S.I., Pories, W.J., Macdonald, K.G. et  al. Increased protein kinase Thr1160 phosphorylation mediates lipid‐induced hepatic insulin resistance.
C  theta in skeletal muscle of diabetic patients. Metabolism, J Clin Invest, 2016;126(11):4361–71.
2001;50(5):553–7. 58. Perry, R.J., Wang, Y., Cline, G.W. et  al. Leptin mediates a glucose‐fatty
34. Szendroedi, J., Yoshimura, T., Phielix, E. et  al. Role of diacylglycerol acid  cycle to maintain glucose homeostasis in starvation. Cell,
­activation of PKCθ in lipid‐induced muscle insulin resistance in humans. 2018;172(1):234–48.
Proc Natl Acad Sci USA, 2014;111(26):9597–602. 59. Taylor, R., Magnusson, I., Rothman, D.L. et al. Direct assessment of liver
35. Nowotny, B., Zahiragic, L., Krog, D. et al. Mechanisms underlying the onset glycogen storage by 13C nuclear magnetic resonance spectroscopy and
of oral lipid‐induced skeletal muscle insulin resistance in humans. Diabetes, ­regulation of glucose homeostasis after a mixed meal in normal subjects.
2013;62(7):2240–8. J  Clin Invest, 1996;97(1):126–32.
36. Shulman, G.I., Rothman, D.L., Jue, T. et al. Quantitation of muscle glycogen 60. Rothman, D.L., Magnusson, I., Katz, L.D. et  al. Quantitation of hepatic
synthesis in normal subjects and subjects with non‐insulin‐dependent ­glycogenolysis and gluconeogenesis in fasting humans with 13C NMR.
­diabetes by 13C nuclear magnetic resonance spectroscopy. N Engl J Med, Science, 1991;254 (5031):573–6.
1990;322(4):223–8. 61. Petersen, K.F., Price, T., Cline, G.W. et  al. Contribution of net hepatic
37. Donnelly, K.L., Smith, C.I., Schwarzenberg, S.J. et al. Sources of fatty acids ­glycogenolysis to glucose production during the early postprandial period.
stored in liver and secreted via lipoproteins in patients with nonalcoholic Am J Physiol, 1996;270(1 Pt 1):E186–91.
fatty liver disease. J Clin Invest, 2005;115(5):1343–51. 62. Magnusson, I., Rothman, D.L., Katz, L.D. et al. Increased rate of gluconeo-
38. Diraison, F., Moulin, P., and Beylot, M. Contribution of hepatic de novo genesis in type II diabetes mellitus. A 13C nuclear magnetic resonance study.
lipogenesis and reesterification of plasma non esterified fatty acids to plasma J Clin Invest, 1992;90(4):1323–7.
triglyceride synthesis during non‐alcoholic fatty liver disease. Diabetes 63. Wajngot, A., Chandramouli, V., Schumann, W.C. et al. Quantitative contribu-
Metab, 2003;29(5):478–85. tions of gluconeogenesis to glucose production during fasting in type 2
39. Lambert, J.E., Ramos‐Roman, M.A., Browning, J.D. et al. Increased de novo ­diabetes mellitus. Metabolism, 2001;50(1):47–52.
lipogenesis is a distinct characteristic of individuals with nonalcoholic fatty 64. Marchesini, G., Brizi, M., Bianchi, G. et al. Nonalcoholic fatty liver disease:
liver disease. Gastroenterology, 2014;146(3):726–35. a feature of the metabolic syndrome. Diabetes, 2001;50(8):1844–50.
40. Petersen, K.F., Dufour, S., Savage, D.B. et al. The role of skeletal muscle 65. Kahn, S.E. The relative contributions of insulin resistance and beta‐cell
insulin resistance in the pathogenesis of the metabolic syndrome. Proc Natl dysfunction to the pathophysiology of Type 2 diabetes. Diabetologia,
­
Acad Sci USA, 2007;104(31):12587–94. 2003;46(1):3–19.
41. Brown, M.S. and Goldstein, J.L. Selective versus total insulin resistance: a 66. Lee, Y.H., Wang, M.‐Y., Yu, X.‐X. et al. Glucagon is the key factor in the
pathogenic paradox. Cell Metab, 2008;7(2):95–6. development of diabetes. Diabetologia, 2016;59(7):1372–5.
42. Vatner, D.F., Majumdar, S.K., Kumashiro, N. et  al. Insulin‐independent 67. Shulman, G.I., Liljenquist, J.E., Williams, P.E. et al. Glucose disposal during
regulation of hepatic triglyceride synthesis by fatty acids. Proc Natl Acad Sci insulinopenia in somatostatin‐treated dogs. The roles of glucose and
USA, 2015;112(4):1143–8. ­glucagon. J Clin Invest, 1978;62(2):487–91.
43. Merkel, M., Weinstock, P.H., Chajek‐Shaul, T. et  al. Lipoprotein lipase 68. Roden, M., Perseghin, G., Petersen, K.F. et al. The roles of insulin and gluca-
expression exclusively in liver. A mouse model for metabolism in the neona- gon in the regulation of hepatic glycogen synthesis and turnover in humans.
tal period and during cachexia. J Clin Invest, 1998;102(5):893–901. J Clin Invest, 1996;97(3):642–8.
44. Kim, J.K., Fillmore, J.J., Chen, Y. et  al. Tissue‐specific overexpression of 69. Reaven, G.M., Chen, Y.D., Golay, A. et al. Documentation of hypergluca-
lipoprotein lipase causes tissue‐specific insulin resistance. Proc Natl Acad gonemia throughout the day in nonobese and obese patients with noninsulin‐
Sci USA, 2001;98(13):7522–7. dependent diabetes mellitus. J Clin Endocrinol Metab, 1987;64(1):106–10.
45. Moitra, J., Mason, M.M., Olive, M. et al. Life without white fat: a transgenic 70. Raskin, P. and Unger, R.H. Hyperglucagonemia and its suppression.
mouse. Genes Dev, 1998;12(20):3168–81. Importance in the metabolic control of diabetes. N Engl J Med,
46. Kim, J.K., Gavrilova, O., Chen, Y. et al. Mechanism of insulin resistance in 1978;299(9):433–6.
A‐ZIP/F‐1 fatless mice. J Biol Chem, 2000;275(12):8456–60. 71. Basu, R., Schwenk, W.F., and Rizza, R.A. Both fasting glucose production
47. Samuel, V.T., Liu, Z.‐X., Qu, X. et  al. Mechanism of hepatic insulin and disappearance are abnormal in people with “mild” and “severe” type 2
resistance in non‐alcoholic fatty liver disease. J Biol Chem, 2004;
­ diabetes. Am J Physiol Endocrinol Metab, 2004;287(1):E55–62.
279(31):32345–53. 72. Petersen, M.C., Vatner, D.F., and Shulman, G.I. Regulation of hepatic
48. Krssak, M., Brehm, A., Bernroider, E. et  al. Alterations in postprandial glucose metabolism in health and disease. Nat Rev Endocrinol,
­
hepatic glycogen metabolism in type 2 diabetes. Diabetes, 2004; 2017;13(10):572–87.
53(12):3048–56. 73. Accili, D. and Arden, K.C. FoxOs at the crossroads of cellular metabolism,
49. Petersen, M.C., and Shulman, G.I. Roles of diacylglycerols and ceramides in differentiation, and transformation. Cell, 2004;117(4):421–6.
hepatic insulin resistance. Trends Pharmacol Sci, 2017;38(7):649–65. 74. Liu, Y., Dentin, R., Chen, D. et  al. A fasting inducible switch modu-
50. Kumashiro, N., Erion, D.M., Zhang, D. et al. Cellular mechanism of insulin lates  gluconeogenesis via activator/coactivator exchange. Nature,
resistance in nonalcoholic fatty liver disease. Proc Natl Acad Sci USA, 2008;456(7219):269–73.
2011;108(39):16381–5. 75. Koo, S.‐H., Flechner, L., Qi, L. et al. The CREB coactivator TORC2 is a key
51. Magkos, F., Su, X., Bradley, D. et al. Intrahepatic diacylglycerol content is regulator of fasting glucose metabolism. Nature, 2005;437(7062):1109–11.
associated with hepatic insulin resistance in obese subjects. Gastroenterology, 76. Dentin, R., Liu, Y., Koo, S.‐H. et al. Insulin modulates gluconeogenesis by
2012;142(7):1444–6. inhibition of the coactivator TORC2. Nature, 2007;449 7160;, 366–9.
37:  Non‐alcoholic Fatty Liver Disease and Insulin Resistance 471

77. Samuel, V.T., Beddow, S.A., Iwasaki, T. et al. Fasting hyperglycemia is not insulin clearance, and gene expression in adipose tissue in patients with
associated with increased expression of PEPCK or G6Pc in patients with type 2 diabetes. Diabetes 2004; 53(8):2169–76.
Type 2 Diabetes. Proc Natl Acad Sci, 2009;106(29):12121–6. 94. Basu, R., Basu, A., Chandramouli, V. et  al. Effects of pioglitazone and
78. Perry, R.J., Camporez, J.‐P.G., Kursawe, R. et al. Hepatic acetyl CoA links metformin on NEFA‐induced insulin resistance in type 2 diabetes.
­
adipose tissue inflammation to hepatic insulin resistance and type 2 diabetes. Diabetologia, 2008;51(11):2031–40.
Cell, 2015;160(4):745–58. 95. Shulman, G.I. Cellular mechanisms of insulin resistance. J Clin Invest,
79. Perry, R.J., Zhang, D., Zhang, X.‐M. et al. Controlled‐release mitochondrial 2000;106(2):171–6.
protonophore reverses diabetes and steatohepatitis in rats. Science, 96. Mayerson, A.B., Hundal, R.S., Dufour, S. et al. The effects of rosiglitazone
2015;347(6227):1253–6. on insulin sensitivity, lipolysis, and hepatic and skeletal muscle triglyceride
80. Zingone, A., Hiraiwa, H., Pan, C.‐J. et  al. Correction of glycogen storage content in patients with type 2 diabetes. Diabetes, 2002;51(3):797–802.
disease type 1a in a mouse model by gene therapy. J Biol Chem, 97. Samuel, V.T. and Shulman, G.I. Nonalcoholic fatty liver disease as a nexus
2000;275(2):828–32. of metabolic and hepatic diseases. Cell Metab, 2018;27(1):22–41.
81. Previs, S.F., Cline, G.W., and Shulman, G.I. A critical evaluation of mass 98. Petersen, K.F., Dufour, S., Befroy, D. et  al. Reversal of nonalcoholic
isotopomer distribution analysis of gluconeogenesis in vivo. Am J Physiol, hepatic steatosis, hepatic insulin resistance, and hyperglycemia by
1999;277(1 Pt 1):E154–60. ­moderate weight reduction in patients with type 2 diabetes. Diabetes,
82. Titchenell, P.M., Chu, Q., Monks, B.R. et  al. Hepatic insulin signalling 2005;54(3):603–8.
is  dispensable for suppression of glucose output by insulin in vivo. 99. Klein, S., Mittendorfer, B., Eagon, J.C. et  al. Gastric bypass surgery
Nat Commun, 2015;12(6):7078. improves metabolic and hepatic abnormalities associated with nonalco-
83. Buettner, C., Patel, R., Muse, E.D. et al. Severe impairment in liver insulin holic fatty liver disease. Gastroenterology, 2006;130(6):1564–72.
signaling fails to alter hepatic insulin action in conscious mice. J Clin Invest, 100. Sato, F., Tamura, Y., Watada, H. et  al. Effects of diet‐induced moderate
2005;115(5):1306–13. weight reduction on intrahepatic and intramyocellular triglycerides and
84. Reaven, G.M., Hollenbeck, C., Jeng, C.Y. et al. Measurement of plasma glu- glucose metabolism in obese subjects. J Clin Endocrinol Metab,
cose, free fatty acid, lactate, and insulin for 24 h in patients with NIDDM. 2007;92(8):3326–9.
Diabetes, 1988;37(8):1020–4. 101. Viljanen, A.P.M., Iozzo, P., Borra, R. et al. Effect of weight loss on liver
85. Puhakainen, I., Koivisto, V.A., and Yki‐Järvinen, H. Lipolysis and gluconeo- free fatty acid uptake and hepatic insulin resistance. J Clin Endocrinol
genesis from glycerol are increased in patients with noninsulin‐dependent Metab, 2009;94(1):50–5.
diabetes mellitus. J Clin Endocrinol Metab, 1992;75(3):789–94. 102. Lean, M.E.J., Leslie, W.S., Barnes, A.C. et  al. Primary care‐led weight
86. Paolisso, G., Tataranni, P.A., Foley, J.E. et al. A high concentration of fasting management for remission of type 2 diabetes (DiRECT): an open‐label,
plasma non‐esterified fatty acids is a risk factor for the development of cluster‐randomized trial. Lancet, 2018;391(10120):541–51.
NIDDM. Diabetologia, 1995;38(10):1213–17. 103. Taylor, R., Al‐Mrabeh, A., Zhyzhneuskaya, S. et al. Remission of human
87. Hundal, R.S., Krssak, M., Dufour, S. et al. Mechanism by which metformin type 2 diabetes requires decrease in liver and pancreas fat content but
reduces glucose production in type 2 diabetes. Diabetes, 2000;49(12):2063–9. is  dependent upon capacity for beta cell recovery. Cell Metab,
88. Madiraju, A.K., Erion, D.M., Rahimi, Y. et al. Metformin suppresses gluco- 2018;28:547–56.
neogenesis by inhibiting mitochondrial glycerophosphate dehydrogenase. 104. Oh, S., Shida, T., Yamagishi, K. et  al. Moderate to vigorous physical
Nature, 2014;510(7506): 542–6. ­activity volume is an important factor for managing nonalcoholic fatty liver
89. El‐Mir, M.Y., Nogueira, V., Fontaine, E. et al. Dimethylbiguanide inhibits disease: A retrospective study. Hepatology, 2015;61(4):1205–15.
cell respiration via an indirect effect targeted on the respiratory chain 105. Zhang, H.‐J., He, J., Pan, L.‐L. et  al. Effects of moderate and vigorous
­complex I. J Biol Chem, 2000;275(1):223–8. exercise on nonalcoholic fatty liver disease: a randomized clinical trial.
90. Foretz, M., Hébrard, S., Leclerc, J. et  al. Metformin inhibits hepatic JAMA Intern Med, 2016;176(8):1074–82.
­gluconeogenesis in mice independently of the LKB1/AMPK pathway via a 106. Bacchi, E., Negri, C., Targher, G. et al. Both resistance training and aerobic
decrease in hepatic energy state. J Clin Invest, 2010;120(7):2355–69. training reduce hepatic fat content in type 2 diabetic subjects with nonalco-
91. Miller, R.A., Chu, Q., Xie, J. et al. Biguanides suppress hepatic glucagon holic fatty liver disease (the RAED2 randomized trial). Hepatol Baltim Md,
signalling by decreasing production of cyclic AMP. Nature, 2013;58(4):1287–95.
2013;494(7436):256–60. 107. Perseghin, G., Lattuada, G., De Cobelli, F. et al. Habitual physical activity
92. Konopka, A.R., Esponda, R.R., Robinson, M.M. et al. Hyperglucagonemia is associated with intrahepatic fat content in humans. Diabetes Care,
mitigates the effect of metformin on glucose production in prediabetes. 2007;30(3):683–8.
Cell Rep, 2016;15(7):1394–1400. 108. Perseghin, G., Price, T.B., Petersen, K.F. et al. Increased glucose transport‐
93. Tiikkainen, M., Häkkinen, A.‐M., Korsheninnikova, E. et  al. Effects of phosphorylation and muscle glycogen synthesis after exercise training in
rosiglitazone and metformin on liver fat content, hepatic insulin resistance, insulin‐resistant subjects. N Engl J Med, 1996;335(18):1357–62.
AMPK: Central Regulator
38 of Glucose and Lipid
Metabolism and Target
of Type 2 Diabetes
Therapeutics
Daniel Garcia, Maria M. Mihaylova, and Reuben J. Shaw
Molecular and Cell Biology Laboratory, The Salk Institute for Biological Studies, La Jolla, CA, USA

AMPK STRUCTURE AND MECHANISM acute effects on metabolic enzymes as well as prolonged adap-
OF ACTIVATION tions in glucose and lipid metabolism through modulation of
transcriptional programs of metabolic enzymes. In addition to
A fundamental requirement of all cells is that they couple the ubiquitous roles as an energy checkpoint, AMPK also plays
availability of nutrients to signals emanating from growth fac- additional key roles in glucose and lipid metabolism in special-
tors to drive proliferation only when nutrients are in sufficient ized metabolic tissues in mammals and higher eukaryotes such
abundance to guarantee successful cell division. Even in non‐ as liver, muscle, and adipose [2]. Thus AMPK not only governs
dividing cells, nutrients in the environment supply the necessary cellular energetics, but indeed overall organismal bioenergetics
building blocks for cellular metabolism and survival, and fuel by coordinating the response between tissues to nutritional
the bioenergetic needs of the cell by providing substrates to pro- input.
duce intracellular ATP to be used for all cellular processes. AMPK was first discovered as a mammalian protein kinase
When nutrient levels fall, ATP levels fall and unless ATP‐con- that is activated by changes in intracellular adenosine nucleotide
suming biosynthetic processes are curtailed, a critical shortage levels [3]. However, it was not until later, that the yeast ortholog
of ATP will cause catastrophic cellular demise. Eukaryotic cells SNF‐1 (sucrose non‐fermenting complex) was annotated from a
all share a highly conserved metabolic checkpoint that acts as a Saccharomyces cerevisiae mutant screen for cells that failed to
sensor of ATP levels in the cell, the AMP‐activated protein grow on non‐fermentable carbon sources or sucrose [4, 5].
kinase (AMPK). As intracellular ATP levels fall due to patho- The AMP‐activated protein kinase (AMPK) is an obligate
logical stresses such as glucose or oxygen shortages, osmotic heterotrimeric kinase complex composed of a catalytic (α) sub-
stress, or disruptions of glycolysis or mitochondrial oxidative unit and two regulatory (β and γ) subunits. AMPK is activated
phosphorylation will all result in decreased ATP. Upon activa- under conditions of energy stress, when intracellular ATP levels
tion under low ATP conditions, AMPK acts a metabolic check- decline and intracellular AMP increases, as occurs during nutri-
point in the cell, suppressing ATP‐consuming biosynthetic ent deprivation or hypoxia [2]. Upon energy stress, AMP
processes while stimulating ATP‐generating processes to repair directly binds to tandem repeats of cystathionine‐β‐synthase
the initiating loss of ATP [1]. Upon activation, AMPK initiates (CBS) domains in the AMPK γ subunit. Binding of AMP is

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
38:  CENTRAL REGULATOR OF GLUCOSE AND LIPID METABOLISM AND TARGET OF TYPE 2 DIABETES THERAPEUTICS 473

thought to prevent dephosphorylation of the critical activation were first revealed (8, 9), but in the past decade our molecular
loop threonine in the α subunit [6]. The phosphorylation of the understanding of the regulation and function of AMPK has sig-
activation loop threonine is absolutely required for AMPK acti- nificantly been advanced by the elucidation of the crystal struc-
vation. In mammals, there are seven mammalian genes encod- tures of a variety of AMPK holoenzymes [10–16]. Although
ing each of the α, β, γ subunits, allowing for 12 distinct some details differ, together these studies provide a detailed
heterotrimeric variations (see Figure 38.1). There are two genes view of the architecture of the AMPK complex. The structure of
encoding catalytic subunits, α1 and α2, two regulatory β and the AMPK trimeric complex consists of three major segments or
three γ subunits that participate in the heterotrimer. Of these, γ3 “modules”: the catalytic module, the carbohydrate‐binding
appears mostly skeletal muscle specific and α2 appears to be module (CBM), and the nucleotide‐binding module (also called
most highly expressed in key metabolic tissues including mus- “regulatory fragment”). The activation loop of the α‐subunit
cle and liver. The catalytic α1 and α2 subunits contain a kinase resides at the interface between the catalytic and nucleotide‐
domain within their N‐terminus, as well as a critical region for binding modules, in close proximity to the C‐terminus of the
binding β and γ subunits within their C‐terminus. All kinases β‐subunit and the CBS repeats of the γ‐subunit. This structural
possess an activation loop which is often a target for upstream arrangement ensures that phosphorylation and dephosphoryla-
kinases, creating a conformation change that allows substrate tion of Thr172 is sensitive to conformational rearrangements
assess to the catalytic pocket, and the activation loop of the induced by nucleotide binding. The catalytic domain exhibits a
AMPK α subunits contains a single threonine (Thr172 in mam- typical eukaryotic serine/threonine kinase domain structure
malian AMPKα) that is the key regulatory site whose phospho- with a small N‐lobe and a large C‐lobe. The CBM directly con-
rylation is absolutely required across all species for AMPK tacts the N‐lobe of the kinase domain and the interface between
activation. It has been shown that α1 catalytic isoform is found these two modules forms a discrete pocket that was identified as
mainly in the cytoplasm, whereas the α2 isoform appears to be the binding site for many direct AMPK‐activating compounds.
nuclear in some cell types. In liver, phosphorylation of both the It is speculated that natural metabolites might bind this site to
α1 and α2 subunits accounts for half of the total AMPK activity regulate AMPK, however, no such metabolite has been yet iden-
and there appears to be no preferential binding of the α1 and α2 tified. The nucleotide‐binding module is made up mostly by the
subunits with the different β or γ subunits [7]. The crystal struc- γ‐subunit, which forms a flattened disk with the CBS repeats
tures of S. cerevisiae Snf1 and the human α2 kinase domains symmetrically arranged around the disk, one in each quadrant.

Length Major Site Chromosome


T172 of Expression Location
βγ Binding
α1 Kinase domain α-CTD 550 ubiquitous 5p11-14

T172

liver, muscle
α2 Kinase domain α-CTD 552 1p31
widespread

glycogen
binding αγ Binding
β1 GBD 270 ubiquitous 12q24.1-.3

β2 GBD 272 ubiquitous? 1q21.1

Bateman domain 1 Bateman domain 2


AMP AMP AMP
γ1 CBS1 CBS2 CBS3 CBS4 331 ubiquitous 12q12-14

γ2 long γ2-NTD CBS1 CBS2 CBS3 CBS4 569 ubiquitous 7q36

γ2 short CBS1 CBS2 CBS3 CBS4 328 ubiquitous 7q36

γ3 long γ3-NTD CBS1 CBS2 CBS3 CBS4 489 skeletal muscle 2q35

γ3 short γ3-NTD CBS1 CBS2 CBS3 CBS4 464 skeletal muscle 2q35

Figure 38.1  Human AMPK subunit isoforms. Domain structure, expression pattern, and alternative splice isoforms of the two catalytic kinase (α)
isoforms, the two beta regulatory subunits which contain a glycogen‐binding domain (GBD), and the three genes encoding the gamma subunits,
which each contain 4 CBS domains which directly bind to AMP as drawn.
474 THE LIVER:  UPSTREAM REGULATORS OF AMPK: LKB1 AND CAMKK

Mechanistically, these crystallographic studies reveal the or oxygen (hypoxia). It is now evident, that as complex organ-
molecular details of how adenine nucleotides and small molecule isms developed, various circulating hormones gained function
activators activate AMPK. In the case of nucleotides, the crystal to act as whole‐organism sensors and are capable of turning on
structures show that when AMP is bound to site 3, the γ‐subunit AMPK in response to metabolic stresses such as starvation. One
forms stable interactions with a few amino acids within the α‐ well‐known adipokine, adiponectin has been shown to activate
linker’s α‐RIM1 and α‐RIM2, which interact with the unoccupied AMPK in liver, leading to fatty‐acid oxidation and decrease in
site 2 and the AMP molecule bound at site 3, respectively [12, 15, blood glucose levels, consistent with previous findings that adi-
16]. The binding of the α‐RIM motifs to the γ‐subunit restricts the ponectin is able to suppress hepatic glucose production [18]. In
flexibility of the α‐linker, resulting in tighter association of the addition to hormonal input from adiponectin, AMPK activity
catalytic and nucleotide‐binding modules, which physically pro- has been shown to be modulated by leptin, resistin, ghrelin, epi-
tects Thr172 from dephosphorylation. Interestingly, the same nephrine, and cannabinoids [19]. Exercise is another metabolic
effect is proposed to occur when ADP binds site 3, raising the stress that has been shown to activate AMPK in response to
possibility that in some contexts ADP might be the relevant muscle contraction. Such activation may be due to increased
AMPK activating signal [14]. Moreover, the binding of the α‐ AMP to ATP ratios caused by movement and muscle contrac-
RIM motifs to the γ‐subunit shifts the autoinhibitory domain tion, correlating to studies in mice where electrical muscle stim-
(AID) in the α‐subunit away from the kinase domain when AMP ulation increased AMP levels and turned on AMPK [20].
is bound, thus releasing the AID’s negative effects on the kinase Two different classes of drugs for treatment of diabetes mel-
domain [10, 12, 13]. This rearrangement of the AID domain may litus type 2, biguanides such as metformin [21] and thiazolidin-
represent the molecular basis for the allosteric activation effect of ediones (TZDs) such as rosiglitazone [22] and pioglitazone [23]
AMP. According to this model, the AID can shift between kinase have been shown to activate hepatic AMPK, most likely through
domain‐bound (inactive AMPK) and nucleotide‐module‐bound perturbation of mitochondrial ATP output via mild inhibition of
states (active AMPK) depending on the nucleotide binding status. complex I of the mitochondrial respiratory chain. Metformin is
In summary, the published crystal structures concur that binding a biguanide that has been structurally modified from galegine, a
of AMP, especially at site 3, induces a conformational change that naturally occurring compound found in the French lilac (Galega
is transmitted to the kinase domain by changes in the interaction officinalis). Although it was not until 1918 that biguanides were
of the α‐RIM motifs and the AID with the nucleotide‐binding discovered to have a blood glucose lowering effect, people have
module. These structural changes, which are opposed by ATP, been using the French lilac to ameliorate a variety of maladies
result in the allosteric activation of AMPK and a compaction of since the Middle Ages [24].
the interface between the catalytic module and the nucleotide‐ Bitter melon or Momordica charantia, like the French lilac,
binding module, which protects Thr172 from dephosphorylation. has been used for hundreds of years in traditional Chinese medi-
However, it is not clear whether these structural rearrangements cine to treat many different ailments and it was not until recently
also promote Thr172 phosphorylation. that scientists isolated triterpenoid compounds from M. charan-
On the other hand, activating compounds, such as A769662, tia that are able to activate AMPK and facilitate fatty acid oxida-
activate AMPK by a different mechanism. Binding of these com- tion and glucose utilization when administrated in mice [25]. In
pounds, together with phosphorylation of serine 108 in the β‐sub- a recent study, AMPK has been also been shown to become
unit, stabilizes the CBM and strengthens the interaction of the active in response to resveratrol, a polyphenol that is found in
CBM with the kinase domain [10, 13, 15]. Specifically, binding the skin of red grapes, certain nuts and berries, and has been
of activating compounds induces the formation of an α‐helix in linked to longevity in model organisms such as yeast, nema-
the β‐subunit, termed the C‐interacting helix, which interacts with todes, drosophila, fish, and mice. Further, it was shown that res-
the so‐called C‐helix of the kinase domain (a conserved helix veratrol could mimic the benefits of dietary restriction in mice
across multiple kinases which is important for ATP binding). This by perturbing fatty liver phenotype and increasing insulin sensi-
conformational change results in a shift toward a closed, active tivity in animals fed high calorie diet [26]. Although direct
conformation of the kinase domain, protection from Thr172 mutations in AMPK have not been found in diabetic patents,
dephosphorylation, and increased substrate affinity. Interestingly, studies show that AMPK is implicated in various pathways
glycogen inhibits the CBM‐KD interaction and this may be the deregulated in such metabolic disorders and makes it an attrac-
mechanism by which glycogen inhibits AMPK [17]. tive therapeutic target in treatment of such diseases.

STRESS, HORMONES, UPSTREAM REGULATORS OF AMPK:


AND THERAPEUTICS ACTIVATE AMPK LKB1 AND CAMKK
Animals, in their multi‐cell complexity and multi‐organ utiliza- There was evidence for an upstream kinase activating AMPK as
tion, have evolved intricate mechanisms to sense and initiate early as 1978 [27] and in the years to follow scientists intently
immediate responses to energy requirement or nutrient depriva- labored to identify the kinase responsible for phosphorylation
tion. Multiple studies have now shown that cellular stress caused and activation of catalytic subunits α1 or α2. However, it was not
by starvation or exercise can activate AMPK. In single‐cell until 1996 that an “AMP‐ activated protein kinase” (AMPKK)
eukaryotes as well as mammalian cell culture conditions, was partially purified from rat liver and shown to phosphorylate
AMPK is activated in response to nutrient depletion of glucose AMPK on Thr172 [28]. Quickly after the budding yeast genome
38:  CENTRAL REGULATOR OF GLUCOSE AND LIPID METABOLISM AND TARGET OF TYPE 2 DIABETES THERAPEUTICS 475

was completed, three upstream kinases Sak1 (Pak1), Elm1, and DOWNSTREAM TARGETS I:
Tos3 were identified by whole genome screening methods and REGULATION OF ACUTE METABOLIC
were shown to act upstream of the yeast AMPK orthologue, the
SNF1 complex. When these kinases were genetically knocked
RESPONSE – ENZYME IN LIPOGENESIS
out in yeast, the results yielded the same phenotype as a snf1
In the liver, AMPK phosphorylates and regulates multiple
mutant, again placing them upstream of the SNF1 complex [29,
downstream targets involved in lipogenesis and lipid homeosta-
30]. In the human genome the closest related kinases to the yeast
sis (Figure 38.2). One of the first identified downstream targets
ones were found to be Calmodulin‐dependent protein kinases,
of AMPK was acetyl‐coenzyme A carboxylase (ACC) [41].
CaMKKα and CaMKKβ and the protein kinase LKB1. Several
ACC is an enzyme involved in the generation of fatty acid pre-
groups simultaneously showed that LKB1 [1, 31, 32] and
cursor malonyl‐CoA, a key metabolite in the regulation of
CAMKKβ [33–35] were indeed the upstream kinases acting on
energy homeostasis. Two genes encoding two different ACC
AMPK in mammals and were capable of phosphorylating both
isoforms are found in mammals – ACC1 and ACC2 – and they
AMPKα1 and AMPKα2 subunits on Thr172.
appear to have distinct tissue specificity. It has been shown that
Interestingly, LKB1 was first identified in humans as a serine/
ACC1 and ACC2 control the synthesis of two different pools of
threonine tumor suppressor kinase that is defective in the cancer
malonyl‐CoA production. ACC1 is thought to suppress the pro-
predisposing Peutz–Jeghers Syndrome [36]. In mammalian
duction of malonyl‐CoA used in fatty acid synthesis whereas
cells, LKB1 exists in a complex with two other proteins, STRAD
ACC2 stimulates fatty acid oxidation (reviewed in [42]).
(sterile‐20‐related adaptor) and MO25 (mouse protein‐25) [31]
AMPK inhibits both ACC1 and ACC2 through direct phos-
and when bound to these accessory proteins, it is stabilized and
phorylation of their homologous residues Ser79 in ACC1 and
constitutively activated. Recent data suggest that when AMP
Ser218 in ACC2. Downregulation of ACC1 activity leads to
nucleotides bind allosterically to the Bateman domains in the γ
reduced malonyl‐CoA levels and a decrease in lipogenesis.
subunit of AMPK complex, conformational changes occur that
AMPK phosphorylation of ACC2 inhibits its enzymatic activ-
protect the LKB1‐mediated phosphorylation Thr172 in the α
ity and decreases cellular levels of malonyl‐CoA levels, which
subunit [6], although effects on localization and complex forma-
leads to direct inhibition of mitochondrial fatty acid uptake
tion between LKB1 and AMPK following energy stress remain
and increased fatty acid oxidation and ATP production through
to be fully explored. In 2005, scientists showed that LKB1 is
carnitine palmitoyltransferase 1 (CPT‐1) (reviewed in [19]).
indeed the major upstream AMPK kinase in liver [37] and that
Interestingly, it has been also shown that calorie restriction can
genetic deletion of hepatic LKB1 almost completely reduces
increase AMPK activity causing a decrease of fatty acid syn-
hepatic AMPK activity. Lack of LKB1 in mouse liver rapidly
thesis or upregulation of fatty acid oxidation through inhibi-
leads to hyperglycemia and increased levels of gluconeogenic
tion of ACC. In a recent study, various AMPK activating
and lipogenic gene expression. It was also shown in these ani-
polyphenol compounds led to lowering of lipid accumulation
mals that activation of AMPK by the antidiabetic therapeutic
in HepG2 cells grown in high glucose and inhibited athero-
metformin is dependent on LKB1, and in the absence of hepatic
sclerosis in diabetic LDLR−/− mice treated with the various
LKB1, metformin is unable to lower blood glucose levels [37].
polyphenols [43].
However, it is important to note that AMPK is not the only sub-
In addition to reporting ACC as a downstream target of AMPK,
strate of LKB1. LKB1 similarly phosphorylates the activation
HMG‐CoA reductase was also found as one of the first key down-
loop of a family of 11 kinases all related to AMPK, also resulting
stream substrates [41, 44]. It had been already known for over ten
in their activation [38]. Importantly, of these 14 LKB1‐depend-
years at the point that HMG‐CoA reductase kinase (3‐hydroxy‐3‐
ent kinases, only AMPKα1 and AMPKα2 are activated under
methyl‐glutaryl‐CoA reductase or HMGR) activity was regulated
low ATP conditions, probably due to the fact that only they inter-
by an upstream kinase [27], however, that kinase had not yet been
act with AMPKγ which contains the AMP‐binding domains
discovered. Today, it is known that HMG‐CoA reductase is a rate‐
[39]. Little is currently known about what stimuli direct LKB1
limiting enzyme involved in the production of cholesterol and
toward any of these AMPK‐related kinases and current evidence
other isoprenoids, and more specifically functions in converting
suggests that LKB1 is constitutively active and these other
HMG‐CoA to mevalonic acid. By phosphorylating HMGR,
kinases may be regulated through phosphorylation at other sites
AMPK blocks anabolic or ATP consuming processes such as cho-
outside of their activation loops. Collectively, these findings map
lesterol synthesis in order to preserve intracellular ATP levels.
AMPK on the axis of a major tumor suppressor pathway and
Strikingly, the AMPK phosphorylation site in HMGR is con-
provide an interesting link between cancer and metabolism.
served throughout eukaryotes including plants.
In addition to LKB1, the calmodulin‐dependent protein kinases,
CaMKKα and CaMKKβ also regulate AMPK activity, though
only in response to calcium flux and not in response to changes in Downstream targets II: regulation of metabolic
AMP, which appears to work completely through LKB1 based on
adaptation: control of transcription
genetic knockout and RNAi studies. Genetic deletion of LKB1
dramatically reduces AMPK activation in liver suggesting little In response to changes in AMP : ATP ratios, AMPK can rapidly
role for CAMKK in this tissue, though in hypothalamic neurons regulate downstream targets via phosphorylation. However, in
controlling food intake, CAMKKβ (CAMKK2) appears to be the addition to these fast post‐translational modifications, AMPK
dominant upstream kinase for AMPK [40]. This is consistent with can also promote long‐term transcriptional changes and repro-
the fact that CaMKKα and CaMKKβ are most highly expressed in gram transcription of certain genes in response to the cellular
neurons, whereas LKB1 is more ubiquitously expressed. state. These effects are thought to be mediated by the direct
476 THE LIVER:  DOWNSTREAM TARGETS I: REGULATION OF ACUTE METABOLIC RESPONSE – ENZYME IN LIPOGENESIS

Exercise
Low Nutrients (glucose, O2) LKB1 adiponectin
Mitochondrial inhibitors: STRAD MO25 Ghrelin, cannabinoids
TZDs, biguanides
Resistin
AICAR (amp mimetic)
Resveratrol, polyphenols AMP CAMKKβ

Glycolysis inhibitors: 2-DG Ca2 +


P
Abbott A769662 γ
AMPKα
β

Glycolysis/ Glucose Uptake


iPFK2 Ser461 Transcriptional Control
TBC1D1 Ser237 of Glucose Metabolism

CRTC2 Ser171
Insulin Sensitivity FOXO3 Ser413
Fatty Acid Lipid Synthesis Protein Synthesis p300 Ser89
Oxidation AREBP Ser470
HMG CoR Ser872 TSC2 Ser1387
ACC2 Ser221 Raptor Ser792 HNF4a Ser313
ACC1 Ser79
IRS1 Ser794 Chrebp Ser568

Figure 38.2  The AMPK signaling pathway. AMPK is phosphorylated on its kinase activation loop threonine and activated by two distinct
upstream kinases in response to different stimuli. LKB1 activates AMPK in response to all stimuli that lower intracellular ATP and increase AMP.
CAMKKb activates AMPK in response to calcium flux in an AMP‐independent manner. AMPK is activated physiologically by exercise, low nutri-
ents such as lowered glucose or lowered oxygen, and hormones including ghrelin, leptin, adiponectin, and cannabinoids. Leptin is reported to
activate AMPK in peripheral tissues but inhibit AMPK in the central nervous system through poorly‐understood mechanisms. In addition, AMPK
is activated by agents that disrupt ATP production by inhibiting or poisoning the mitochondria, including uncouplers, or agents that inhibit glycoly-
sis such as the glucose analog 2‐deoxyglucose that as a competitive inhibitor for hexokinase. Additional pharmacological agents that activate
AMPK include resveratrol and related polyphenols as well as the cell‐permeable AMP‐mimetic AICAR and the first small molecule direct activator
Abbott A‐769662. Upon activation, AMPK serves to inhibit anabolic, ATP‐consuming biosynthetic processes such as protein, lipid, and glucose
synthesis, while upregulating catabolic processes to generate ATP production, including increased glycolysis, glucose uptake, and fatty acid oxida-
tion. AMPK modulates cell growth and insulin sensitivity through the mTOR signaling pathway. All current in best established in vivo direct AMPK
substrates and their AMPK phosphorylation sites are listed. All the sites listed conform to the identified optimal AMPK substrate motif.

phosphorylation of sequence‐specific transcription factors and LKB1‐dependent. In mice lacking hepatic LKB1, CRTC2 is
transcriptional coactivators by AMPK. Of the two best‐studied hypophosphorylated and predominantly nuclear compared to
transcriptional programs controlled by AMPK in liver, gluco- wild‐type mice [37]. Furthermore, LKB1−/− mice in liver had
neogenesis and lipogenesis, a number of direct substrates for dramatic increases in fasting blood glucose levels, which were
AMPK in gluconeogenesis have been identified. greatly attenuated upon introduction of an shRNA reducing
Interestingly, two different coactivators which modulate CRTC2 levels in the liver, reinforcing the idea that gluconeo-
cyclic AMP‐responsive element binding protein (CREB)‐ genesis in liver is controlled by LKB1‐dependant kinases and
dependent transcription in response to glucagon induced by that CRTC2 is a key downstream target of these kinases.
fasting are the histone acetyltransferase p300 and the CREB Consistent with these findings in the LKB1 liver‐specific
coactivator CRTC2 (also known as TORC2 or transducer of knockout mice, animals bearing a liver‐specific knockout of the
regulated CREB activity 2) [45–47]. CRTC2 is a critical rate‐ AMPKα2 isoform also exhibited elevated hepatic glucose out-
limiting transcriptional regulator of gluconeogenesis in mice put, glucose intolerance, impaired leptin‐ and adiponectin‐reg-
via effects on hepatic CREB targets including the PGC‐1α pro- ulated hepatic glucose production [50] (see Table 38.1). Future
moter. When AMPK is active, it can phosphorylate CRTC2 on studies will be needed to dissect the temporal and spatial regu-
residue Ser171, which in turn allows CRCT2 to associate to lation of the contexts in which AMPKa2 or SIK1 control these
14–3–3 proteins and be sequestered from the nucleus to the key modulators of CREB and gluconeogenesis. Interestingly,
cytoplasm, an event that shuts off transcription of gluconeo- SIK1 itself is a CREB target, providing a time‐delayed mecha-
genic enzymes [48, 49]. If AMPK activity is decreased, nism to attenuate chronic CREB‐dependent transcription.
hypophosphorylated CRTC2 can conversely move into the Phosphorylation of p300 and CRTC2 by AMPK and SIK
nucleus where it binds to transcriptional coactivator CREB and kinases may be key effectors of metformin in the control of
promotes transcription of PGC1α and downstream gluconeo- type 2 diabetes. How AMPK phosphorylation of p300 may
genic targets PEPCK and G6Pase (see Figure 38.3). Strikingly, regulate its many other downstream interacting transcription
both p300 and CRTC2 have been shown to be phosphorylated factors key to hepatic metabolism remains to be examined,
at the same regulatory sites by either AMPK or its related fam- although mice lacking LKB1 or AMPK in liver provide excel-
ily member SIK1 (salt‐inducible kinase 1), both of which are lent tools for such future studies.
38:  CENTRAL REGULATOR OF GLUCOSE AND LIPID METABOLISM AND TARGET OF TYPE 2 DIABETES THERAPEUTICS 477

14-3-3

LKB1 P

CRTC2
cytoplasm
P
P
SIK
14-3-3 AMPK nucleus
P
P
AMPK
p300

P PGC1α
P CRTC2 P P
ChREBP PEPCK
FASN CREB FoxO3 HNF4a AREBP
PGC-1α G6Pase

Inhibition of Inhibition of
Lipogenesis Gluconeogenesis

Figure 38.3  LKB1 and AMPK‐mediated control of metabolic transcriptional programs. AMPK and its related family members SIK1 and SIK2
(not shown) phosphorylate a common set of substrates including the CREB coactivator CRTC2 (CREB‐regulated transcriptional coactivator 2),
previously known as TORC2 (transducer of regulated CREB 2) and the histone acetyltransferase p300. Phosphorylation of CRTC2 creates a 14‐3‐3
docking site which then results in 14‐3‐3 mediated nuclear export of CRTC2. This cytoplasmic sequestration of CRTC2 by AMPK or SIK kinases
causes an inhibition of CREB‐dependent transcriptional targets including key mediators of gluconeogenesis like the PGC‐1α coactivator. In addi-
tion to suppressing PGC‐1a mRNA expression, AMPK also directly phosphorylates a handful of transcription factors (FOXO3, HNF4a, and
AREBP) that directly bind to the promoters of the two key gluconeogenic enzymes PEPCK and G6Pase. In addition to these gluconeogenesis regu-
lators, AMPK also is reported to phosphorylate Chrebp, a key lipogenic transcription factor that controls levels of fatty acid synthase (FASN),
acetyl‐CoA carboxylase 1 (ACC1), and L‐pyruvate kinase mRNA.

Table 38.1  Mouse models of AMPK/LKB1 function in liver It has been also shown that a subset of patients with MODY
Mouse model Metabolic phenotype Reference (maturity‐onset diabetes of the young) form of diabetes harbor
AMPKα1 knockout None [48]
mutations in HNF4α [53]. Interestingly, HNF4α liver specific
AMPKα2 knockout Glucose uptake in muscle [49] knockout mice do not develop hyperglycemia unlike MODY
Hyperglycemia, low insulin [48] patients, but do develop lipid accumulation, consistent with
AMPKα2 liver knockout Hyperglycemia [46] altered triglyceride levels in MODY patients [54]. These find-
Glucose intolerance
Hyperlipidemia ings further reinforce the possible role of AMPK in lipid and
LKB1 liver knockout Hyperglycemia [30] glucose homeostasis through modulation of downstream tran-
Glucose intolerance scriptional regulators such as HNF4α.
Hyperlipidemia
In addition to HNF4α, in a 2006 report AMPK was shown to
directly phosphorylate another transcription factor named the
In addition to direct phosphorylation of coactivators, AMPK AICAR response‐element binding protein (AREBP) that
also phosphorylates the hepatocyte nuclear factor 4 alpha directly binds the PEPCK promoter [55]. AMPK phosphoryla-
(HNF4α), which is a key transcription factor of the nuclear tion of AREBP on Ser470 reduces its ability to bind DNA and in
receptor superfamily which binds to the promoters of the two turn lowers expression of PEPCK. Thus, like HNF4α, AMPK
key gluconeogenic enzymes, phosphoenolpyruvate carboxykin- phosphorylation of AREBP controls its ability to bind to DNA
ase (PEPCK) and G6Pase. It is hypothesized that AMPK phos- and promote transcription of downstream target genes.
phorylation of HNF4α on residue Ser304 decreases the protein’s In addition to effects on gluconeogenesis, AMPK activation
stability by interfering with its ability to dimerize and bind to is also known to inhibit hepatic lipogenesis. While some of that
DNA and may further promote its degradation [51, 52]. In addi- is due to acute effects on lipogenic enzymes as previously dis-
tion to PEPCK and G6Pase, HNF4α controls gene expression of cussed, it is also known that AMPK activation leads to decreased
glucose transporter GLUT2 and glycolytic enzymes such as transcription of key hepatic lipogenic enzymes, including fatty
aldolase B and liver‐type pyruvate kinase (L‐PK), which are acid synthase (FASN) as well as ACC1. Two sequence–sequence
diminished when AMPK is activated by 5‐aminoimidazole‐4‐ transcription factors are known to coregulate many of these
carboxamide‐1‐β‐D‐ribofuranoside (AICAR) in hepatocytes [51]. lipogenic enzymes: the sterol‐responsive binding protein
478 THE LIVER:  DOWNSTREAM TARGETS I: REGULATION OF ACUTE METABOLIC RESPONSE – ENZYME IN LIPOGENESIS

(SREBP‐1c) and the carbohydrate response binding protein [63]. Indeed, we previously demonstrated that AMPK‐related
(ChREBP). ChREBP is highly expressed in liver and has essen- kinases downstream of LKB1 can phosphorylate HDAC4,
tial roles in glucose‐induced transcription of liver pyruvate HDAC5, and HDAC7 in mouse liver, which suppresses their
kinase (L‐PK), in addition to its effects on lipogenic enzyme ability to enhance FOXO‐dependent gluconeogenesis [64].
promoters. Phosphorylation by AMPK on Ser568 of ChREBP
[56], promotes decreased DNA binding which causes a decrease Downstream targets III: regulation of cell
in the transcription of glycolytic and lipogenic genes. It is
growth and insulin signaling via mTOR
known now that glucose metabolism can be repressed by fatty
acids, which act as another source of energy when glucose needs Environmental factors cue cells to cease growing and dividing
to be preserved. This effect has been named the fatty acid “spar- when conditions are unfavorable. These mechanisms are con-
ing effect” on glucose. served from the smallest and simplest of eukaryotes to the most
Like ChREBP, SREBP‐1 is also directly regulated by AMPK complex ones. When nutrients are scarce, cellular energy sensor
[21, 57, 61]. Due to its rate‐limiting effects on lipogenic AMPK gets activated and inhibits energy demanding processes
enzyme expression, SREBP‐1 has been linked to insulin resist- such as protein synthesis and cell growth. One of the ways
ance, dyslipidemia and type 2 diabetes [58, 59]. In one study, AMPK accomplishes that task is by negatively regulating the
treatment of rat hepatocytes with AMPK activators such as mechanistic target of rapamycin (mTOR) pathway. mTOR is a
metformin or AICAR led to suppressed SREBP‐1 mRNA serine/threonine kinase highly conserved in all eukaryotes and a
expression, and also lowered hepatic mRNA levels for central regulator of cell growth. Whereas AMPK is active under
SREPB‐1 controlled genes FAS and S14 [21]. This suggests nutrient‐poor conditions and inactive under nutrient‐rich condi-
that AMPK activation through metformin can inhibit the tions, mTOR is activated in the inverse pattern. In higher eukar-
expression of lipogenic genes. Indeed, metformin treatment or yotes, mTOR activation requires positive signals from both
overexpression of an activated allele of AMPK was found to be nutrients (glucose, amino acids) and growth factors. mTOR, like
sufficient to reduce triglyceride content in insulin resistant its budding yeast orthologs, is found in two biochemically and
HepG2 cells [60]. Mice lacking hepatic AMPK function due to functionally discrete signaling complexes [65]. In mammals,
liver‐specific LKB1 deletion show elevated SREBP1 and the mTORC1 complex is composed of four known subunits:
SREBP1 target genes resulting in lipid accumulation and raptor (regulatory associated protein of mTOR), PRAS40,
hepatic steatosis [37]. In 2011, Srebp1 was found to be directly mLST8, and mTOR. The mTORC2 complex contains rictor
phosphorylated by AMPK and its highly related kinase SIK1, (rapamycin insensitive companion of mTOR), mSIN1, PRR5/
both of which appear to suppress activation of full‐length Protor, mLST8, and mTOR [66]. Signaling from mTOR com-
Srebp1a in the ER membrane, preventing appearance of plex 1 (mTORC1) is nutrient‐sensitive, acutely inhibited by the
Srebp1c in the nucleus [61]. While the activity of the SREBP bacterial macrolide rapamycin, and controls cell growth, angio-
proteins is primarily regulated through intracellular concentra- genesis, and metabolism. In contrast, mTORC2 is not sensitive
tions of unsaturated fatty acids and cholesterol, AMPK phos- to nutrients, nor acutely inhibited by rapamycin, and its known
phorylation of Ser372 on SREBP1 may also inhibit its activity substrates include the hydrophobic motif phosphorylation sites
by preventing proteolytic processing and thus transcriptional in AGC kinases including Akt and PKC family members.
activity. Additional studies are needed however to truly sort out Downstream of the AMPK‐ and rapamycin‐sensitive raptor‐
the requirement and roles for AMPK‐dependent phosphoryla- mTOR (mTORC1) complex are its two well‐characterized sub-
tion of Srebp1 in vivo. As seen from the above‐mentioned stud- strates: 4EBP1 and the p70 ribosomal S6 kinase. Phosphorylation
ies, AMPK plays a major role in the gluconeogenic of 4EBP‐1 by mTORC1 suppresses its ability to bind and inhibit
transcriptional program through p300, CRTC2, HNF4α, and the translation initiation factor eIF4E. mTORC1 mediates phos-
AREBP, as well as the lipogenic transcriptional program phorylation of S6K at a Thr residue in a hydrophobic motif at
through ChREBP and SREBP‐1. the C‐terminus of the kinase domain. A specific motif (TOS
Finally, studies have hinted that AMPK may play a more motif) found in both 4EBP1 and S6K was shown to mediate
global role in transcription through direct phosphorylation and direct binding of these proteins to raptor allowing them to be
regulation of transcriptional repressors histone deacetylases phosphorylated in the mTORC1 complex. Mechanistic details
class IIa (HDACs IIa) enzymes [62] and transcriptional activa- of how mTORC1 regulates the assembly of translational initia-
tor histone acetyltransferase p300 [45]. It has been shown that tion complexes via a number of ordered phosphorylation events
AMPK can phosphorylate class IIa HDAC5 on residues Ser259 were recently discovered [67]. mTORC1‐dependent translation
and Ser498 in human primary myotubes, promoting 14‐3‐3 is known to control a number of specific cell growth regulators,
binding and dissociation from DNA binding transcription part- including cyclin D1, the HIF‐1α transcriptional factor and c‐
ner myocyte enhancer factor‐2 (MEF2), which in turn allows myc, which in turn promote processes including cell cycle pro-
expression of downstream target genes [62]. Although direct gression, cell growth, glycolysis, and angiogenesis, all
regulation of class IIa HDACs by AMPK has not yet been impli- contributing to enhanced tumorigenesis [66].
cated in liver metabolism, this provides an attractive postulate Upstream components of the mTORC1 complex were initially
where AMPK can simultaneously control multiple downstream discovered through classical cancer genetics. The TSC2 tumor
transcriptional events in response to upstream metabolic suppressor tuberin and its obligate binding partner hamartin
stresses. Such analysis will not happen without challenges, (TSC1), are mutated in a familial tumor syndrome called tuber-
since it has been shown that multiple upstream kinases are capa- ous sclerosis complex (TSC). TSC patients are predisposed to
ble of phosphorylating the same critical residues of HDAC 5 widespread benign tumors termed hamartomas in kidney, lung,
38:  CENTRAL REGULATOR OF GLUCOSE AND LIPID METABOLISM AND TARGET OF TYPE 2 DIABETES THERAPEUTICS 479

brain, and skin. Genetic studies in Drosophila and mammalian backup mechanism which allows AMPK to inhibit cell growth
cells identified the TSC tumor suppressors as critical upstream and proliferation through the mTORC1 pathway. Collectively,
inhibitors of the mTORC1 complex. TSC2 (also known as these observations prompted the discovery of yet another novel
tuberin) contains a GTPase‐activating protein (GAP) domain at mechanism of inhibition. The mTOR kinase exists in complex
its C‐terminus that inactivates the small Ras‐like GTPase Rheb, consisting of mLST8/Gbl, PRAS40, and the scaffold protein
which has been shown to associate with and directly activate the Raptor. In a recent study, it was shown that AMPK is able to
mTORC1 complex in vitro [68]. Loss of TSC1 or TSC2 therefore directly phosphorylate Raptor on two conserved residues Ser722
leads to hyperactivation of mTORC1. Phosphorylation of TSC1 and Ser792, which in turn induces binding to 14‐3‐3 proteins
and TSC2 serves as an integration point for a wide variety of envi- and inactivation the mTORC1 complex [75]. Taken together
ronmental signals that regulate mTORC1 [69]. One of the key with previous studies, these findings indicate that AMPK
activators of the mTORC1 pathway is PI3‐kinase, which plays a directly phosphorylates both TSC2 and raptor to inhibit
key role in promoting cell growth and insulin‐mediated effects on mTORC1 activity by a dual‐pronged mechanism (see
metabolism. PI3‐kinase activates the serine/threonine kinase Akt Figure 38.4). Importantly, metformin treatment of mice leads to
which directly phosphorylates and inactivates both TSC2 and an robust phosphorylation of raptor Ser792 in murine liver, an
inhibitor of the mTORC1 complex named PRAS40 [68, 70]. effect which is ablated in the LKB1‐liver specific knockout
In addition to these growth factor cues that activate mTORC1, mice [75]. LKB1‐liver specific knockout mice which lack
the complex is rapidly inactivated by a wide variety of cell AMPK activity in liver exhibit hyperactivation of mTORC1
stresses, thereby ensuring that cells do not continue to grow signaling in the liver including increased phosphorylation of
under unfavorable conditions. One of the unique aspects of the S6K1 and 4EBP1 under ad libitum fed conditions [37]. In addi-
mTORC1 complex is that unlike many of the aforementioned tion, hormones that activate AMPK in liver including glucagon
growth factor activated kinases, it is dependent on nutrient avail- [76] and adiponectin [77] have been reported to suppress
ability for its kinase activity. Withdrawal of glucose, amino mTORC1 signaling. Very recently, our laboratory collaborating
acids, or oxygen leads to rapid suppression of mTORC1 activity with the lab of Brendan Manning demonstrated that AMPK is
even in the presence of full growth factors [69]. Upon LKB1‐ genetically required in primary hepatocytes and in the livers of
and AMP‐dependent activation of AMPK by nutrient loss, mice for metformin to suppress mTORC1 signaling [78]. Using
AMPK directly phosphorylates the TSC2 tumor suppressor on liver‐specific knockouts of AMPKα1 and AMPKα2 we found
conserved serine sites distinct from those targeted by other that metformin was completely unable to suppress mTORC1
kinases, which constitutes one mechanism by which glucose after metformin treatment in mice. However, in cell culture, at
and oxygen control mTORC1 activation [71–74]. Interestingly, higher doses and at later timepoints, metformin induced sup-
TSC2 orthologs are absent from lower eukaryotes such as S. pression of mTORC1 signaling suggestive of a transcriptional
cerevisiae and Caenorhabditis elegans and mammalian cells response which may be due to activation of the mito ER stress
lacking TSC2 still remain partially sensitive to AMPK activa- pathway and ATF4‐dependent activation of REDD1 mRNAs
tion, indicating that there may be an alternative and more ancient and other mTORC1 inhibitor mRNAs [79].

Insulin R

IRS1/2

PI3K
mTOR rictor
LKB1 PIP3 PTEN
AMP FOXO
Akt
GSK3
metformin AMPK TSC2
TSC1
14

Rheb
-3
-3

-3
-3
14

raptor PRAS40
rapamycin mTOR

4EBP1 S6K
Srebp1
EIF4E S6

Figure 38.4  AMPK regulates the mechanistic target of rapamycin (mTOR) pathway to control cell growth and insulin sensitivity. AMPK directly
phosphorylates the TSC2 (tuberous sclerosis 2) tumor suppressor and the key mTOR scaffold subunit raptor to inhibit the activity of the mTOR‐rap-
tor kinase complex (mTORC1) toward its downstream substrates 4EBP1 and S6K1. PI3K‐activity activates mTORC1 by Akt‐dependent phospho-
rylation of TSC2 and the mTORC1 inhibitor PRAS40. Under conditions of nutrient excess, overstimulated mTORC1 and its substrate S6K both
directly phosphorylate the insulin receptor substrate 1 and 2 proteins, resulting in their degradation. Thus too much mTORC1 activity attenuates
insulin signaling, leading to cellular insulin resistance. AMPK reverses this resistance by inactivating mTORC1 and can also directly phosphorylate
IRS1 itself. mTORC1 has also been recently shown to play a role in the control of lipogenesis through regulation of SREBP1.
480 THE LIVER:  DOWNSTREAM TARGETS I: REGULATION OF ACUTE METABOLIC RESPONSE – ENZYME IN LIPOGENESIS

Beyond effects on cell growth, mTOR also has effects on degradation. It is a process used by cells both for normal turno-
lipid metabolism that may be particularly important in liver. ver and for the generation of nutrients in response to energy
One key regulator of lipogenesis is the aforementioned SREBP‐1 shortages. AMPK potently promotes autophagy through several
transcription factor. Recently, mTORC1 signaling was shown to mechanisms. AMPK phosphorylates and activates ULK1
be required for nuclear accumulation of SREBP1 and the induc- (unc‐51‐like autophagy‐activating kinase 1), which triggers the
tion of SREBP1 target genes [80]. Consistent with previous initiation of the autophagic cascade [85–87]. Importantly,
results with metformin, treatment with the AMPK activators mTOR strongly suppresses autophagy, in part by directly phos-
AICAR and 2DG, or the mTORC1 inhibitor rapamycin resulted phorylating and inhibiting ULK1 [86]. Accordingly, AMPK
in suppression of nuclear SREBP1 accumulation [80]. In future promotes autophagy not only by direct activation of ULK1 but
studies, it will be important to define how much of the lipid‐ also by negatively regulating mTORC1 and blocking its inhibi-
reducing effects of AMPK are due to direct phosphorylation of tory effect on ULK1. Thus, ULK1 is yet another node at which
lipogenic enzymes such as acetyl‐CoA carboxylase (ACC), and AMPK and mTOR regulate a specific metabolic process in
how much are due to effects on SREBP‐1 or Chrebp‐dependent opposing fashion. AMPK also stimulates autophagy initiation
transcription through effects of AMPK on mTORC1, versus by differential regulation of VPS34 (vacuolar protein sorting
direct phosphorylation of Srebp‐1 and Chrebp by AMPK. 34) containing complexes [88], which are important for the ini-
One final aspect of liver physiology that may be under control of tiation and formation of autophagosomes. AMPK was reported
the AMPK–mTOR axis is insulin signaling. A major route by to directly phosphorylate and inhibit VPS34 in non‐autophagic
which excess nutrients downregulate insulin signaling leading to complexes that do not contain autophagy adaptor proteins,
cellular insulin resistance is through hyperactivation of the while enhancing VPS34 activity in pro‐autophagic complexes
mTORC1 complex. Excess glucose leads to hyperactivation of that contain Beclin‐1 by directly phosphorylating Beclin‐1 [88].
mTOR via suppression of AMPK, and excess fats and excess In this way, AMPK presumably suppresses nonessential vesicle
amino acids also act to hyperactivate mTORC1 [81]. The mTOR/ trafficking in favor of membrane trafficking into the autophagy
Raptor complex, along with its key downstream substrate S6K, pathway during nutrient starvation. Given that both AMPK and
have been shown to directly phosphorylate the insulin receptor sub- ULK1 have been reported to directly phosphorylate distinct
strate‐1 and 2 (IRS1 and IRS2), leading to their proteasome degra- sites in both Beclin‐1 and Vps34, much remains to be clarified
dation. The same is observed under conditions of hyperinsulinemia, about the temporal and spatial control of autophagy initiation in
as insulin signaling itself also leads to increases in mTORC1 activ- response to different stresses. In addition, AMPK and ULK1
ity as described earlier. The net effect is a negative feedback loop have also both been reported to phosphorylate and control the
whereby too much mTOR/raptor activity leads to hyperphospho- localization of Atg9, a transmembrane protein involved in early
rylation of IRS1/IRS2 and suppression of PI3‐kinase and Akt sign- autophagosome formation [87, 89, 90].
aling downstream of the insulin receptor [81]. This nutrient induced AMPK has also recently been shown to promote autophagy
hyperactivation of mTORC1 and consequent downregulation of through transcriptional mechanisms, via regulation of Tfeb
Akt signaling is observed in a cultured cell systems and also in (transcription factor EB), a master transcriptional regulator of
peripheral tissues of mice on a high fat diet. Illustrating its impor- lysosomal genes and autophagy. Although no direct link
tance downstream of mTORC1 in the IRS1/IRS2 inhibition, this between AMPK and Tfeb has been reported, AMPK can acti-
effect is lost in peripheral tissues from an S6K1−/− mouse [82]. vate Tfeb through inhibition of mTORC1, thus blocking the
As one of the key direct substrates of AMPK is raptor and ability of mTOR to phosphorylate and translocate Tfeb out of
TSC2, AMPK activation leads to a inhibition of mTORC1 and the nucleus [91]. Furthermore, through phosphorylation and
its phosphorylation of IRS1 Ser636/639 and negative feedback activation of the transcription factor FOXO3a (Forkhead box
loop on PI3‐kinase/Akt signaling. Exogenous LKB1 expression O3) [92], AMPK has been reported to increase the levels of
in HEK293 cells and metformin treatment of human hepatocel- CARM1 (coactivator‐associated arginine methyltransferase 1),
lular carcinoma HepG2 cells can suppress phosphorylation of an important cofactor for Tfeb transcription [93].
IRS‐1 on Ser636/639 and induce Akt phosphorylation [83]. In In addition to general autophagy, several lines of evidence
addition to suppressing phosphorylation of IRS‐1 by mTORC1, indicate that AMPK promotes mitophagy, the process of degra-
AMPK has also been shown to directly phosphorylate IRS1 dation of defective mitochondria. Indeed, activation of ULK1
itself on S789, though the functional consequence of that phos- by AMPK was shown to be required for proper removal of dam-
phorylation event on IRS function remains uncertain [83, 84]. aged mitochondria via mitophagy, though the details of how
Taken altogether, these studies demonstrate a mechanism by ULK1 regulates mitophagy are not fully resolved [85]. A neces-
which AMPK activation can promote PI3‐kinase/Akt activity sary step preceding removal of damaged mitochondria is the
while simultaneously reducing mTORC1 activity. This provides fragmentation of mitochondria in response to mitochondrial
cells with a negative feedback switch that integrates upstream insults, in order to separate and target damaged mitochondrial
signals from both nutrients and growth factors and allows the fragments to turnover via the mitophagy pathway. This highly
cells to maintain energy homeostasis and insulin sensitivity. conserved process is known as mitochondrial fission. Recently,
a novel mechanism was elucidated by which AMPK promotes
Downstream targets IV: autophagy mitochondrial fission [94]. In this study, AMPK was demon-
and mitophagy strated to induce mitochondrial fission during energy stress
through direct phosphorylation of MFF (mitochondrial fission
Autophagy is a cellular process in which proteins, organelles, factor), which then serves as a receptor for DRP1 (dynamin‐
and other macromolecules are delivered to the lysosomes for related protein 1), the enzyme that catalyzes mitochondrial
38:  CENTRAL REGULATOR OF GLUCOSE AND LIPID METABOLISM AND TARGET OF TYPE 2 DIABETES THERAPEUTICS 481

fission [94]. Once at the mitochondria, DRP1 splits damaged respiratory chain in mitochondria [109]. This leads to a change in
mitochondria into smaller fragments that are presumably more the ATP‐to‐AMP ratio and canonical AMPK activation. There
efficiently cleared by autophagosomes. Furthermore, AMPK has been extensive research to evaluate the contribution of
activates PGC1α (peroxisome proliferator‐activated receptor AMPK to the effect of metformin on circulating glucose and
gamma, coactivator 1α), a master regulator of mitochondrial lipids. AMPK phosphorylation of ACCs has been proposed as a
biogenesis, reportedly via direct phosphorylation of PGC1α master contributor to the changes in lipid synthesis that are
[95] but also by promoting NAD+‐dependent activation of induced by metformin, which in turn modulates insulin sensitiv-
PGC1α by Sirt1 (sirtuin 1) [96]. Interestingly, Tfeb, similar to ity and glucose uptake in muscle. One key piece of evidence in
its family member Tfe3, was recently reported to drive mito- favor of this hypothesis came from the generation of a mouse
chondrial biogenesis as well [97, 98], which offers the possibil- knock‐in mutant lacking the AMPK site on both ACC1 and
ity that activation of Tfeb, or Tfe3, might be yet another ACC2 [110]. This mouse model revealed that these phosphoryla-
mechanism by which AMPK can promote the regeneration of tion events mediate the insulin‐sensitizing effect of metformin,
mitochondria. In all, AMPK coordinates mitochondrial fission thus establishing AMPK as a relevant target in the action of met-
and mitophagy in the acute response to mitochondrial insults, formin. Despite controversy [111], AMPK is widely viewed as
and after sustained energy stress, AMPK promotes transcrip- an essential component of the action of metformin at physiologi-
tional induction of mitochondrial biogenesis. In this fashion, cal concentrations [78, 112]. Given the role of the AMPK–ACC
AMPK serves as a central mediator of mitochondrial quality, pathway in regulating fatty acid synthesis, AMPK activation is
ensuring metabolic efficiency in cells and tissues. also an attractive treatment option for conditions associated with
increased fatty acid production, such as non‐alcoholic fatty liver
Downstream targets V: polarization disease (NAFLD) [113]. In addition to diabetes, retrospective
studies have revealed that patients taking metformin had a
of hepatocytes decreased occurrence of cancer [114]. However, whether direct
The predominant upstream activating kinase for AMPK, LKB1, activation of AMPK would be sufficient to replicate the benefi-
is a well‐conserved key regulator of cell polarity and trafficking, cial effect of metformin was not known until recent advances in
and metabolism, due at least in part to its ability to phosphoryl- the generation of potent and specific small‐molecule activators
ate and activate the MARK/Par1 subfamily of AMPK related of AMPK. Most notably, two recent studies have revealed that
kinases [37]. In an intestinal cell line in culture, genetic activa- direct AMPK activation indeed improves symptoms of type 2
tion of LKB1 induced apical–basolateral polarity in the absence diabetes in multiple animal models. One study revealed that a
of cell–cell or cell–matrix cues [99], though evidence of which pan‐specific AMPK activator improved diabetes symptoms in
of the 14 AMPK‐related kinase family were involved has not several animal models, including rodents and monkeys [115]. In
been determined. Nonetheless, AMPK itself has been connected another study, a direct AMPK activator increased glucose uptake
to cell polarization, particularly in hepatocytes and its activation in muscle and reduced blood glucose in type 2 diabetes models
enhanced bile canalicular formation, whereas inhibition resulted [116]. Interestingly, inactivation of AMPK in the liver had no
in loss of polarity and mislocalization of apical transporters effect on the efficacy of the drug, whereas muscle‐specific dele-
[100–102]. In MDCK cells, AMPK regulates tight junction tion of AMPK abolished the effect of the drug, establishing mus-
assembly and disassembly in response to calcium depletion cle AMPK as a key therapeutic target in type 2 diabetes [116].
[103–104]. LKB1‐AMPK activation phosphorylates the core Although hepatic AMPK may not be as central as originally
tight junction protein cingulin on Ser137 [105] while at the hypothesized for metformin effects on glucose homeostasis in
same time stabilizing existing cell junctions to maintain cell type 2 diabetes, hepatic AMPK is still a very attractive target in
polarity through phosphorylation of Gα‐interacting vesicle‐ the emerging field of therapeutics for NASH and NAFLD [113].
associated protein (GIV/Girdin) on Ser245 [106]. Interestingly, Indeed in every system from C. elegans to mammals, all studies
GIV/Girdin was previously reported as an Akt substrate [107], concur that AMPK activation suppresses de novo lipogenesis and
reinforcing the idea that AMPK and mTORC1/Akt converge on lipid accumulation. While much has been learned, there is still
a common small set of effector proteins (ULK1, MFF, Srebp1, even more yet to be learned about this ancient energy sensor and
Foxo, Girdin) central to metabolism and growth [108]. central metabolic regulator.
Additional studies are needed to delineate the roles of AMPK
and mTORC1/Akt in hepatocyte polarity and liver zonation in
vivo. REFERENCES
1. Shaw, R.J., Kosmatka, M., Bardeesy, N. et al. The tumor suppressor LKB1
Therapeutics and future perspectives kinase directly activates AMP‐activated kinase and regulates apoptosis in
response to energy stress. Proc Natl Acad Sci USA, 2004;101:3329–35.
AMPK has received a lot of attention as a potential target for
2. Kahn, B.B., Alquier, T., Carling, D., and Hardie, D.G. AMP‐activated protein
treating diseases associated with metabolic perturbation. This kinase: ancient energy gauge provides clues to modern understanding of
includes diabetes, obesity, and fatty liver diseases and also can- metabolism. Cell Metab, 2005;1:15–25.
cer, which is often associated with changes in metabolism. The 3. Carlson, C.A. and Kim, K.H. Regulation of hepatic acetyl coenzyme A car-
type 2 diabetes drug metformin had been used for decades when bozylase by phosphorylation and dephosphorylation. J Biol Chem,
1973;248:378–80.
a study showed that its mechanism of action involved the activa- 4. Mitchelhill, K.I. et al. Mammalian AMP‐activated protein kinase shares struc-
tion of AMPK in hepatocytes [21]. Mechanistically, metformin tural and functional homology with the catalytic domain of yeast Snf1 protein
induces energy stress through inhibition of complex I of the kinase. J Biol Chem, 1994;269:2361–4.
482 THE LIVER:  REFERENCES

5. Woods, A., Munday, M.R., Scott, J., Yang, X., Carlson, M., and Carling, D. Yeast 31. Hawley, S.A., Boudeau, J., Reid, J.L. et al. Complexes between the LKB1
SNF1 is functionally related to mammalian AMP‐activated protein kinase and tumor suppressor, STRADalpha/beta and MO25alpha/beta are upstream
regulates acetyl‐CoA carboxylase in vivo. J Biol Chem, 1994;269:19509–15. kinases in the AMP‐activated protein kinase cascade. J Biol, 2003;2:28.
6. Sanders, M.J., Grondin, P.O., Hegarty, B.D., Snowden, M.A., and Carling, 32. Woods, A., Johnstone, S.R., Dickerson, K. et al. LKB1 is the upstream kinase
D. Investigating the mechanism for AMP activation of the AMP‐activated in the AMP‐activated protein kinase cascade. Curr Biol, 2003;13:2004–8.
protein kinase cascade. Biochem J, 2007;403:139–48. 33. Hawley, S.A., Pan, D.A., Mustard, K.J. et al. Calmodulin‐dependent protein
7. Cheung, P.C., Salt, I.P., Davies, S.P., Hardie, D.G., and Carling, D. kinase kinase‐beta is an alternative upstream kinase for AMP‐activated pro-
Characterization of AMP‐activated protein kinase gamma‐subunit isoforms tein kinase. Cell Metab, 2005;2:9–19.
and their role in AMP binding. Biochem J, 2000;346(3):659–69. 34. Hurley, R.L., Anderson, K.A., Franzone, J.M., Kemp, B.E., Means, A.R.,
8. Amodeo, G.A., Rudolph, M.J., and Tong, L. Crystal structure of the hetero- and Witters, L.A. The Ca2+/calmodulin‐dependent protein kinase kinases
trimer core of Saccharomyces cerevisiae AMPK homologue SNF1. Nature, are AMP‐activated protein kinase kinases. J Biol Chem, 2005;280:29060–6.
2007;449:492–5. 35. Woods, A., Dickerson, K., Heath, R. et al. C(Ca2+)/calmodulin‐dependent
9. Xiao, B., Heath, R., Saiu, P. et al. Structural basis for AMP binding to mam- protein kinase kinase‐beta acts upstream of AMP‐activated protein kinase in
malian AMP‐activated protein kinase. Nature, 2007;449:496–500. mammalian cells. Cell Metab, 2005;2:21–33.
10. Calabrese, M.F., Rajamohan, F., Harris, M.S. et al. Structural basis for AMPK 36. Hemminki, A., Markie, D., Tomlinson, I. et  al. A serine/threonine kinase
activation: natural and synthetic ligands regulate kinase activity from opposite gene defective in Peutz‐Jeghers syndrome. Nature, 1998;391:184–7.
poles by different molecular mechanisms. Fold Des, 2014;22:1161–72. 37. Shaw, R.J., Lamia, K.A., Vasquez, D. et al. The kinase LKB1 mediates glu-
11. Chen, L., Wang, J., Zhang, Y.‐Y. et al. AMP‐activated protein kinase under- cose homeostasis in liver and therapeutic effects of metformin. Science,
goes nucleotide‐dependent conformational changes. Nat Struct Mol Biol, 2005;310:1642–6.
2012;19:716–8. 38. Lizcano, J.M., Goransson, O., Toth, R. et al. LKB1 is a master kinase that
12. Chen, L., Xin, F.‐J., Wang, J. et al. Conserved regulatory elements in AMPK. activates 13 kinases of the AMPK subfamily, including MARK/PAR‐1.
Nature, 2013;498:E8–10. EMBO J, 2004;23:833–43.
13. Li, X., Wang, L., Zhou, X.E. et al. Structural basis of AMPK regulation by 39. Al‐Hakim, A.K., Goransson, O., Deak, M. et  al. 14‐3‐3 cooperates with
adenine nucleotides and glycogen. Cell Res, 2015.25:50–66. LKB1 to regulate the activity and localization of QSK and SIK. J Cell Sci,
14. Xiao, B., Sanders, M.J., Underwood, E. et  al. Structure of mammalian 2005;118:5661–73.
AMPK and its regulation by ADP. Nature, 2011;472:230–3. 40. Anderson, K.A., Ribar, T.J., Lin, F. et al. Hypothalamic CaMKK2 contrib-
15. Xiao, B., Sanders, M.J., Carmena, D. et al. Structural basis of AMPK regula- utes to the regulation of energy balance. Cell Metab, 2008;7:377–88.
tion by small molecule activators. Nat Commun, 2013;4:3017. 41. Carling, D., Zammit, V.A., and Hardie, D.G. A common bicyclic protein
16. Xin, F.‐J., Wang, J., Zhao, R.‐Q., Wang, Z.‐X., and Wu, J.‐W. Coordinated kinase cascade inactivates the regulatory enzymes of fatty acid and choles-
regulation of AMPK activity by multiple elements in the α‐subunit. Cell Res, terol biosynthesis. FEBS Lett, 1987;223:217–22.
2013;23:1237–40. 42. Saggerson, D. Malonyl‐CoA a key signaling molecule in mammalian cells.
17. McBride, A., Ghilagaber, S., Nikolaev, A., and Hardie, D.G. The glycogen‐ Annu Rev Nutr, 2008;28:253–72.
binding domain on the AMPKβ subunit allows the kinase to act as a glycogen 43. Zang, M., Xu, S., Maitland‐Toolan, K.A. et al. Polyphenols stimulate AMP‐
sensor. Cell Metab, 2009;9:23–34. activated protein kinase, lower lipids, and inhibit accelerated atherosclerosis
18. Yamauchi, T., Kamon, J., Minokoshi, Y. et al. Adiponectin stimulates glu- in diabetic LDL receptor‐deficient mice. Diabetes, 2006;55:2180–91.
cose utilization and fatty-acid oxidation by activating AMP-activated protein 44. Sato, R., Goldstein, J.L., and Brown, M.S. Replacement of serine‐871 of
kinase. Nat Med, 2002;8:1288–95. hamster 3‐hydroxy‐3‐methylglutaryl‐CoA reductase prevents phosphoryla-
19. Hardie, D.G. AMP‐activated protein kinase as a drug target. Annu Rev tion by AMP‐activated kinase and blocks inhibition of sterol synthesis
Pharmacol Toxicol, 2007;47:185–210. induced by ATP depletion. Proc Natl Acad Sci USA, 1993;90:9261–5.
20. Sakamoto, K., McCarthy, A., Smith, D. et al. Deficiency of LKB1 in skeletal 45. Yang, W., Hong, Y.H., Shen, X.Q., Frankowski, C., Camp, H.S., and Leff, T.
muscle prevents AMPK activation and glucose uptake during contraction. Regulation of transcription by AMP‐activated protein kinase: phosphoryla-
EMBO J, 2005;24:1810–20. tion of p300 blocks its interaction with nuclear receptors. J Biol Chem,
21. Zhou, G., Myers, R., Li, Y. et al. Role of AMP‐activated protein kinase in 2001;276:38341–4.
mechanism of metformin action. J Clin Invest, 2001;108:1167–74. 46. Koo, S.H. et al. The CREB co‐activator TORC2 is a key regulator of fasting
22. Fryer, L.G., Parbu‐Patel, A., and Carling, D. The anti‐diabetic drugs rosigli- glucose metabolism. Nature, 2005;437:1109–11.
tazone and metformin stimulate AMP‐activated protein kinase through dis- 47. Liu, Y., Dentin, R., Chen, D. et al. A fasting inducible switch modulates glu-
tinct signaling pathways. J Biol Chem, 2002;277:25226–32. coneogenesis via activator/coactivator exchange. Nature, 2008;456:269–73.
23. Saha, A.K., Avilucea, P.R., Ye, J.M., Assifi, M.M., Kraegen, E.W., and 48. Bittinger, M.A., McWhinnie, E., Meltzer, J. et  al. Activation of cAMP
Ruderman, N.B. Pioglitazone treatment activates AMP‐activated protein response element‐mediated gene expression by regulated nuclear transport
kinase in rat liver and adipose tissue in vivo. Biochem Biophys Res Commun, of TORC proteins. Curr Biol, 2004;14:2156–61.
2004;314:580–5. 49. Screaton, R.A., Conkright, M.D., Katoh, Y. et  al. The CREB coactivator
24. Witters, L.A. The blooming of the French lilac. J Clin Invest, TORC2 functions as a calcium‐ and cAMP‐sensitive coincidence detector.
2001;108:1105–7. Cell, 2004;119:61–74.
25. Tan, M.J., Ye, J.M., Turner, N. et al. Antidiabetic activities of triterpenoids 50. Andreelli, F., Foretz, M., Knauf, C. et al. Liver adenosine monophosphate‐
isolated from bitter melon associated with activation of the AMPK pathway. activated kinase‐alpha2 catalytic subunit is a key target for the control of
Chem Biol, 2008;15:263–73. hepatic glucose production by adiponectin and leptin but not insulin.
26. Baur, J.A., Pearson, K.J., Price, N.L. et al. Resveratrol improves health and Endocrinology, 2006 147, 243241.
survival of mice on a high‐calorie diet. Nature, 2006;444:337–42. 51. Leclerc, I., Lenzner, C., Gourdon, L., Vaulont, S., Kahn, A., and Viollet, B.
27. Ingebritsen, T.S., Lee, H.S., Parker, R.A., and Gibson, D.M. Reversible modula- Hepatocyte nuclear factor‐4a involved in type 1 maturity‐onset diabetes of the
tion of the activities of both liver microsomal hydroxymethylglutaryl coenzyme young is a novel target of AMP‐activated protein kinase. Diabetes, 2001;50.
A reductase and its inactivating enzyme. Evidence for regulation by phospho- 52. Hong, Y.H., Varanasi, U.S., Yang, W., and Leff, T. AMP‐activated protein
rylation‐dephosphorylation. Biochem Biophys Res Commun, 1978;81:1268–77. kinase regulates HNF4alpha transcriptional activity by inhibiting dimer for-
28. Hawley, S.A., Davison, M., Woods, A. et al. Characterization of the AMP‐ mation and decreasing protein stability. J Biol Chem, 2003;278:27495–501.
activated protein kinase from rat liver and identification of threonine 172 as 53. Doria, A., Patti, M.E., and Kahn, C.R The emerging genetic architecture of
the major site at which it phosphorylates AMP‐activated protein kinase. J type 2 diabetes. Cell Metab, 2008;8:186–200.
Biol Chem, 1996;271:27879–87. 54. Hayhurst, G. P., Lee, Y.H., Lambert, G., Ward, J.M., and Gonzalez, F.J.
29. Hong, S.P., Leiper, F.C., Woods, A., Carling, D., and Carlson, M. Activation Hepatocyte nuclear factor 4a (nuclear receptor 2a1) is essential for mainte-
of yeast Snf1 and mammalian AMP‐activated protein kinase by upstream nance of hepatic gene expression and lipid homeostasis. Mol Cell Biol,
kinases. Proc Natl Acad Sci USA, 2003;100:8839–43. 2001:1393–1403.
30. Sutherland, C.M., Hawley, S.A., McCartney, R.R. et  al. Elm1p is one of 55. Inoue, E. and Yamauchi, J. AMP‐activated protein kinase regulates PEPCK
three upstream kinases for the Saccharomyces cerevisiae SNF1 complex. gene expression by direct phosphorylation of a novel zinc finger transcrip-
Curr Biol, 2003;13:1299–305. tion factor. Biochem Biophys Res Commun, 2006;351:793–9.
38:  CENTRAL REGULATOR OF GLUCOSE AND LIPID METABOLISM AND TARGET OF TYPE 2 DIABETES THERAPEUTICS 483

56. Kawaguchi, T., Osatomi, K., Yamashita, H., Kabashima T., and Uyeda, K. 81. Manning, B.D. Balancing Akt with S6K: implications for both metabolic
Mechanism of fatty acid “sparing” effect on glucose‐induced transcription: diseases and tumorigenesis. J Cell Biol, 2004;167:399–403.
regulation of carbohydrate‐responsive element‐binding protein by AMP‐ 82. Um, S.H., Frigerio, F., Watanabe, M. et  al. Absence of S6K1 protects
activated protein kinase. J Biol Chem, 2002;277:3829–35. against age‐ and diet‐induced obesity while enhancing insulin sensitivity.
57. Foretz, M., Ancellin, N., Andreelli, F. et al. Short‐term overexpression of a Nature, 2004;431:200–5.
constitutively active form of AMP‐activated protein kinase in the liver leads 83. Tzatsos, A. and Kandror, K.V. Nutrients suppress phosphatidylinositol 3‐
to mild hypoglycemia and fatty liver. Diabetes, 2005;54:1331–9. kinase/Akt signaling via raptor‐dependent mTOR‐mediated insulin recep-
58. Shimomura, I. et  al. Decreased IRS‐2 and increased SREBP‐1c lead to tor substrate 1 phosphorylation. Mol Cell Biol, 2006;26:63–76.
mixed insulin resistance and sensitivity in livers of lipodystrophic and ob/ob 84. Jakobsen, S.N., Hardie, D.G., Morrice, N., and Tornqvist, H.E 5’‐AMP‐
mice. Mol Cell, 2000;6:77–86. activated protein kinase phosphorylates IRS‐1 on Ser‐789 in mouse C2C12
59. Kakuma, T. et al. Leptin, troglitazone, and the expression of sterol regulatory myotubes in response to 5‐aminoimidazole‐4‐carboxamide riboside. J Biol
element binding proteins in liver and pancreatic islets. Proc Natl Acad Sci Chem, 2001;276:46912–6.
USA, 2000;97:8536–41. 85. Egan, D.F., Shackelford, D.B., Mihaylova, M.M. et al. Phosphorylation of
60. Zang, M., Zuccollo, A., Hou, X. et  al. AMP‐activated protein kinase is ULK1 (hATG1) by AMP‐activated protein kinase connects energy sensing
required for the lipid‐lowering effect of metformin in insulin‐resistant human to mitophagy. Science, 2011;331:456–61.
HepG2 cells. J Biol Chem, 2004;279:47898–905. 86. Kim, J., Kundu, M., Viollet, B., and Guan, K.‐L. AMPK and mTOR regu-
61. Li, Y., Xu, S., Mihaylova, M. et  al. AMPK phosphorylates and inhibits late autophagy through direct phosphorylation of Ulk1. Nat Cell Biol,
SREBP activity to attenuate hepatic steatosis and atherosclerosis in diet‐ 2011;13:132–41.
induced insulin resistant mice. Cell Metab, 2011;13:376–88. 87. Mack, H.I.D., Zheng, B., Asara, J.M., and Thomas, S.M. AMPK‐dependent
62. McGee, S.L., van Denderen, B.J., Howlett, K.F. et al. AMP‐activated protein phosphorylation of ULK1 regulates ATG9 localization. Autophagy,
kinase regulates GLUT4 transcription by phosphorylating histone deacety- 2012;8:1197–1214.
lase 5. Diabetes, 2008;57:860–7. 88. Kim, J., Kim, Y.C., Fang, C. et al. Differential regulation of distinct Vps34 com-
63. Chang, S., Bezprozvannaya, S., Li, S., and Olson, E.N. An expression screen plexes by AMPK in nutrient stress and autophagy. Cell, 2013;152:290–303.
reveals modulators of class II histone deacetylase phosphorylation. Proc 89. Weerasekara, V.K., Panek, D.J., Broadbent, D.G. et  al. Metabolic‐stress‐
Natl Acad Sci USA, 2005;102:8120–5. induced rearrangement of the 14–3‐3ζ interactome promotes autophagy via
64. Mihaylova, M.M., Vasquez, D.S., Ravnskjaer, K. et al. Class IIa histone dea- a ULK1‐ and AMPK‐regulated 14–3‐3ζ interaction with phosphorylated
cetylases are hormone‐activated regulators of FOXO and mammalian glu- Atg9. Mol Cell Biol, 2014;34:4379–88.
cose homeostasis. Cell, 2011;145:607–21. 90. Zhou, C., Ma, K., Gao, R. et al. Regulation of mATG9 trafficking by Src‐
65. Wullschleger, S., Loewith, R., and Hall, M.N. TOR signaling in growth and and ULK1‐mediated phosphorylation in basal and starvation‐induced
metabolism. Cell, 2006;124:471–84. autophagy. Cell Res, 2017;27:184–201.
66. Guertin, D.A. and Sabatini, D.M. Defining the role of mTOR in cancer. 91. Young, N.P., Kamireddy, A., Van Nostrand, J.L. et al. AMPK governs line-
Cancer Cell, 2007;12:9–22. age specification through Tfeb‐dependent regulation of lysosomes. Genes
67. Holz, M.K., Ballif, B.A., Gygi, S.P., and Blenis, J. mTOR and S6K1 mediate Dev, 2016;30:535–52.
assembly of the translation preinitiation complex through dynamic protein 92. Greer, E.L., Oskoui, P.R., Banko, M.R. et al. The energy sensor AMP‐acti-
interchange and ordered phosphorylation events. Cell 2005;123:569–80. vated protein kinase directly regulates the mammalian FOXO3 transcrip-
68. Sancak, Y., Thoreen, C.C., Peterson, T.R. et al. PRAS40 is an insulin‐regu- tion factor. J Biol Chem, 2007;282:30107–19.
lated inhibitor of the mTORC1 protein kinase. Mol Cell, 2007;25:903–15. 93. Shin, H.‐J.R., Kim, H., Oh, S. et al. AMPK–SKP2–CARM1 signalling cas-
69. Shaw, R.J. and Cantley, L.C. Ras, PI(3)K and mTOR signalling controls cade in transcriptional regulation of autophagy. Nature, 2016;534:553–7.
tumour cell growth. Nature, 2006;441:424–30. 94. Toyama, E.Q., Herzig, S., Courchet, J. et al. Metabolism. AMP‐activated
70. Vander Haar, E., Lee, S.I., Bandhakavi, S., Griffin, T.J., and Kim, D.H. protein kinase mediates mitochondrial fission in response to energy stress.
Insulin signalling to mTOR mediated by the Akt/PKB substrate PRAS40. Science, 2016;351:275–81.
Nat Cell Biol, 2007;9:316–23. 95. Jäger, S., Handschin, C., St‐Pierre, J., and Spiegelman, B.M. AMP‐acti-
71. Inoki, K., Zhu, T., and Guan, K.L. TSC2 mediates cellular energy response vated protein kinase (AMPK) action in skeletal muscle via direct phospho-
to control cell growth and survival. Cell, 2003;115:577–90. rylation of PGC‐1alpha. PNAS, 2007;104:12017–22.
72. Corradetti, M.N., Inoki, K., Bardeesy, N., DePinho, R.A., and Guan, K.L 96. Cantó, C., Gerhart‐Hines, Z., Feige, J.N. et  al. AMPK regulates energy
Regulation of the TSC pathway by LKB1: evidence of a molecular link expenditure by modulating NAD+ metabolism and SIRT1 activity. Nature,
between tuberous sclerosis complex and Peutz–Jeghers syndrome. Genes 2009;458:1056–60.
Dev, 2004;18:1533–8. 97. Mansueto, G., Armani, A., Viscomi, C. et al. Transcription factor EB con-
73. Shaw, R.J., Bardeesy, N., Manning, B.D. et al. The LKB1 tumor suppressor trols metabolic flexibility during exercise. Cell Metab, 2017;25:182–96.
negatively regulates mTOR signaling. Cancer Cell, 2004;6:91–9. 98. Wada, S., Neinast, M., Jang, C., et al. The tumor suppressor FLCN medi-
74. Liu, L., Cash, T.P., Jones, R.G., Keith, B., Thompson, C.B., and Simon, M.C. ates an alternate mTOR pathway to regulate browning of adipose tissue.
Hypoxia‐induced energy stress regulates mRNA translation and cell growth. Genes Dev, 2016;30:2551–64.
Mol Cell, 2006;21:521–31. 99. Baas, A.F. et al. Complete polarization of single intestinal epithelial cells
75. Gwinn, D.M., Shackelford, D.B., Egan, D.F. et al. AMPK phosphorylation upon activation of LKB1 by STRAD. Cell, 2004;116:457–66.
of raptor mediates a metabolic checkpoint. Mol Cell, 2008;30:214–26. 100. Fu, D., Wakabayashi, Y., Ido, Y., Lippincott‐Schwartz, J., and Arias, I.M.
76. Kimball, S.R., Siegfried, B.A., and Jefferson, L.S. Glucagon represses sign- Regulation of bile canalicular network formation and maintenance by
aling through the mammalian target of rapamycin in rat liver by activating AMP‐activated protein kinase and LKB1. J Cell Sci, 2010;123:3294–302.
AMP‐activated protein kinase. J Biol Chem, 2004;279:54103–9. 101. Fu, D., Wakabayashi, Y., Lippincott‐Schwartz, J., and Arias, I.M. Bile acid
77. Wang, C., Mao, X., Wang, L. et al. Adiponectin sensitizes insulin signaling stimulates hepatocyte polarization through a cAMP‐Epac‐ MEK‐LKB1‐
by reducing p70 S6 kinase‐mediated serine phosphorylation of IRS‐1. J Biol AMPK pathway. Proc Natl Acad Sci USA, 2011;108:1403–8.
Chem, 2007;282:7991–6. 102. Homolya, L., Fu, D., Sengupta, P. et al. LKB1/AMPK and PKA control
78. Howell, J.J., Hellberg, K., Turner, M. et  al. Metformin inhibits hepatic ABCB11 Trafficking and polarization in hepatocytes. PLoS One, 2014:9.
mTORC1 signaling via dose‐dependent mechanisms involving AMPK and 103. Zhang, L., Li, J., Young, L.H., and Caplan, M.J. AMP‐activated protein
the TSC complex. Cell Metab, 2017;25:463–71. kinase regulates the assembly of epithelial tight junctions. Proc Natl Acad
79. Kimball, S.R. and Jefferson, L.S. Induction of REDD1 gene expression in Sci USA, 2006;103:17272–7.
the liver in response to endoplasmic reticulum stress is mediated through a 104. Zheng B. and Cantley, L.C. Regulation of epithelial tight junction assembly
PERK eIF2α phosphorylation, ATF4‐dependent cascade. Biochem Biophys and disassembly by AMP‐activated protein kinase. Proc Natl Acad Sci
Res Commun, 2012;427:485–9. USA, 2007;104:819–22.
80. Porstmann, T., Santos, C.R., Griffiths, B. et al. SREBP activity is regulated 105. Yano, T., Matsui, T., Tamura, A., Uji, M., and Tsukita, S. The association of
by mTORC1 and contributes to Akt‐dependent cell growth. Cell Metab, microtubules with tight junctions is promoted by cingulin phosphorylation
2008;8:224–36. by AMPK. J Cell Biol, 2013;203:605–14.
484 THE LIVER:  REFERENCES

106. Aznar, N., Patel, A., Rohena, C.C. et al. AMP‐activated protein kinase forti- 112. An, H. and He, L. Current understanding of metformin effect on the control
fies epithelial tight junctions during energetic stress via its effector GIV/ of hyperglycemia in diabetes. J Endocrinol, 2016;228:R97–106.
Girdin. Elife, 2016;5:e20795. 113. Smith, B.K. Marcinko, K., Desjardins, E.M., Lally, J.S., Ford, R.J., and
107. Enomoto, A., Murakami, H., Asai, N. et al. Akt/PKB regulates actin organi- Steinberg, G.R. Treatment of nonalcoholic fatty liver disease: role of
zation and cell motility via Girdin/APE. Dev Cell, 2005;9:389–402.111. AMPK. Am J Physiol Endocrinol Metab, 2016;311:E730–40.
108. Mihaylova, M.M. and Shaw, R.J. The AMPK signalling pathway coordinates 114. Quinn, B.J., Kitagawa, H., Memmott, R.M., Gills, J.J., and Dennis, P.A.
cell growth, autophagy, and metabolism. Nat Cell Biol, 2011;13:1016–23. Repositioning metformin for cancer prevention and treatment. Trends
109. Owen, M.R., Doran, E., and Halestrap, A.P. Evidence that metformin exerts Endocrinol Metab, 2013;24:469–80.
its anti‐diabetic effects through inhibition of complex 1 of the mitochon- 115. Myers, R.W. et al. Systemic pan‐AMPK activator MK‐8722 improves glucose
drial respiratory chain. Biochem J, 2000;348(3):607–14. homeostasis but induces cardiac hypertrophy. Science, 2017;357:507–11.
110. Fullerton, M.D., Galic, S., Marcinko, K. et al. Single phosphorylation sites 116. Cokorinos, E.C. et al. Activation of skeletal muscle AMPK promotes glu-
in Acc1 and Acc2 regulate lipid homeostasis and the insulin‐sensitizing cose disposal and glucose lowering in non‐human primates and mice. Cell
effects of metformin. Nat Med, 2013;19:1649–54. Metab. 2017;25:1147–59.
111. Foretz, M. et al. Metformin inhibits hepatic gluconeogenesis in mice inde-
pendently of the LKB1/AMPK pathway via a decrease in hepatic energy
state. J Clin Invest, 2010;120:2355–69.
Insulin‐Mediated PI3K
39 and AKT Signaling
Hyokjoon Kwon1 and Jeffrey E. Pessin2,3
1
Department of Medicine, Division of Endocrinology, Metabolism and Nutrition, Rutgers‐Robert Wood
Johnson Medical School, New Brunswick, NJ, USA
2
Albert Einstein‐Mount Sinai Diabetes Research Center and the Fleischer Institute for Diabetes and
Metabolism, Albert Einstein College of Medicine, Bronx, NY, USA
3
Department of Medicine and Department of Molecular Pharmacology, Albert Einstein College of Medicine,
Bronx, NY, USA

INTRODUCTION CHARACTERISTICS OF PI3K AND AKT


Obesity is a pandemic in modern society and closely linked PI3K and AKT/PKB are key signaling molecules, mediating
with diverse metabolic diseases such as cardiovascular dis­ insulin signaling in diverse tissues including liver, muscle, and
ease, type 2 diabetes mellitus (T2D) and non‐alcoholic fatty adipose tissue. In general, the intracellular responses to extracel­
liver disease (NAFLD). Thus, the cost of managing obesity lular signals are mediated by small second‐messenger molecules
and related metabolic diseases is a burden on public health such as cAMP, Ca2+, and lipid molecules that are responsible for
care systems in modern society. T2D is a quickly growing transducing signals between the extracellular environment and
global metabolic disorder characterized by impaired insulin intracellular compartments. In this pathway, PI3K activation
secretion from pancreatic β cells and insulin resistance in induces the formation of insotitol‐3,4,5‐trisphosphate (PI(3,4,5)
peripheral tissues such as liver, muscle, and adipose tissue. P3) that serves as a second‐messenger to activate AKT in the
Insulin resistance diminishes insulin‐stimulated glucose insulin signaling pathway, resulting in glucose disposal, glyco­
uptake in muscle and glycogen synthesis in liver and also gen synthesis, and suppression of lipolysis. To understand the
impairs insulin‐mediated suppression of gluconeogenesis in physiologic role of these signaling molecules in insulin action
liver, resulting in hyperglycemia and hyperinsulinemia. As and glucose homeostasis, in the following sections we review the
insulin is a crucial endocrine hormone for modulating glucose biochemical characteristics of the PI3K and AKT kinases.
homeostasis, the molecular mechanism of insulin signaling in
glucose homeostasis has been a central theme in biomedical
research. Insulin signaling is initiated by the activation of the
Phosphoinositide 3‐kinase (PI3K)
insulin receptor (IR) through autophosphorylation of the Phosphatidylinositol has free hydroxyl groups that are
tyrosine residue in the IR, and then many signaling molecules ­phosphorylated by diverse series of kinases including phosphoi­
including insulin receptor substrate 1 and 2 (IRS1/2), phosph­ nositide 3‐kinase (PI3K) for the generation of crucial second‐
oinositide‐3‐kinase (PI3K) and AKT/protein kinase B (PKB) messenger in cell signaling. Early studies demonstrated that the
are involved to modulate downstream signaling pathways. In cleavage of phosphatidylinositol‐4,5‐bisphosphate (PI(4,5)P2)
this chapter, we will focus on the role of PI3K and AKT/PKB by a membrane bound phospholipase C (PLC) generated diacylg­
in insulin signaling related to pathophysiology in T2D, lycerol (DAG) and inositol‐1,4,5‐trisphosphate. Diacylglycerol
NAFLD, and liver cancer. activates protein kinase C (PKC), and inositol‐1,4,5‐trisphosphate

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
486 THE LIVER:  CHARACTERISTICS OF PI3K AND AKT

(a) Class 1A
Catalytic p110α/β/δ ABD RBD C2 Helical Kinase

Regulatory p85α/β SH3 BH N-SH2 iSH2 C-SH2

p55α/p50α N-SH2 iSH2 C-SH2

p55γ N-SH2 iSH2 C-SH2

Class 1B
Catalytic p110γ RBD C2 Helical Kinase

Regulatory p101 GβγBD

p84

(b) T308 T450 S473


AKT1 1 PH Kinase Domain HD 480

T309 T451 S474


AKT2 1 PH Kinase Domain HD 481

T305 T447 S472


AKT3 1 PH Kinase Domain HD 479

Figure 39.1  Schematic structure of PI3K and AKT/PKB. (a) Domain structure of class I PI3K catalytic and regulatory subunits. ABD, adaptor‐
binding domain; RBD, Ras‐binding domain; SH2, Src homology‐2; BH, breakpoint cluster region homology; iSH, coiled‐coil inter‐SH2.
(b) Domain structure AKT isoforms. PH, pleckstrin homology domain; HD, hydrophobic domain.

induces Ca2+ influx from intracellular calcium stores. In contrast, heterodimer with Vps15/p150 to locate on the intracellular
PI3K mediates phosphorylation of phosphoinositides at the membrane, and the biological function of Vps34 in mammals
D3  position to generate various 3‐phosphorylated phosphoi­ relates to the regulation of vesicle trafficking such as autophagy,
nositides such as phosphatidyl‐3,5‐bisphosphate (PI(3,5)P2) and endocytosis, and phagocytosis. In mammals, class I PI3K is
phosphatidyl‐3,4,5‐triphosphate (PI(3,4,5)P3) in the absence of present in all cell types and mediates insulin signaling. Class I
phospholipase‐mediated cleavage. In insulin signaling, PI3K PI3K consists of a heterodimer with a catalytic subunit (110–
generates plasma membrane‐bound PI(3,4,5)P3 to activate AKT 120 kDa) and regulatory subunit, and phosphatidylinositol‐4,
through both the binding of the AKT pleckstrin homology (PH) 5‐bisphosphate (PI(4,5)P2) is a preferred substrate in vivo.
domain to PI(3,4,5)P3 and AKT site‐specific phosphorylations Catalytic subunits contain C‐terminal catalytic and phosphati­
by the 3‐phosphoinositide‐dependent kinase 1 (PDK1) and by the dylinositol kinase domains, N‐terminal adaptor‐binding
mechanistic target of rapamycin (mTOR) in complex 2 (mTORC2). domain (ABD), and Ras‐binding domain (RBD) (Figure 39.1a).
PI3K is classified into three classes (class I, class II, and Low concentrations (nanomolar) of Wortmannin irreversibly
class III) according to structure and lipid substrate preferences inhibits the catalytic subunit of class I PI3K by Schiff base for­
[1]. Class II PI3K including PI3K‐C2α, PI3K‐C2β and PI3K‐ mation with a lysine in the C‐terminal kinase domain [2].
C2γ was discovered in mammals on the basis of sequence Regulatory subunits have C‐terminal Src homology‐2 (SH2)
homology with class I and III PI3K instead of functional con­ domains that are separated by an inter‐SH2 (iSH2) region to
text. Thus, their functional role is unclear although class II provide a binding site for the ABD of the catalytic subunit. The
PI3K is constitutively associated with intracellular membranes. SH2 domain is a module of about 100 amino acids to bind
Class II PI3K consists of two subclasses α and β and contains phosphotyrosine‐containing motifs. Regulatory subunits also
a C‐terminal C2 domain as observed in PKC molecules for have N‐terminal proline‐binding SH3 and a breakpoint cluster
phospholipid binding. Vps34 was originally identified in region homology domain (BH). Class I PI3K catalytic p110
Saccharomyces cerevisiae for the endosomal sorting and is the subunits are divided into a class IA group such as p110α,
only member of class III PI3K. Vps34 generates a constitutive p110β, and p110δ, which bind the regulatory p85 type subunit
39:  Insulin‐Mediated PI3K and AKT Signaling 487

and into a class IB p110γ to interact with regulatory p101 and cytosol. Indeed, mTORC2‐induced phosphorylation of Thr‐450 in
p84. The p110α and p110β are expressed ubiquitously, whereas the AKT1 turn motif localizes AKT1 to the cytosol as an inactive
p110γ and p110δ are expressed in immune cells. Each catalytic conformation through the interaction between PH and kinase
subunit generates a dimer with a regulatory subunit to modulate domains [13]. AKT activity is regulated through the phospho­
the catalytic activity and subcellular localization. The p110α rylation in kinase domain at Thr‐308 by PDK1 and in the hydro­
catalytic subunit generates a heterodimer with one of five dif­ phobic domain at Ser‐473 by mTORC2, necessary for the full
ferent regulatory subunits including p85α, p85δ, p55α, p55γ, induction of Akt substrate kinase activity [14].
and p50α. In the p85α/p110α heterodimer, the ABD of p110α
catalytic subunit interacts with coiled‐coil inter‐SH2 (iSH)
domain of the p85α regulatory subunit to maintain a stable het­
erodimer with a low kinase activity state [3]. Thus, binding of INSULIN‐MEDIATED PI3K AND AKT
the SH2 domain in a p85α regulatory subunit with phosphoty­ SIGNALING AND INSULIN RESISTANCE
rosine residues such as phosphotyrosine in IRS1/2 releases
inhibitory interactions between p85α and p110α, resulting in Glucose homeostasis is maintained by delicate regulation of the
the activation of PI3K in insulin signaling [4, 5]. pancreatic exocrine and endocrine systems. The pancreatic exo­
crine cells composed of acinar and ductal cells secrete digestive
enzymes into the duodenum for nutrient digestion. In contrast,
AKT/protein kinase B the pancreatic endocrine cells in the islets of Langerhans secrete
The v‐AKT oncogene known as AKT8 was identified from endocrine hormones into the blood circulation to regulate nutri­
transforming retrovirus in a spontaneous thymoma of AKR ent metabolism. Pancreatic islets include several endocrine cells
mice, and then cellular homologue serine and threonine kinase such as α cells (glucagon production), β cells (insulin produc­
AKT (roughly 57 kDa), also termed as protein kinase B (PKB), tion), δ cells (somatostatin production), PP cells (pancreatic
was cloned and characterized [6–9]. Human AKT has three polypeptide production), and ε cells (ghrelin production) for
isoforms (AKT1/PKBα, AKT2/PKBβ, AKT3/PKBγ), and specific endocrine functions. The secretion of glucagon and
each isoform is encoded from separate human genes: AKT1/ insulin is tightly regulated in response to circulating glucose
PKBα in Chr14 q32.33, AKT2/PKBβ in Chr19 q13.2, and levels to maintain normoglycemia during fasting and feeding,
AKT3/PKBγ in Chr1 q44 (Figure 39.1b). For functional stud­ respectively. In peripheral tissues, insulin stimulates glucose
ies, murine models demonstrate that each AKT isoform has uptake (skeletal muscle and adipose tissue) and glycogen syn­
distinct and overlapping signaling functions. AKT1 deficient thesis (skeletal muscle and liver) and inhibits gluconeogenesis
mice show reduced body size, AKT2 deficiency impairs glu­ and glycogenolysis (liver). Insulin also increases lipogenesis in
cose homeostasis, and AKT3 deficient mice show diminished hepatocytes and adipocytes and diminishes lipolysis in adipo­
brain size. In contrast, all AKT isoforms contribute to phos­ cytes to reduce circulating free fatty acid and glycerol [15].
phorylation of GSK3α and GSK3β in 3T3‐L1 adipocytes and Thus, impaired regulation of insulin secretion and IR‐mediated
regulation of forkhead box O1 (FOXO1), TSC2, and GSK3β signaling pathway results in T2D.
in H157 cells.
AKT kinases belong to a class of AMP/GMP kinase and protein
Insulin‐mediated PI3K and AKT signaling
kinase C (AGC) kinases. AKT has a PH domain, kinase domain,
and hydrophobic domain (HD) with PH and kinase domains con­ Identification and characterization of the IR initiated a huge
nected with an α‐helical linker. The PH domain found as a major effort to understand the molecular mechanisms of insulin action
phosphorylation site by PKC in pleckstrin proteins at its N‐termi­ in glucose homeostasis. At the molecular level, insulin binds to
nus interacts with PI(3,4,5)P3 to localize AKT into the plasma the cell surface IR to initiate intracellular signaling cascades
membrane and then to interact with 3‐phosphoinositide‐dependent that ultimately results in specific cellular biological responses
kinase 1 (PDK1) that also has a PH domain to be enriched in [16]. The IR is a transmembrane tyrosine kinase receptor
plasma membrane. The PH domain is a 100–120 amino acids motif encoded by Chr19 p13.2 in humans and has two isoforms, IR‐A
with seven anti‐parallel β‐sheets forming a hydrophobic pocket that and IR‐B, generated by alternative splicing of exon11. The IR‐A
interacts with the C‐terminal amphipathic helix. The PH domain is isoform excludes exon 11 and is predominantly expressed in
a primary lipid binding site although it is also involved in protein– fetal tissues and the brain with high affinity for insulin and insu­
protein interactions. As the PH domain has different phosphoi­ lin‐like growth factor 2 (IGF‐2), whereas IR‐B includes exon 11
nositide binding specificity, PH domains of dynamin bind to PI(4) and is highly expressed in liver. To interact with insulin, the
P1 and PI(4,5)P2, whereas PH domains of AKT and PDK1 bind to insulin receptor consists of two α subunits and two β subunits
PI(3,4)P2 and PI(3,4,5)P3 [10, 11]. The kinase domain of AKT that are bound by a disulfide link into an α2β2 heterotetrameric
shares high homology with other AGC kinase members, and the complex. The α and β subunits are derived from a single large
kinase domain of three AKT kinase isoforms displays around 87% precursor by one or more proteolytic cleavages. The extracellu­
sequence homology. HD is a characteristic feature of AGC kinases lar α subunits directly bind insulin that allows for a transmem­
including protein kinase A (PKA), protein kinase C (PKC), and brane conformational change that activates the intracellular
PDK1 and contains a F‐X‐X‐F/Y‐S/T‐Y/F motif where X is any tyrosine kinase domain of the IR β subunits, resulting in intra­
amino acid. Upon biosynthesis of AKT, the nascent AKT poly­ molecular transphosphorylation of the β subunit to phosphoryl­
peptide is phosphorylated at the ribosome by mTORC2 at the ate its adjacent partner on specific tyrosine residues [17, 18].
C‐terminal turn motif [12], enhancing the protein stability in Autophosphorylated tyrosine residues in the tyrosine kinase
488 THE LIVER:  INSULIN‐MEDIATED PI3K AND AKT SIGNALING AND INSULIN RESISTANCE

domain of β subunits recruit receptor substrates such as insulin Insulin‐mediated PI3K activation
receptor substrate (IRS) that is a scaffold for organizing the
proximal signaling complex. Phosphotyrosine residues in acti­ PI3K‐AKT/PKB signaling pathway modulates most metabolic
vated IRs interact with IRS through its phosphotyrosine binding functions of insulin. Activation of PI3K by extracellular stimu­
domain (PTB) [19]. IRS also has several tyrosine residues, lation including insulin induces the activation of AKT. In the
which are phosphorylated by activated IR tyrosine kinase activ­ basal state, p85α/p110α heterodimeric PI3K has interactions
ity, for the interaction with the SH2 domain of the regulatory between ABD of the p110α catalytic subunit and coiled‐coil
subunit of PI3K [5]. IRS isoforms (IRS1‐IRS4) show differen­ iSH domain of p85α regulatory subunit to maintain stable PI3K
tial function in physiology. IRS‐1 knockout mice have growth with a low kinase activity state. However, insulin‐induced phos­
retardation and insulin resistance particularly in muscle, albeit phorylation on tyrosine residue in IRS1/2 mediates interaction
with normal whole body glucose tolerance [20], whereas IRS‐2 between phosphotyrosine of IRS1/2 and the SH2 domain in
knockout mice show impaired insulin action in liver and pancre­ p85α regulatory subunit to release inhibitory interactions
atic β cell loss, resulting in T2D [21]. IRS1/2 phosphorylated on between p85α and p110α [5, 22]. Thus, PI3K accesses the
specific tyrosine residues activates two major signaling path­ ­membrane to produce phosphatidylinositol‐3,4,5‐triphosphate
ways; (i) the PI3K‐AKT/PKB pathway and (ii) Ras/mitogen‐ (PI(3,4,5)P3) from PI(4,5)P2 at the plasma membrane. Although
activated protein kinase (MAPK) pathway. In addition, there are p110α or p110β knockout mice are embryonic lethal, liver spe­
inhibitory molecules for insulin signaling such as protein tyros­ cific p110β deficient mice showed little effect in insulin signal­
ine phosphatase 1B (PTP1B), suppressor of cytokine signaling ing. However, liver specific p110α deficient mice have glucose
(SOCS), and growth factor receptor bound protein 10 (Grb10) intolerance and insulin resistance, suggesting that p110α is criti­
that suppress insulin signaling by inducing IR dephosphoryla­ cal to mediate insulin signaling in hepatocytes [23]. Deletion of
tion, physical blocking substrate phosphorylation, and degrada­ the p85 regulatory subunits in heterozygous deletion of p85α,
tion of the IR and/or IRS substrates. Metabolites such as DAG p85β, or p50α/p55α showed enhanced insulin sensitivity as
also mediate the inhibition of insulin signaling (Figure 39.2). ­regulatory subunits are in excess concentrations to catalytic
subunits, thereby competing with IRS for the formation of

Figure 39.2  Insulin signaling and insulin resistance. IRS1/2 phosphorylated on specific tyrosine residues activates the phosphoinositide 3‐kinase
(PI3K)‐AKT/protein kinase B (PKB) pathway and Ras/mitogen‐activated protein kinase (MAPK) pathway. PI3K‐AKT signaling pathway regulates
metabolic processes such as glycogen synthesis (muscle and liver), glucose uptake (muscle and adipocytes), protein synthesis (muscle and liver),
and gluconeogenesis (liver). Inflammatory signals such as TNF‐α, saturated free fatty acid, IL‐6, LPS, and diacylglycerol (DAG) activate inhibitory
molecules such as suppressor of cytokine signaling (SOCS) and cJun N‐terminal kinase (JNK) to suppress insulin signaling, resulting in insulin
resistance.
39:  Insulin‐Mediated PI3K and AKT Signaling 489

p85α/p110α heterodimer [24]. PI(3,4,5)P3 generated by insulin‐ that induces FOXO1 association with nuclear 14‐3‐3 proteins to
induced activation of PI3K is metabolized rapidly into PI(4,5)P2 exclude FOXO1 from the nucleus to the cytosol in an inactive
by lipid phosphatase including tumor suppressor phosphatase state. In the liver, this suppresses gluconeogenic gene expres­
and tensin homologue (PTEN) and SH2‐containing inositol sion and thereby inhibits hepatic glucose production in feeding.
5′‐phosphatase‐2 (SHIP2) to terminate proximal signaling [25, AKT phosphorylates tuberous sclerosis complex 1 and 2
26]. PTEN encoded in chromosome 10q23 was identified as a (TSC1/2), which releases the inhibition of Ras homologue
tumor suppressor that is inactivated in diverse tumors including enriched in brain (Rheb) for the activation of mTORC1 complex
endometrial, prostate, and mammary carcinomas. [14], that in turn enhances protein synthesis through the acti­
vation of eukaryotic translation initiation factor 4E binding
Insulin‐mediated AKT activation protein‐1 (4E‐BP) and p70 ribosomal protein S6 kinase 1
(p70S6K1). Furthermore, mTORC1 activation induces
PI(3,4,5)P3 generated by PI3K triggers the activation of 3‐phos­ SREBP1c activation for the expression of lipogenic genes
phoinositide‐dependent protein kinase 1 (PDK1) that is respon­ including FAS and ACC to induce lipogenesis in liver [31].
sible for the phosphorylation and activation of AKT. PDK1 Fasting‐induced β‐adrenergic receptor signaling activates PKA
contains two domains, an N‐terminal kinase domain and a C‐ to phosphorylate perilipin1 (PLIN1) and hormone sensitive
terminal PH domain. Autophosphorylation of Ser‐241 in kinase lipase (HSL), facilitating lipolysis in adipocytes. In feeding,
domain by PDK1 is required for kinase activity, and the small however, insulin‐induced AKT activation mediates phospho­
PH domain binds to membrane bound PI(3,4,5)P3 and PI(3,4)P2. rylation of phosphodieterase‐3B (PDE‐3B) to decrease cAMP,
As AKT also has a PH domain to interact with membrane‐bound resulting in the suppression of PKA activity and HSL activity
PI(3,4,5)P3, AKT is subsequently recruited from the cytosol to in adipocytes [32]. Thus, insulin‐induced suppression of
the plasma membrane through binding of its PH domain to ­adipose tissue lipolysis mediates acute suppression of gluconeo­
PI(3,4,5)P3, resulting in a conformational change that separates genesis in liver as suppressed lipolysis reduces acetyl‐CoA
the PH and kinase domains from inactive conformation and ­levels in liver [33, 34].
exposes two key regulatory residues whose phosphorylations
are required for maximal activation of AKT kinase. The acti­
vated PKD1 closely localized with AKT through PI(3,4,5)P3 Insulin resistance in liver
binding phosphorylates Thr‐308 in the activation loop of AKT The molecular mechanisms accounting for insulin resistance are
[13]. AKT is also phosphorylated on Ser‐473 by the mTORC2, still unclear although there is substantial progress in our under­
and this dual phosphorylation results in full activation of AKT standing of insulin signaling. Insulin resistance is the integral
kinase activity [14]. Deletion of mTORC2‐specific subunits result of alterations in insulin secretion in pancreatic β cells, IR
such as Rictor or Sin1 abrogates AKT phosphorylation at both expression, ligand binding, and downstream IR signaling,
the C‐terminal turn motif (Thr‐450 of AKT1) and HD (Ser‐473 resulting in diverse metabolic disease such as T2D and cardio­
of AKT1), resulting in the impaired proximal signaling of AKT vascular and NAFLD diseases. Liver is one of major tissues for
[14, 27]. Interestingly, the AKT isoforms show differential func­ regulating glucose homeostasis in response to pancreatic hor­
tions according to the distinctive expression profiles in tissues. mones such as insulin and glucagon produced by pancreatic β
AKT1 and AKT2 are broadly expressed with AKT2 being more and α cells, respectively. Insulin‐mediated PI3K and AKT acti­
closely linked to metabolic processes. AKT1 deficient mice vation regulate hepatic glucose and lipid metabolism. In insulin
have growth retardation without defects in metabolism. In sensitive individuals, regular meals increase the blood glucose
contrast, AKT2 deficient mice show insulin resistance due to that in turn releases pancreatic insulin to activate insulin signal­
interrupted insulin signaling [28]. ing and glucose uptake through IR and GLUT4 in muscle and
adipose tissue, respectively. In the liver, PI3K and AKT activa­
Proximal signaling of insulin‐induced tion enhance glycogen synthesis and de novo lipogenesis (DNL)
and also suppresses gluconeogenesis through FOXO inactiva­
AKT activation
tion (Figure  39.3a). Thus, hepatic insulin signaling may sup­
Activated AKT/PKB regulates diverse insulin‐mediated meta­ press gluconeogenesis through transcriptional suppression of
bolic pathways such as glucose transport, glycogen synthesis, gluconeogenic genes such as PCK1 and G6PC for long‐term
gluconeogenesis, protein synthesis, and cell growth. Several regulation. However, the ability of insulin to acutely suppress
AKT substrates have been identified by the AKT consensus gluconeogenesis is mostly mediated by an inhibition of adipose
motif (R‐X‐R‐X‐X‐S/T‐B) where X is any amino acid, and B tissue lipolysis [34]. First, insulin reduces hepatic glucose pro­
represents bulky hydrophobic amino acids [29]. AKT promotes duction within 30 minutes but does not reduce gluconeogenic
cell growth and proliferation through the suppression of p27 to protein levels [35]. Second, insulin suppresses adipose tissue
activate cyclin D1 along with the activation of MAPK signaling lipolysis to regulate hepatic gluconeogenesis as acetyl‐CoA
pathway (Figure 39.2). AKT phosphorylates AKT substrate of (Ac‐CoA) generated by adipose tissue lipolysis allosterically
160 kDa (AS160) to activate the Rab family of small GTPases activates pyruvate carboxylase (PyC) activity for hepatic gluco­
that initiate the translocation of the glucose transporter 4 neogenesis. Thus, suppressed adipose tissue lipolysis lowers
(GLUT4), resulting in glucose uptake in muscle and adipocytes. hepatic acetyl‐CoA and glycerol contents to reduce PyC activity
AKT also suppresses glycogen synthase kinase‐3 (GSK3) via and glucose production [33, 34, 36]. In contrast, in insulin
phosphorylation on Ser‐21 or Ser‐9 to activate glycogen syn­ resistance status, glycogenolysis and gluconeogenesis are
thase in muscle and liver [30]. AKT phosphorylates FOXO1 enhanced to produce hepatic glucose due to dysfunctional
490 THE LIVER:  INSULIN‐MEDIATED PI3K AND AKT SIGNALING AND INSULIN RESISTANCE

(a)

(b)

Figure 39.3  Insulin resistance in liver. (a) In insulin sensitive individuals, insulin derived from pancreatic β cells activates AKT to enhance gly­
cogen synthesis, de novo lipogenesis, and suppress gluconeogenesis through FOXO1 degradation. (b) In insulin resistance, impaired insulin signal­
ing increases glycogenolysis and gluconeogenesis, and flux of free fatty acids and glycerol from adipose tissue and intestine results in glucose
production and fat accumulation in liver. DNL, de novo lipogenesis; GS, glycogen synthase; GP, glycogen phosphorylase; PyC, pyruvate carboxy­
lase; Ac‐CoA, acetyl‐CoA; FA‐CoA, fatty acyl‐CoA; TAG, triacylglycerol; CM, chylomicron; VLDL, very low density lipoprotein.

insulin signaling (Figure  39.3b). Impaired activation of PI3K triglyceride lipase (ATGL) deficient mice that have reduced
and AKT induces activation of glycogen phosphorylase (GP) to ability to convert triglyceride to DAG show enhanced glucose
induce glycogenolysis along with suppressed glycogen syn­ tolerance and insulin sensitivity [42]. More recently, an alterna­
thase (GS) activity. Impaired AKT‐mediated phosphorylation of tive model of increased ceramide levels has also been shown to
FOXO1 also generates nuclear active FOXO1 to mediate hepatic be associated with insulin resistance [43]. Decreased IRS pro­
gluconeogenesis. Flux of free fatty acids and glycerol from adi­ tein levels also contribute to insulin resistance in rodents and
pose tissue and intestine supplies substrate for glucose produc­ humans although complete molecular understanding of the
tion and fat accumulation in liver, resulting in hyperglycemia mechanisms of reduced IRS levels are still under investigation
and hyperlipidemia. [44]. SOCS1/3 increased by obesity‐induced inflammatory
Hepatic insulin resistance is reproducibly correlated with cytokines such as TNF‐α and IL‐6 enhances the degradation of
increased hepatic triglyceride content as shown in NAFLD. IRS1/2 through E3 ubiquitin ligase activation, resulting in insu­
High levels of DAG generated by incomplete synthesis to tri­ lin resistance [45, 46]. IRS phosphorylation on serine residues is
glyceride or the breakdown of triglyceride to DAG have been another mechanism that induces insulin resistance as the IRS
proposed to inhibit insulin signaling through protein kinase proteins contain several serine residues that are phosphorylated
Cε (PKCε) activation, which phosphorylates IR at Thr‐1160 by kinases such as extracellular signal regulated kinase (ERK),
(Thr‐1150 in mouse) to suppress tyrosine kinase activity of IR cJun N‐terminal kinase (JNK), protein kinase Cζ (PKCζ), and
[37–41]. Thus, intervention of accumulation of triglyceride in p70S6K [47]. The phosphorylation of IRS on Ser‐307 is a
liver results in the reversal of hepatic insulin resistance in ­typical inhibitory signal to suppress insulin signaling as Ser‐307
humans and rodent models with NAFLD. In this regard, adipose locates in the phosphotyrosine‐binding (PTB) domain of IRS [48].
39:  Insulin‐Mediated PI3K and AKT Signaling 491

Thus, increased TNF‐α, saturated free fatty acids, and endoplas­ participate glucose production in liver. Although acetyl‐CoA
mic reticulum (ER) stress in obese individuals activate JNK and generated by β‐oxidation of NEFA in mitochondria does not
inhibitor of nuclear factor κB kinase β (IKKβ) to phosphorylate contribute to the substrates for glucose production, acetyl‐CoA
Ser‐307 of IRS. In addition, ERK activated by insulin also phos­ allosterically activates pyruvate carboxylase to enhance gluco­
phorylates IRS1 on Ser‐612 to attenuate AKT activation [49]. neogenesis [33]. Thus, insulin‐mediated suppression of gluca­
gon secretion and adipose tissue lipolysis is important in the
regulation of hepatic gluconeogenesis. Experimentally, infusion
of insulin in fasted rats decreased the concentration of glycerol
PI3K AND AKT IN GLUCOGENOGENESIS and acetyl‐CoA along with the suppression of hepatic glucose
AND LIPOGENESIS production. However, when infusion of glycerol and acetate was
added to the infusion of insulin, hepatic glucose production was
The secretion of insulin and glucagon are tightly regulated to restored as increased glycerol and acetyl‐CoA enhanced gluco­
modulate gluconeogenesis and lipogenesis in hepatocytes, neogenesis [33–35]. These results suggest that lipolysis is also
and dysregulation of the proximal signaling pathways of these involved in the regulation of gluconeogenesis.
ligands results in the hyperglycemia and hyperlipidemia seen in FOXO1 and transcriptional coactivator peroxisome prolif­
metabolic diseases. In fasting, glucagon and catecholamines erator‐activated receptor‐γ coactivator‐1α (PGC‐1α) enhance
regulate gluconeogenesis through the activation of cAMP‐ the expression of gluconeogenic genes to mediate gluconeo­
dependent PKA. Activated PKA phosphorylates CREB‐Ser133 genesis in liver. To suppress gluconeogenesis in the post­
to induce the expression of gluconeogenic genes such as Pck1 prandial state, insulin‐dependent PI3K and AKT activation
and G6pc and also phosphorylates liver pyruvate kinase (L‐PK) phosphorylates FOXO1 (Thr‐24, Ser‐253, and Ser‐316 for
to suppress glycolysis. Glucagon decreases the level of fruc­ murine FoxO1) to exclude FOXO1 from the nucleus that is
tose‐2,6‐bisphosphate, an allosteric activator of phosphofruc­ then subsequently ubiquitinated and degraded by proteasome,
tokinase and an inhibitor of fructose‐1,6‐bisphosphatase to resulting in the suppression of gluconeogenesis [54, 55].
suppress glycolysis via a phosphorylation of phosphofructoki­ However, the role of insulin in gluconeogenesis through
nase 2 (PFK2) [50]. Glucagon‐induced activation of PKA regu­ FOXO1‐PGC1α is still unclear as liver specific IR and FOXO1
lates gene expression mediated by the CREB‐CBP‐CRTC2 gene double knockout mice show normal glucose homeostasis
complex, and CRTC2 is a key regulator in CREB‐CBP‐CRTC2 although liver specific IR knockout mice have glucose intoler­
complex regulation. Glucagon‐induced PKA activation sup­ ance and insulin resistance [56, 57]. In addition, T2D patients
presses constitutively active serine/threonine kinase SIK2 to and insulin resistant high‐fat diet fed rodent models do not
decrease CRTC2‐Thr171 phosphorylation and also phospho­ show differential expression of gluconeogenic genes including
rylates inositol‐1,4,5‐triphsophate receptors (IP3R) to increase PCK1 and G6PC [58]. In contrast to the high‐fat diet, fructose
Ca2+ influx for CRTC2‐specific phosphatase calcineurin acti­ fed rodents display increased G6pc expression that is the tran­
vation, resulting in the formation of the CREB‐CBP‐CRTC2 scriptional target of both FOXO1 (primarily a gluconeogenic
complex in the nucleus to induce gluconeogenic target gene gene transcriptional activator) and ChREBPβ (primarily a
expression [51, 52]. In contrast, insulin suppresses the secretion lipogenic gene transcriptional regulator) [59]. Since insulin
of glucagon from pancreatic α‐cells, and phosphorylated inhibits G6pc expression through suppression of FOXO1, and
CRTC2 through insulin signaling interacts with 14‐3‐3 protein fructose is a poor inducer of insulin secretion, it is hypothe­
to be sequestered in cytoplasm, thereby suppressing the forma­ sized that one mechanism for enhanced gluconeogenesis by
tion of the CREB‐CBP‐CRTC2 complex [53]. Thus, insulin fructose is due to an imbalance between FOXO1 and ChREBPβ
regulates glucose homeostasis after feeding through the activa­ regulation.
tion of PI3K and AKT to suppress gluconeogenesis and enhance
lipogenesis in hepatocytes.
PI3K and AKT‐mediated modulation
in lipogenesis
PI3K and AKT‐mediated modulation
Glucose from excess dietary carbohydrate undergoes glycoly­
in gluconeogenesis sis in liver and is eventually converted into fatty acids through
Gluconeogenesis in liver contributes approximately half of the DNL and then esterified to triglyceride for very low density
hepatic glucose production in humans during overnight fasting lipoprotein (VLDL) secretion into blood circulation. DNL is
and is the primary mechanism responsible for the increased abnormally increased in NAFLD and closely linked to the
fasting glucose levels in T2D patients. Gluconeogenesis is
­ pathogenesis of T2D. Regulation of DNL is mediated by tran­
­regulated by complex mechanisms: (i) availability of gluconeo­ scriptional regulation of enzymes such as fatty acid synthase
genic substrate such as lactate, alanine and glycerol, (ii) meta­ (FAS) and acetyl‐CoA carboxylase (ACC) for fatty acid syn­
bolic intermediate‐induced allosteric regulation in metabolic thesis and allosteric regulation of ACC. The transcriptional
enzymes, and (iii) the balance of hormones such as insulin, regulation of critical enzymes in DNL is modulated by two
glucagon, and catecholamines. Glycerol and non‐esterified fatty major transcriptional regulators including sterol regulatory ele­
acids (NEFA) generated by lipolysis are involved in the regula­ ment binding protein 1c (SREBP1c) and carbohydrate response
tion of gluconeogenesis. Glycerol, one of major substrates element binding protein (ChREBP) that are activated by
in  gluconeogenesis, is converted to glycerol‐3‐phosphate by enhanced insulin signaling and glucose concentration, respec­
glycerol kinase and then dihydroxyacetone‐3‐phosphate to tively [60, 61]. Thus, insulin‐induced PI3K and AKT activation
492 THE LIVER:  PI3K AND AKT IN LIVER DISEASES: NAFLD AND CANCER

closely regulate the DNL in liver. mTORC1 activated by AKT PI3K and AKT signaling in non‐alcoholic
suppresses Lipin1, a phosphatidic acid phosphatase, to increase fatty liver disease
phosphorylation of nascent SREBP1c in ER, resulting in the
activation of SREBP1c to induce the expression of fatty acid The development of NAFLD is an early step in the pathogenesis
synthesis enzymes such as FAS and ACC in liver [62, 63]. In of T2D, and most obese T2D patients have NAFLD. As lipid
contrast, ChREBP regulation is mediated by glucose imported accumulation induces insulin resistance in liver, weight loss
by insulin independent GLUT2. Thus, glucose metabolites leads to the resolution of NAFLD and normalization of fasting
such as glucose‐6‐phosphate and fructose‐2,6‐bisphosphate glucose level [72]. Thus, hyperinsulinemia associated with
and dephosphorylation of Ser‐196 in ChREBP are proposed to hepatic insulin resistance is a hallmark feature of NAFLD, espe­
regulate ChREBP activity [64, 65]. ACC is also one of the key cially in humans with T2D. Diverse tissue specific IR knockout
regulators modulating DNL in liver as ACC converts acetyl‐ mice demonstrate that muscle specific IR knockout mice or
CoA to malonyl‐CoA to provide monomer for fatty acid syn­ muscle/adipose tissue specific IR knockout mice have normal
thesis in DNL. ACC has two isoforms: ACC1 in adipose tissue, glucose levels. However, liver specific IR knockout mice result
mammary gland, and liver and ACC2 in skeletal and cardiac in hyperglycemia with peripheral insulin resistance and NASH,
muscle. ACC1 activity is regulated by several different levels. suggesting that liver insulin resistance is critical in pathogenesis
First, ACC expression is enhanced by insulin signaling through of NAFLD [73]. Peripheral insulin resistance induces lipolysis
SREBP1c activation along with ChREBP and LXRs [61]. in adipose tissue to increase free fatty acid flux into the liver,
Second, ACC1 exists as a low activity dimer, and allosteric and liver insulin resistance induces DNL through SREBP1c
activators such as citrate and glutamate stimulate polymeriza­ activation, resulting in fat accumulation in liver.
tion of ACC1 to generate high activity polymers to enhance The liver‐specific overexpression in mice of a mutated cata­
DNL [66], whereas malonyl‐CoA and fatty acyl‐CoA inhibits lytic subunit p110α (Pik3ca) found in human HCC results in
ACC1 polymerization as a feedback inhibition mechanism. severe liver steatosis in six months and liver tumors in a year
Third, dephosphorylation and phosphorylation of ACC1 by [74]. In contrast liver specific Pik3ca deficient mice show sup­
insulin and glucagon are also important for the regulation of pressed liver steatosis with a high‐fat diet [75], suggesting that
ACC1 activity although the molecular mechanism is still PI3K is involved in NAFLD. Unlike Pik3ca, Pik3cb (encoding
unclear [67–69]. Furthermore, malonyl CoA, a product of for p110β) deficiency mice show normal hepatic lipid levels,
ACC, inhibits carnitine : palmitoyl‐CoA transferase‐1 (CPT1) indicating that Pik3ca is specifically involved in NAFLD and
to modulate the entry of long‐chain fatty acyl‐CoA into mito­ HCC development. PTEN dephosphorylates PI(3,4,5)P3 to sup­
chondria for β‐oxidation. Thus, ACC suppression is important press PI3K‐AKT signaling pathways, and PTEN deficient mice
to reduce lipid stores in muscle and liver and hence to enhance demonstrate early onset of hepatic microvesicular steatosis due
insulin sensitivity. to increased DNL and free fatty acid uptake that progresses into
NASH, liver fibrosis, and cancer [76]. As Pik3ca deficient mice
show steatosis without NASH and liver fibrosis, PTEN deficient
mice are useful rodent models to recapitulate human NAFLD
PI3K AND AKT IN LIVER DISEASES: pathogenesis. AKT, a downstream PI3K signaling molecule, is
NAFLD AND CANCER also involved in NAFLD development. Adenoviral overexpres­
sion of AKT significantly induces micro and macrovesicular
NAFLD is a common hepatic disorder characterized by fat steatosis and NASH in 12 weeks and then results in HCC [77].
accumulation in the liver in the absence of excessive alcohol High‐fat diet‐induced hepatic fibrosis enhances deposition of
intake. Steatosis, which has benign fat accumulation within the ECM, resulting in interactions between the ECM and AKT
cytoplasm of at least 5% of the hepatocytes in the liver, pro­ through the integrin‐linked protein kinase (ILK) [78]. Thus,
gresses to non‐alcoholic steatohepatitis (NASH), fibrosis, cir­ liver‐specific ILK deletion shows improved HFD‐induced liver
rhosis, and liver cancer. For the diagnosis of NASH, hepatocyte steatosis and insulin resistance.
ballooning and inflammation appear with steatosis. However,
NASH is not necessary for liver fibrosis, and in general NASH
and fibrosis frequently occur together. Liver fibrosis character­
PI3K and AKT signaling in cancer
ized by the deposition of excess extracellular matrix (ECM) Platelet‐derived growth factor (PDGF) stimulates PI3K to gener­
such as collagen by activated hepatic stellate cells progresses to ate PI(3,4)P2 and PI(3,4,5)P3 in smooth muscle, suggesting that
cirrhosis, which demonstrates the loss of hepatic structure, por­ PI3K is important for cellular response to growth factors and
tal hypertension, and the formation of regenerative nodules. tumorigenesis [79], and the PI3K pathway is one of the most fre­
Cirrhosis is irreversible and induces liver organ failure and quently activated signaling pathways in human cancer. Thus,
hepatocellular carcinoma (HCC). NAFLD is closely linked with PI3K mutations are commonly found in a variety of cancers, and
hepatic and adipose tissue insulin resistance, showing that about development of PI3K inhibitors is currently being conducted to
50% reduction in glucose disposal and impaired insulin‐medi­ treat various cancers. Systemic identification of somatic muta­
ated suppression in endogenous gluconeogenesis, and approxi­ tions in cancer genomes uncovers diverse mutations in human
mately 80% of the T2D patients demonstrates NAFLD [70, 71]. cancer to help understand the molecular mechanisms of tumori­
Thus, understanding the molecular mechanisms linking these genesis and develop targeted therapeutics. Class I PI3K, espe­
pathophysiologic symptoms is one of major targets of current cially p110α/p85α heterodimer, activation is closely linked with
biomedical research. tumorigenesis [80, 81]. Most of the reported mutations in p110α
39:  Insulin‐Mediated PI3K and AKT Signaling 493

encoded by the PIK3CA cluster are conserved in the region for [87]. Pan‐PI3K inhibitors such as buparlisib (BKM120) and
the helical (exon 9, E542K and E545K) and kinase domains XL147 suppress all p110 isoforms, whereas isoform‐specific
(exon20, H1047R) of p110α [33]. These mutations in PIK3CA PI3K inhibitors are developed to minimize the side‐effects of
constitutively activate p110α kinase activity without upstream pan‐PI3K inhibitors. p110α‐specific alpelisib and MLN1117
stimuli by growth factors, promoting uncontrolled proliferative and p110β‐specific taselisib (GDC‐0032) are more effective to
signaling through the constitutive activation of AKT, S6K, and tumors with PIK3CA mutation. In contrast, p110α‐specific
4E‐BP in tumorigenesis. PIK3CA mutations in the helical and GSK2636771 is more effective in PTEN‐deficient tumors. AKT
kinase domains also show a distinct pattern in gender and tissue inhibitors such as AZD‐5363, GDC‐0068, and MK‐2206 are in
specificity. In colorectal cancer, PIK3CA mutations are more fre­ clinical trials. MK‐2206, an allosteric pan‐AKT inhibitor, is in
quently found in women, and helical domain mutations (exon 9) clinical trial to treat cancers such as breast and colorectal can­
demonstrate more effects than kinase domain mutations (exon cers. MK‐2206 suppresses phosphorylation of AKT Thr‐308
20). The mutations of PIK3CB encoding for the p110β subunit on and Ser‐473, resulting in the suppression of cell proliferation in
E633K in the helical domain are linked with increased kinase cell lines, which has PIK3CA activating mutations, PTEN inac­
activity and enhanced cell proliferation. Like the p110α H1047R tivation, and amplification of AKT [88]. As MK‐2206 treatment
mutation, the p110β E633K mutation enhances membrane tar­ demonstrates limited antitumor activity in phase II clinical trial,
geting to activate downstream signaling. Increased expression combined treatment with chemotherapeutics and small mole­
of p110γ has been reported in chronic myeloid leukemia, invasive cule inhibitors are in trials. Although rapalogs, the rapamycin
breast carcinoma, and pancreatic ductal adenocarcinoma. analogs to suppress mTOR, are developed as immunosuppres­
Recently, somatic mutations in PIK3R1, which encodes p85α, are sants, they also suppress cell proliferation and angiogenesis to
also found in cancers, and these mutations cluster in the iSH2 treat cancer. Temsirolimus is the first rapalog that is approved
region (Figure 39.1a) releasing p110α to enhance kinase activity, by the Food and Drug Administration (FDA) for the treatment
resulting in the phosphorylation of AKT on S473 [82]. Tumor of renal cell carcinoma as it arrests the cell cycle in the G1 phase
suppressor PTEN suppresses PI3K and AKT by dephosphoryla­ and suppresses angiogenesis by reducing vascular endothelial
tion of PI(3,4,5)P3 and PI(3,4)P2, thus suppressing tumorigenesis. growth factor (VEGF) synthesis.
In addition to the alteration in PI3Ks, mutations of other signaling
molecules in PI3K/AKT/mTORC axis such as PTEN, AKT,
TSC1/,2 and mTOR are also present. PTEN is frequently mutated SUMMARY
in human cancers including HCC (about 5% of HCC), and
decreased PTEN levels are observed in human hepatic steatosis Numerous studies over the past decade have begun to unravel
[83, 84]. Thus, liver‐specific deletion of the PTEN gene results in the complex molecular details and specific signaling pathways
hyperplastic, fatty degeneration of the liver and increased cell that regulate normal liver function with respect to hepatic glu­
proliferation [76]. Thus, the class I PI3K, p100α, appears to be an cose and fatty acid metabolism. These efforts have also begun to
ideal target for drug development, whereas little is known about elucidate the primary basis for dysregulated liver metabolism
the genetic modification in class II and III PI3K in cancer. that occurs during states of insulin resistance and T2D. We now
AKT regulates cell proliferation and survival and is one of the know that elevated fatty acid levels can serve as a significant
most activated downstream effectors of PI3K activation in tumo­ contributor to hepatic glucose production. When coupled with
rigenesis. AKT phosphorylates GSK3β to prevent degradation of liver insulin resistance this combination becomes a strong driver
cyclin D1 and activates the mTORC pathway to enhance the of unrestrained gluconeogenesis and hyperglycemia. On the
translation of cyclin D1 and D3, resulting in cell proliferation. other hand, hyperinsulinemia associated with insulin resistance
AKT also prevents programmed cell death through the phospho­ activates the transcriptional network that drives lipogenic gene
rylation of Bcl‐2‐associated death promoter (BAD) to suppress expression and DNL. When coupled with other dietary factors,
cytochrome c release from mitochondria and caspase activation fructose further enhances DNL through ChREBP that at the
by phosphorylation of pro‐caspase 9. AKT promotes additional same time exacerbates hepatic glucose production through tran­
processes such as angiogenesis by increased nitric oxide produc­ scriptional activation of G6PC. Several challenges remain but
tion and metastasis through the increased secretion of matrix further understanding of the network and interplay between
metalloproteinases and epithelial‐mesenchymal transition liver gluconeogenic and lipogenic signaling should provide new
(EMT) [85]. Thus, AKT is closely linked with tumorigenesis. avenues to develop specific therapeutic approaches for the treat­
Indeed, the AKT activation is observed in human cancers through ment of metabolic liver disease.
the amplification, overexpression, or point mutation of the AKT
kinase genes. AKT1 amplification causes the resistance to cis­
platinum treatment in gastric carcinomas. Somatic mutations of
AKT1 on E17K causes localization of AKT1 at the plasma mem­ REFERENCES
brane to increase phosphorylation on Ser‐473 and Thr‐308,
resulting in leukemia in the mouse model [86]. In HCC, enhanced 1. Vanhaesebroeck, B., Guillermet‐Guibert, J., Graupera, M., and Bilanges, B.
expression of AKT2 protein was detected in about 40% tumors, The emerging mechanisms of isoform‐specific PI3K signalling. Nat Rev Mol
whereas AKT1 expression was similar in all cases. Cell Biol, 2010;11(5):329–41.
2. Wymann, M.P., Bulgarelli‐Leva, G., Zvelebil, M.J. et  al. Wortmannin
As the PI3K/AKT/mTORC signaling axis is a major determi­ ­inactivates phosphoinositide 3‐kinase by covalent modification of Lys‐802,
nant of tumorigenesis, several molecules for suppressing this a  residue involved in the phosphate transfer reaction. Mol Cell Biol,
pathway have been developed and evaluated in clinical trials 1996;16(4):1722–33.
494 THE LIVER:  REFERENCES

3. Backer, J.M. The regulation of class IA PI 3‐kinases by inter‐subunit interac­ 27. Jacinto, E., Facchinetti, V., Liu, D. et al. SIN1/MIP1 maintains rictor‐mTOR
tions. Curr Top Microbiol Immunol, 2010;346:87–114. complex integrity and regulates Akt phosphorylation and substrate specific­
4. Carpenter, C.L., Auger, K.R., Chanudhuri, M. et  al. Phosphoinositide 3‐ ity. Cell, 2006;127(1):125–37.
kinase is activated by phosphopeptides that bind to the SH2 domains of the 28. Cho, H., Mu, J., Kim, J.K., Thorvaldsen, J.L. et al. Insulin resistance and a
85‐kDa subunit. J Biol Chem, 1993;268(13):9478–83. diabetes mellitus‐like syndrome in mice lacking the protein kinase Akt2
5. Myers, M.G., Jr., Backer, J.M., Sun, X.J. et al. IRS‐1 activates phosphati­ (PKB beta). Science, 2001;292(5522):1728–31.
dylinositol 3’‐kinase by associating with src homology 2 domains of p85. 29. Manning, B.D. and Cantley, L.C. AKT/PKB signaling: navigating down­
Proc Natl Acad Sci USA, 1992;89(21):10350–4. stream. Cell, 2007;129(7):1261–74.
6. Bellacosa, A., Testa, J.R., Staal, S.P., and Tsichlis, P.N. A retroviral onco­ 30. Cross, D.A., Alessi, D.R., Cohen, P., Andjelkovich, M., and Hemmings, B.A.
gene, akt, encoding a serine‐threonine kinase containing an SH2‐like region. Inhibition of glycogen synthase kinase‐3 by insulin mediated by protein
Science, 1991;254(5029):274–7. kinase B. Nature, 1995;378(6559):785–9.
7. Coffer, P.J. and Woodgett, J.R. Molecular cloning and characterisation of a 31. Saxton, R.A. and Sabatini, D.M. mTOR signaling in growth, metabolism,
novel putative protein‐serine kinase related to the cAMP‐dependent and pro­ and disease. Cell, 2017;169(2):361–71.
tein kinase C families. Eur J Biochem, 1991;201(2):475–81. 32. Czech, M.P., Tencerova, M., Pedersen, D.J., and Aouadi, M. Insulin
8. Jones, P.F., Jakubowicz, T., Pitossi, F.J., Maurer, F., and Hemmings, B.A. ­signalling mechanisms for triacylglycerol storage. Diabetologia, 2013;56(5):
Molecular cloning and identification of a serine/threonine protein kinase of the 949–64.
second‐messenger subfamily. Proc Natl Acad Sci USA, 1991;88(10):4171–5. 33. Perry, R.J., Camporez, J.P., Kursawe, R. et al. Hepatic acetyl CoA links adi­
9. Staal, S.P. Molecular cloning of the akt oncogene and its human homologues pose tissue inflammation to hepatic insulin resistance and type 2 diabetes.
AKT1 and AKT2: amplification of AKT1 in a primary human gastric adeno­ Cell, 2015;160(4):745–58.
carcinoma. Proc Natl Acad Sci USA, 1987;84(14):5034–7. 34. Rebrin, K., Steil, G.M., Mittelman, S.D., and Bergman, R.N. Causal linkage
10. Klarlund, J.K., Guilherme A., Holik, J.J., Virbasius, J.V., Chawla, A., and between insulin suppression of lipolysis and suppression of liver glucose
Czech, M.P. Signaling by phosphoinositide‐3,4,5‐trisphosphate through pro­ output in dogs. J Clin Invest, 1996;98(3):741–9.
teins containing pleckstrin and Sec7 homology domains. Science, 35. Ramnanan, C.J., Edgerton, D.S., Rivera, N. et al. Molecular characterization
1997;275(5308):1927–30. of insulin‐mediated suppression of hepatic glucose production in vivo.
11. Stephens, L., Anderson, K., Stokoe, D. et al. Protein kinase B kinases that Diabetes, 2010;59(6):1302–11.
mediate phosphatidylinositol 3,4,5‐trisphosphate‐dependent activation of 36. Perry, R.J., Zhang, X.M., Zhang, D. et  al. Leptin reverses diabetes by
protein kinase B. Science, 1998;279(5351):710–4. ­suppression of the hypothalamic‐pituitary‐adrenal axis. Nat Med, 2014;20(7):
12. Keranen, L.M., Dutil, E.M., and Newton, A.C. Protein kinase C is regulated 759–63.
in vivo by three functionally distinct phosphorylations. Curr Biol, 37. Petersen, M.C., Madiraju, A.K., Gassaway, B.M. et  al. Insulin receptor
1995;5(12):1394–403. Thr1160 phosphorylation mediates lipid‐induced hepatic insulin resistance.
13. Calleja, V., Alcor, D., Laguerre, M. et al. Intramolecular and intermolecular J Clin Invest, 2016;126(11):4361–71.
interactions of protein kinase B define its activation in vivo. PLoS Biol, 38. Chin, J.E., Liu, F., and Roth, R.A. Activation of protein kinase C alpha
2007;5(4):e95. inhibits insulin‐stimulated tyrosine phosphorylation of insulin receptor
­
14. Sarbassov, D.D., Guertin, D.A., Ali, S.M., and Sabatini, D.M. ­substrate‐1. Mol Endocrinol, 1994;8(1):51–8.
Phosphorylation and regulation of Akt/PKB by the rictor‐mTOR complex. 39. Badin, P.M., Vila, I.K., Louche, K. et al. High‐fat diet‐mediated lipotoxicity
Science, 2005;307(5712):1098–101. and insulin resistance is related to impaired lipase expression in mouse skel­
15. Pessin, J.E. and Saltiel, A.R. Signaling pathways in insulin action: molecular etal muscle. Endocrinology, 2013;154(4):1444–53.
targets of insulin resistance. J Clin Invest, 2000;106(2):165–9. 40. Griffin, M.E., Marcucci, M.J. et al. Free fatty acid‐induced insulin resistance
16. Taniguchi, C.M., Emanuelli, B., and Kahn, C.R. Critical nodes in signalling is associated with activation of protein kinase C theta and alterations in the
pathways: insights into insulin action. Nat Rev Mol Cell Biol, 2006;7(2): insulin signaling cascade. Diabetes, 1999;48(6):1270–4.
85–96. 41. Jornayvaz, F.R. and Shulman, G.I. Diacylglycerol activation of protein
17. Ullrich, A., Bell, J.R., Chen, E.Y. et al. Human insulin receptor and its rela­ kinase Cepsilon and hepatic insulin resistance. Cell Metab, 2012;15(5):
tionship to the tyrosine kinase family of oncogenes. Nature, 574–84.
1985;313(6005):756–61. 42. Haemmerle, G., Lass, A., Zimmermann, R. et  al. Defective lipolysis and
18. Ebina, Y., Ellis, L., Jarnagin, K. et al. The human insulin receptor cDNA: the altered energy metabolism in mice lacking adipose triglyceride lipase.
structural basis for hormone‐activated transmembrane signalling. Cell, Science, 2006;312(5774):734–7.
1985;40(4):747–58. 43. Chavez, J.A. and Summers, S.A. A ceramide‐centric view of insulin resist­
19. Voliovitch, H., Schindler, D.G., Hadari, Y.R., Taylor, S.I., Accili, D., and ance. Cell Metab, 2012;15(5):585–94.
Zick, Y. Tyrosine phosphorylation of insulin receptor substrate‐1 in vivo 44. Shimomura, I., Matsuda, M., Hammer, R.E., Bashmakov, Y., Brown, M.S.,
depends upon the presence of its pleckstrin homology region. J Biol Chem, and Goldstein, J.L. Decreased IRS‐2 and increased SREBP‐1c lead to mixed
1995;270(30):18083–7. insulin resistance and sensitivity in livers of lipodystrophic and ob/ob mice.
20. Araki, E., Lipes, M.A., Patti, M.E. et al. Alternative pathway of insulin sig­ Mol Cell, 2000;6(1):77–86.
nalling in mice with targeted disruption of the IRS‐1 gene. Nature, 45. Rui, L., Yuan, M., Frantz, D., Shoelson, S., and White, M.F. SOCS‐1 and
1994;372(6502):186–90. SOCS‐3 block insulin signaling by ubiquitin‐mediated degradation of IRS1
21. Withers, D.J., Gutierrez, J.S., Towery, H. et al. Disruption of IRS‐2 causes and IRS2. J Biol Chem, 2002;277(44):42394–8.
type 2 diabetes in mice. Nature, 1998;391(6670):900–4. 46. Kwon, H. and Pessin, J.E. Adipokines mediate inflammation and insulin
22. Zhang, X., Vadas, O., Perisic, O. et al. Structure of lipid kinase p110beta/ resistance. Front Endocrinol, 2013;4:71.
p85beta elucidates an unusual SH2‐domain‐mediated inhibitory mechanism. 47. Boura‐Halfon, S. and Zick, Y. Phosphorylation of IRS proteins, insulin
Mol Cell, 2011;41(5):567–78. action, and insulin resistance. Am J Physiol Endocrinol Metab, 2009;296(4):
23. Sopasakis, V.R., Liu, P., Suzuki, R. et  al. Specific roles of the p110alpha E581–91.
isoform of phosphatidylinsositol 3‐kinase in hepatic insulin signaling and 48. Hirosumi, J., Tuncman, G., Chang, L. et al. A central role for JNK in obesity
metabolic regulation. Cell Metab, 2010;11(3):220–30. and insulin resistance. Nature, 2002;420(6913):333–6.
24. Terauchi, Y., Tsuji, Y., Satoh, S. et al. Increased insulin sensitivity and hypo­ 49. Bard‐Chapeau, E.A., Hevener, A.L., Long, S., Zhang, E.E., Olefsky, J.M.,
glycaemia in mice lacking the p85 alpha subunit of phosphoinositide 3‐ and Feng, G.S. Deletion of Gab1 in the liver leads to enhanced glucose toler­
kinase. Nat Genet, 1999;21(2):230–5. ance and improved hepatic insulin action. Nat Med, 2005;11(5):567–71.
25. Wijesekara, N., Konrad, D., Eweida, M. et al. Muscle‐specific Pten deletion 50. Rider, M.H., Bertrand, L., Vertommen, D., Michels, P.A., Rousseau, G.G.,
protects against insulin resistance and diabetes. Mol Cell Biol, and Hue, L. 6‐phosphofructo‐2‐kinase/fructose‐2,6‐bisphosphatase: head‐
2005;25(3):1135–45. to‐head with a bifunctional enzyme that controls glycolysis. Biochem J,
26. Sleeman, M.W., Wortley, K.E., Lai, K.M. et  al. Absence of the lipid 2004;381(Pt 3):561–79.
phosphatase SHIP2 confers resistance to dietary obesity. Nat Med,
­ 51. Koo, S.H., Flechner, L., Qi, L. et al. The CREB coactivator TORC2 is a key
2005;11(2):199–205. regulator of fasting glucose metabolism. Nature, 2005;437(7062):1109–11.
39:  Insulin‐Mediated PI3K and AKT Signaling 495

52. Wang, Y., Li G., Goode, J. et  al. Inositol‐1,4,5‐trisphosphate receptor 70. Lazo, M. and Clark, J.M. The epidemiology of nonalcoholic fatty liver dis­
­regulates hepatic gluconeogenesis in fasting and diabetes. Nature, 2012; ease: a global perspective. Semin Liver Dis, 2008;28(4):339–50.
485(7396):128–32. 71. Marchesini, G., Brizi, M., Bianchi, G. et al. Nonalcoholic fatty liver disease:
53. Dentin, R., Liu, Y., Koo, S.H. et al. Insulin modulates gluconeogenesis by a feature of the metabolic syndrome. Diabetes, 2001;50(8):1844–50.
inhibition of the coactivator TORC2. Nature, 2007;449(7160):366–9. 72. Henry, R.R., Scheaffer, L., and Olefsky, J.M. Glycemic effects of intensive
54. Brunet, A., Bonni, A., Zigmond, M.J. et al. Akt promotes cell survival by caloric restriction and isocaloric refeeding in noninsulin‐dependent diabetes
phosphorylating and inhibiting a Forkhead transcription factor. Cell, mellitus. J Clin Endocrinol Metab, 1985;61(5):917–25.
1999;96(6):857–68. 73. Biddinger, S.B. and Kahn, C.R. From mice to men: insights into the insulin
55. Tzivion, G., Dobson, M., and Ramakrishnan, G. FoxO transcription resistance syndromes. Annu Rev Physiol, 2006;68:123–58.
­factors;  regulation by AKT and 14‐3‐3 proteins. Biochim Biophys Acta, 74. Kudo, Y., Tanaka, Y., Tateishi, K. et al. Altered composition of fatty acids
2011;1813(11):1938–45. exacerbates hepatotumorigenesis during activation of the phosphatidylinosi­
56. O’Sullivan, I., Zhang, W., Wasserman, D.H. et al. FoxO1 integrates direct tol 3‐kinase pathway. J Hepatol, 2011;55(6):1400–8.
and indirect effects of insulin on hepatic glucose production and glucose 75. Chattopadhyay, M., Selinger, E.S., Ballou, L.M., and Lin, R.Z. Ablation of
utilization. Nat Commun, 2015;6:7079. PI3K p110‐alpha prevents high‐fat diet‐induced liver steatosis. Diabetes,
57. Titchenell, P.M., Chu, Q., Monks, B.R., and Birnbaum, M.J. Hepatic insulin 2011;60(5):1483–92.
signalling is dispensable for suppression of glucose output by insulin in vivo. 76. Horie, Y., Suzuki, A., Kataoka, E. et al. Hepatocyte‐specific Pten deficiency
Nat Commun, 2015;6:7078. results in steatohepatitis and hepatocellular carcinomas. J Clin Invest,
58. Samuel, V.T., Beddow, S.A., Iwasaki, T. et al. Fasting hyperglycemia is not 2004;113(12):1774–83.
associated with increased expression of PEPCK or G6Pc in patients with 77. Calvisi, D.F., Wang, C., Ho, C. et al. Increased lipogenesis, induced by AKT‐
Type 2 diabetes. Proc Natl Acad Sci USA, 2009;106(29):12121–6. mTORC1‐RPS6 signaling, promotes development of human hepatocellular
59. Kim, M.S., Krawczyk, S.A., Doridot, L. et al. ChREBP regulates fructose‐ carcinoma. Gastroenterology, 2011;140(3):1071–83.
induced glucose production independently of insulin signaling. J Clin Invest, 78. Williams, A.S., Trefts, E., Lantier, L. et al. Integrin‐linked kinase is neces­
2016;126(11):4372–86. sary for the development of diet‐induced hepatic insulin resistance. Diabetes,
60. Solinas, G., Boren, J., and Dulloo, A.G. De novo lipogenesis in metabolic 2017;66(2):325–34.
homeostasis: more friend than foe? Mol Metab, 2015;4(5):367–77. 79. Auger, K.R., Serunian, L.A., Soltoff, S.P., Libby, P., and Cantley, L.C.
61. Wang, Y., Viscarra, J., Kim, S.J., and Sul, H.S. Transcriptional regulation of PDGF‐dependent tyrosine phosphorylation stimulates production of novel
hepatic lipogenesis. Nat Rev Mol Cell Biol, 2015;16(11):678–89. polyphosphoinositides in intact cells. Cell, 1989;57(1):167–75.
62. Hegarty, B.D., Bobard, A., Hainault, I., Ferre, P., Bossard, P., and Foufelle, F. 80. Yuan, T.L. and Cantley, L.C. PI3K pathway alterations in cancer: variations
Distinct roles of insulin and liver X receptor in the induction and cleavage on a theme. Oncogene, 2008;27(41):5497–510.
of  sterol regulatory element‐binding protein‐1c. Proc Natl Acad Sci USA, 81. Samuels, Y., Wang, Z., Bardelli, A. et al. High frequency of mutations of the
2005;102(3):791–6. PIK3CA gene in human cancers. Science, 2004;304(5670):554.
63. Peterson, T.R., Sengupta, S.S., Harris, T.E. et al. mTOR complex 1 regulates 82. Jaiswal, B.S., Janakiraman, V., Kljavin, N.M. et  al. Somatic mutations in
lipin 1 localization to control the SREBP pathway. Cell, 2011;146(3): p85alpha promote tumorigenesis through class IA PI3K activation. Cancer
408–20. Cell, 2009;16(6):463–74.
64. Dentin, R., Tomas‐Cobos, L., Foufelle, F. et al. Glucose 6‐phosphate, rather 83. Xu, Z., Hu, J., Cao, H. et al. Loss of Pten synergizes with c‐Met to promote
than xylulose 5‐phosphate, is required for the activation of ChREBP in hepatocellular carcinoma development via mTORC2 pathway. Exp Mol
response to glucose in the liver. J Hepatol, 2012;56(1):199–209. Med, 2018;50(1):e417.
65. Uyeda, K. and Repa, J.J. Carbohydrate response element binding protein, 84. Whittaker, S., Marais, R., and Zhu, A.X. The role of signaling pathways in
ChREBP, a transcription factor coupling hepatic glucose utilization and lipid the development and treatment of hepatocellular carcinoma. Oncogene,
synthesis. Cell Metab, 2006;4(2):107–10. 2010;29(36):4989–5005.
66. Brownsey, R.W., Boone, A.N., Elliott, J.E., Kulpa, J.E., and Lee, W.M. 85. Grille, S.J., Bellacosa, A., Upson, J. et  al. The protein kinase Akt induces
Regulation of acetyl‐CoA carboxylase. Biochem Soc Trans, epithelial mesenchymal transition and promotes enhanced motility and inva­
2006;34(2):223–7. siveness of squamous cell carcinoma lines. Cancer Res, 2003;63(9):2172–8.
67. Peng, I.C., Chen, Z., Sun, W. et al. Glucagon regulates ACC activity in adi­ 86. Carpten, J.D., Faber, A.L., Horn, C. et  al. A transforming mutation in the
pocytes through the CAMKKbeta/AMPK pathway. Am J Physiol Endocrinol pleckstrin homology domain of AKT1 in cancer. Nature, 2007;448(7152):
Metab, 2012;302(12):E1560–8. 439–44.
68. Witters, L.A., Watts, T.D., Daniels, D.L., and Evans, J.L. Insulin stimulates 87. Mayer, I.A. and Arteaga, C.L. The PI3K/AKT pathway as a target for cancer
the dephosphorylation and activation of acetyl‐CoA carboxylase. Proc Natl treatment. Annu Rev Med, 2016;67:11–28.
Acad Sci USA, 1988;85(15):5473–7. 88. Sangai, T., Akcakanat, A., Chen, H. et  al. Biomarkers of response to
69. Zammit, V.A. Role of insulin in hepatic fatty acid partitioning: emerging Akt  inhibitor MK‐2206 in breast cancer. Clin Cancer Res, 2012;18(20):
concepts. Biochem, J., 1996;314(1):1–14. 5816–28.
Ca2+ Signaling in the Liver
40 Mateus T. Guerra1, M. Fatima Leite2, and Michael H. Nathanson1
Departments of Medicine and Cell Biology, Yale University School of Medicine, New Haven, CT, USA
1

Department of Physiology and Biophysics, UFMG, Belo Horizonte, Brazil


2

MECHANISMS OF Ca2+ SIGNALING membrane pool of PIP2 [1], but several lines of evidence now
suggest it instead hydrolyzes nuclear PIP2, which generates
Hormone receptors and initiation InsP3 in the nucleus to increase free Ca2+ within the nucleo-
plasm [3]. The sequence of events linking hormone receptors
of Ca2+ signals and RTKs to Cai2+ signaling is summarized in Figure 40.1.
Hormones, neurotransmitters, and growth factors initiate cyto-
solic Ca2+ (Cai2+) signaling through a variety of mechanisms [1].
Most commonly, these factors trigger intracellular signaling
Inositol 1,4,5‐trisphosphate receptor
cascades by binding to either G protein‐coupled receptors The InsP3R is an InsP3‐gated Ca2+ channel located on the endo-
(GPCRs) or receptor tyrosine kinases (RTKs) on the surface of plasmic reticulum (ER), although there may also be active chan-
liver cells. Several GPCRs in the liver have been particularly nels at the plasma membrane. In hepatocytes, increases in Cai2+
well studied. These include the V1a vasopressin receptor, the α1B are initiated by binding of InsP3 to the InsP3R. Full length
adrenergic receptor, several subtypes of the P2Y class of sequences for three distinct InsP3R genes have been determined
purinergic receptors, and the angiotensin receptors. RTKs for and knockout mice have been generated for each of these three
several types of growth factors are expressed in liver as well, isoforms [4]. These isoforms share considerable sequence
such as insulin, epidermal growth factor (EGF), and hepatocyte homology, but each subtype is expressed and regulated in a dis-
growth factor (HGF). Upon activation, each of these receptors tinct fashion. There also are isoform‐specific differences in tis-
initiates signaling events that lead to an increase in cytosolic sue expression and subcellular distribution, suggesting that the
and/or nuclear Ca2+. The binding of Ca2+ mobilizing hormones isoforms serve distinct roles in Cai2+ signaling. InsP3Rs are
and growth factors to their specific plasma membrane receptors homotetramers consisting of 313, 307, or 304 KDa subunits,
activates phospholipase C (PLC), which is membrane‐associ- corresponding to the type I, II, or III isoforms, respectively, and
ated. α, β, and γ subtypes of PLC are recognized [2]. Many iso- heterotetramers can form as well [5]. The InsP3R has six mem-
forms of these subtypes have been identified; PLCβ1 and PLCβ2 brane spanning domains, oriented so that the N‐terminus of the
are the isoforms activated by G proteins, while PLCγ is acti- protein is in the cytoplasm. The receptor has an uneven bell
vated by RTKs [2]. Activation of PLC by GPCRs hydrolyzes the shape as deduced by cryo‐electron microscopy with the bulkier
pool of phospholipid phosphatidylinositol‐4,5‐bisphosphate N‐terminus directed toward the cytoplasm and the narrower end
(PIP2) within the plasma membrane. The hydrolysis of PIP2 facing the ER lumen.
by  PLC results in the formation of diacylglycerol (DAG) and Deletion analysis studies of the mouse InsP3R1 have revealed
inositol 1,4,5 trisphosphate (InsP3). DAG remains at the plasma three functional regions within the InsP3R: an N‐terminal
membrane to activate protein kinase C (PKC), while InsP3 dif- InsP3‐binding domain, a Ca2+ channel‐forming C‐terminal
fuses into the cytosol to release Ca2+ from intracellular stores via domain, and a regulatory domain flanked by the InsP3‐binding
its interaction with the InsP3 receptor (InsP3R). Activation of domain and the channel region. Several phosphorylation sites
PLC by RTKs was thought to similarly act on the plasma are found along the InsP3R’s amino acid sequence. Moreover,

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
40: Ca2+ SIGNALING IN THE LIVER 497

Growth factor
Hormone

PLC
PLC γ
β Receptor
InsP3 tyrosine kinase
G-
protein
G-protein
coupled receptor
Ca2+

InsP3R

Endoplasmic
reticulum

To nucleus

Ca2+

Figure 40.1  Mechanisms of hormone‐ and growth factor‐induced Ca2+ signaling in hepatocytes. Upon binding to its specific G protein‐coupled
plasma membrane receptor, a hormone induces phospholipase C (PLC)‐beta to hydrolyze phosphatidylinositol 4,5 bisphosphate (PIP2) to form
diacylglycerol (DAG) and inositol 1,4,5‐trisphosphate (InsP3). Alternatively, upon binding to its specific receptor tyrosine kinase (RTK), a growth
hormone induces PLC‐gamma to hydrolyze PIP2 to form DAG and InsP3. InsP3 then binds to its tetrameric receptor (InsP3R) in the ER, which acts
as a Ca2+ channel to allow Ca2+ to enter the cytosol. RTKs such as c‐Met and the insulin receptor may also translocate to the nucleus in hepatocytes
to activate nuclear PLC and increase Ca2+ within the nucleoplasm.

the regulatory domain interacts with several protein partners reversible addition of N‐acetylglucosamine monosaccharide to
responsible for modulating channel activity. The C‐terminus serine/threonine residues. It is speculated that O‐GlcNAcylation
also interacts with protein partners, which influences the subcel- of InsP3R in liver may be particularly important in states of high
lular localization of the receptor [6]. nutrient availability such as NAFLD or diabetes. InsP3R activity
The InsP3 binding domain includes multiple sequences scat- can also be modulated by protein degradation through the protea-
tered throughout the N‐terminal region and key residues respon- some pathway or by selective proteolysis, providing yet another
sible for the interaction between InsP3 and its receptor have been level of regulation of this Ca2+ release channel.
identified by site‐directed mutagenesis and X‐ray crystallography Many cell types express more than one InsP3R isoform.
[7]. Upon InsP3 binding, the receptor undergoes a conformational Hepatocytes normally express only InsP3R1 and InsP3R2.
change that opens the Ca2+ channel, so that Ca2+ in the ER is InsP3R3 is not detected in hepatocytes, but is the predominant
released into the cytosol. Although each of the three InsP3R iso- isoform in bile duct epithelia [10]. Moreover, InsP3R2 is most
forms acts as an InsP3‐gated Ca2+ channel, the isoforms are not concentrated in the apical region of hepatocytes while InsP3R1
uniformly sensitive to InsP3. The relative order of affinity is type is dispersed throughout the cell (Figure 40.2a) [11]. In contrast,
II > type I > type III. The Kd for InsP3R2 is 27 nM, which is twice in bile duct epithelia InsP3R3 is most concentrated in the apical
as great as the affinity of InsP3R1, and ten times the affinity of region, while InsP3R1 and InsP3R2 are dispersed throughout
InsP3R3. InsP3 is absolutely required for Ca2+ release via the the cell (Figure 40.2b) [12]. The apical localization of InsP3R2
InsP3R, but the concentration of Ca2+ in the cytosol modulates the in hepatocytes depends upon lipid rafts in the canalicular mem-
open probability of the Ca2+ channel [8]. This dependence of the brane [13], but the mechanism by which InsP3Rs associate with
InsP3R on the cytosolic Ca2+ concentration is important for lipid rafts is not yet known.
organizing the spatial and temporal pattern of Cai2+ signals.
The activity of InsP3R Ca2+ channels is highly dependent on
post‐translational modifications such as phosphorylation, binding
Ryanodine receptor
of protein cofactors [6], and O‐linked β‐N‐acetylglucosamine The other major class of intracellular Ca2+ release channels is
glycosylation (O‐GlcNAcylation) [9], which operates by the ryanodine receptor (RyR). Like InsP3R, RyR has three
498 THE LIVER:  MECHANISMS OF Ca2+ SIGNALING

(a) (b)

Figure 40.2  InsP3 receptors (InsP3Rs) are concentrated in the apical region of hepatocytes and cholangiocytes. (a) Confocal immunofluorescence
image of rat liver shows the distribution of the InsP3R2 (green) in hepatocytes. Cells also are labeled with the actin stain phalloidin (red), which
outlines individual hepatocytes and labels their apical region most intensely. InsP3R2 colocalizes with periapical actin (arrowheads), and thus is
most concentrated in the apical region. (b) Confocal immunofluorescence image of a bile duct from a biopsy of a normal human liver shows the
distribution of InsP3R3 (green). Cholangiocytes also are labeled to reveal CFTR (red), which resides on their apical membrane. The InsP3R3 is
most concentrated in the apical region, just beneath CFTR. Reproduced from [105] with permission of Elsevier.

family members (RyR1, RyR2, and RyR3). These channels membrane to ions and small molecules. Irreversible formation
contribute to Cai2+ signaling by releasing Ca2+ from the lumen of of the PTP can dissipate the potential gradient across the mito-
the sarcoplasmic/endoplasmic reticulum into the cytosol [1]. chondrial membrane, leading to mitochondrial swelling and
RyRs autocatalytically release Ca2+ via Ca2+‐induced Ca2+ irreversible cell injury, including apoptosis and necrosis. Indeed,
release (CICR). Rat hepatocytes express only a truncated RyR1 liver regeneration after partial hepatectomy is halted due to mas-
isoform that by itself is not able to elicit Ca2+ release but can sive necrotic cell death in mice lacking MICU1 in hepatocytes,
increase the frequency of InsP3‐induced Ca2+ oscillations [14]. which undergo sustained elevations in mitochondrial Ca2+ [17].
These findings suggest there is a novel RyR‐like protein which Mitochondria, like the ER, are densely distributed in hepato-
contributes to Ca2+ signaling in hepatocytes. cytes. There are regions of close proximity between these two
Nicotinic acid adenine dinucleotide phosphate (NAADP) is a organelles (Figure 40.3), and InsP3R1 clustered in these regions
second messenger that mediates release of Ca2+ from intracel- are largely responsible for mitochondrial Ca2+ signaling in
lular acidic compartments such as lysosomes and secretory hepatocytes [18].
granules. The role of NAADP in hepatocyte Cai2+ signaling is The discovery of interactions between apoptotic proteins,
not clear, although there is in vitro evidence demonstrating InsP3R, and mitochondrial proteins helped uncover an essential
NAADP‐dependent Ca2+ release from reconstituted hepatocyte interplay between ER and mitochondrial Ca2+ signaling in the
lysosomes and microsomes [15]. process of programmed cell death. The current view is that anti‐
apoptotic proteins belonging to the Bcl‐2 family such as Bcl‐XL
and Bcl‐2 induce ER Ca2+ leak through direct interaction with
Mitochondria InsP3R1. This Ca2+ leak then reduces the ER Ca2+ available to be
Mitochondria are best known for the metabolic and respiratory released into the cytosol and consequently, decreases the amount
role they play in cells. However, mitochondria strongly influ- of Ca2+ taken up by surrounding mitochondria. Because mito-
ence Cai2+ signaling as well, by taking up Ca2+ from and releas- chondria are less likely to become overloaded with Ca2+, PTP
ing it back into the cytosol. Mitochondria have their own Ca2+ formation is hampered and there is diminished cellular sensitiv-
transport machinery, involving Ca2+ influx through a uniporter, ity to apoptotic stimuli [19]. Accordingly, buffering of mito-
and Ca2+ efflux via both a Na+ exchanger and a H+ exchanger chondrial matrix Ca2+ prevents apoptotic death of hepatocytes
[16]. The uniporter is driven by the potential gradient across the and accelerates liver regeneration after partial hepatectomy
mitochondrial membrane, while the Ca2+ efflux mechanisms [20]. Mcl‐1 is another anti‐apoptotic protein that acts in part
are  active transport systems. The CCDC109A gene encodes by diminishing mitochondrial Ca2+ signals. Unlike other Bcl‐2
the  mitochondrial calcium uniporter (MCU) protein, and Ca2+ family members, Mcl‐1 is expressed in mitochondria and
influx in mitochondria occurs via a macromolecular complex on ­inhibits mitochondrial Ca2+ signals directly [21]. Mcl‐1 is the
the inner mitochondrial matrix composed of pore‐forming subu- principal anti‐apoptotic protein in cholangiocytes, and its over-
nits (MCU, MCUb, and essential MCU regulator [EMRE]), expression may promote the development of cholangiocarci-
plus regulatory proteins MICU1/2 that modulate the activity of noma. In contrast, overexpression of the pro‐apoptotic proteins
the uniporter according to the concentration of cytosolic Ca2+ Bax and Bak lead to ER Ca2+ overload and greater susceptibility
in  the vicinity of the uniporter [16]. Pathological Ca2+ efflux to Ca2+‐mediated apoptosis.
from mitochondria also can occur through the permeability Mitochondria can sequester significant amounts of Ca2+ from
transition pore (PTP). Formation of this pore results in a sudden, the cytosol and mitochondrial Ca2+ can closely parallel the
marked increase in the permeability of the mitochondrial inner ­cytosolic Ca2+ increase induced by receptor activation. This
40: Ca2+ SIGNALING IN THE LIVER 499

(a) (b)

Figure 40.3  Mitochondria and endoplasmic reticulum form regions that are in close proximity in hepatocytes. (a) Histologically normal human liver
biopsy specimen stained with markers for an ER protein (PDI, green) and a mitochondrial protein (Tom‐22, red) show the close association between
these two organelles in hepatocytes. Regions in white represent areas of colocalization, scale bar = 20 μM. (b) EM micrograph of a mouse hepatocyte
showing regions of close (less than 40 nm) contact between the ER (green) and mitochondria (magenta). Mitochondrial Ca2+ signals are derived from
Ca2+ that is released from InsP3Rs that cluster in these regions. Reproduced from [18] with permission of John Wiley and Sons.

functional effect is consistent with electron microscopy data in turn to act as a second messenger that regulates multiple cell
that show a close proximity between mitochondria and ER and functions simultaneously.
a dynamic physical interaction between these two organelles Ca2+ signaling was initially studied with fluorescent dyes and
[22]. There is also functional evidence for microdomains where the bioluminescent protein aequorin, but genetically encoded
mitochondria and InsP3Rs are in close apposition, so that mito- fluorescent Ca2+ indicators (GECI) are now more widely used.
chondria take up a significant fraction of the Ca2+ released These GECI are based on modified versions of the circularly
by the InsP3R [23]. In hepatocytes, InsP3R1 is the isoform in permuted (cp) green fluorescent protein (GFP) and other fluo-
proximity to mitochondria and is responsible for mitochondrial rescent proteins such as blue fluorescent protein (BFP) and
Ca2+ signals [18]. mApple red fluorescent protein (RFP) [26]. These cp proteins
Ca2+ uptake by mitochondria affects multiple factors in cell are fused with the Ca2+ binding protein calmodulin (CaM) and a
metabolism, including the mitochondrial proton motive force, calmodulin (CaM)–binding region of chicken myosin light
electron transport, and the activity of dehydrogenases associ- chain kinase (M13). Ca2+ binding induces a conformational
ated with the TCA cycle, adenine nucleotide translocase, the change that results in increased fluorescence. Additional refine-
F1‐ATPase, and PDH phosphorylase. Slow or small Cai2+ eleva- ments have yielded ratiometric GECI and sensors with different
tions are not transmitted effectively into mitochondria, and affinities for Ca2+. Other GECI have been developed that are
are less able to activate mitochondrial metabolism. In contrast, based on the principle of fluorescence resonance energy transfer
Cai2+ oscillations trigger mitochondrial Ca2+ oscillations and (FRET). These indicators provide the advantage of better signal
sustained NAD(P)H formation [24]. Thus, the frequency rather to noise ratio and photostability as compared to the aequorin
than the amplitude of Cai2+ oscillations regulates mitochondrial reporters and yet retain the ability to be expressed in diverse tis-
metabolism [24]. sues and subcellular locations.

Ca2+ signaling patterns in hepatocytes


ORGANIZATION OF Ca SIGNALS 2+
Peptide hormones such as vasopressin or angiotensin generally
induce biphasic increases in Cai2+ in populations of isolated
Detection of Ca signals in hepatocytes
2+
hepatocytes (Figure 40.4). These increases typically consist of
Our understanding of the complexity of Cai2+ signals has evolved two components. The first component is the rapid peak, then fall
dramatically, largely due to two technical advances: (i) develop- in Cai2+ which takes place over a period of seconds. This is due
ment of Ca2+‐sensitive fluorescent dyes and proteins has permit- to release of Ca2+ from InsP3‐sensitive stores and occurs even in
ted Cai2+ to be monitored continuously in live cells, and (ii) Ca2+‐free medium. The second component is the sustained pla-
improvements in fluorescence imaging techniques allows Cai2+ teau in Cai2+, which follows the rapid peak. This occurs only in
to be detected not only in cell populations but in single cells and the presence of extracellular Ca2+ and is due mostly to influx
in distinct subcellular regions of individual cells [25]. As obser- of Ca2+ to replenish depleted intracellular stores. Although this
vations move from cell populations to single cells to subcellular population response is highly reproducible, Ca2+ signaling pat-
regions, the complexity of Cai2+ signaling patterns increases and terns vary markedly among single hepatocytes. Different stimuli
the time scale of signaling events decreases (Figure 40.4). It is evoke distinct responses, and additional variation occurs among
now appreciated that Cai2+ signaling is regulated at the subcel- hepatocytes stimulated under identical conditions [27]. The
lular level, and that this level of regulation is necessary for Cai2+ range of signaling patterns seen among single hepatocytes
500 THE LIVER:  ORGANIZATION OF Ca2+ SIGNALS

(a) phenylephrine‐induced oscillations as well. Differences in the


700 frequency of Cai2+ oscillations regulate gene transcription
600 in  some cell systems [29], although this has not yet been
Cytosolic Ca2+ (nM)

500 ­demonstrated in hepatocytes.


It is currently thought that Cai2+ oscillations do not depend
400
upon InsP3 oscillations. Instead, oscillations are thought to
300 result from the bell‐shaped dependence of the open probability
200 of the InsP3R on Cai2+. Extracellular Ca2+ contributes to Cai2+
100 oscillations in hepatocytes, because Cai2+ oscillations gradually
dissipate in Ca2+‐free medium. Extracellular Ca2+ thus serves to
0
−50 0 50 100 150 200 250 maintain internal Ca2+ stores, which are the primary source of
Time (sec) Ca2+ for oscillations in hepatocytes. The mechanism by which
(b) Ca2+ release from the ER triggers Ca2+ influx through the plasma
250
200
membrane has been described in many cell types including
Fluorescence intensity

150 hepatocytes [30]. The process of store‐operated Ca2+ entry


200 (SOCE) involves the interaction of the integral ER membrane
100
(pixel value)

50
0 100 200 300
protein stromal interacting molecule 1 (STIM1) and its binding
150 Time (sec) partner on the plasma membrane, the ORAI calcium release‐
activated calcium modulator (ORAI) channels [31]. STIM1
100 works as a Ca2+ sensor in the ER; once ER Ca2+ is reduced due
to InsP3‐dependent activation or pharmacological inhibition of
50 the sarco/endoplasmic reticulum Ca2+ ATPase (SERCA) pump,
0 10 20 30 40 STIM1 forms homodimers that interact with and gate ORAI
Relative time (sec) channels on the plasma membrane, leading to influx of Ca2+ into
(c)
250 the cytosol. This Ca2+ influx ultimately assists in replenishing
ER Ca2+ stores.
Fluorescence intensity

200
Increases in Cai2+ in hepatocytes typically begin near the
apical membrane, where InsP3R2 is most concentrated [32].
(pixel value)

Both vasopressin‐ and phenylephrine‐induced Cai2+ waves


150
began there, then spread in a non‐diminishing fashion across the
cell. Immunofluorescence studies in rat liver, isolated rat hepat-
100
ocyte couplets, and hepatocytes in collagen sandwich culture
established that InsP3R2 is concentrated subapically [33, 34]
50 and that this localization depends on intact lipid rafts (special-
0.0 0.5 1.0 1.5
ized membrane patches rich in cholesterol) [13] (Figure 40.2).
Relative time (sec)
Examination of Cai2+ signaling in polarized preparations of
Figure 40.4  Different views of cytosolic Ca2+ signaling in hepato- isolated hepatocytes has shown that this is the region where
cytes. In each case, isolated rat hepatocytes were stimulated with the Cai2+ waves originate, and that this depends on local clustering
α1B‐adrenergic agonist phenylephrine. (a) In a population of hepato- of InsP3R2 [13, 32].
cytes, a single transient peak is observed, followed by a sustained eleva-
tion. (b) In a single hepatocyte, a series of repetitive peaks (oscillations)
Ca2+ signals in cholangiocytes begin in the apical region, just
are observed. (c) In different regions of the same hepatocyte, the as they do in hepatocytes, even though InsP3R3 rather than
increase in Ca2+ occurs at different times. This represents a Ca2+ wave. InsP3R2 is concentrated near their apical surface, where the
Notice that these Ca2+ signaling events occur over progressively Ca2+ waves originate [12].
shorter  time intervals as the level of focus moves from populations
to  single cells to subcellular regions. Reproduced from [25] with
­permission of Elsevier. Spread of Ca2+ signals among hepatocytes
Cai2+ signals occur asynchronously among isolated hepatocytes.
includes single transient or sustained Cai2+ increases and repeti- For example, the lag time between stimulation with vasopressin
tive Cai2+ spikes (i.e. oscillations). For example, lower concen- and initiation of Cai2+ signaling varies among isolated hepato-
trations of vasopressin induce Cai2+ oscillations, while higher cytes by up to several seconds, while the frequency of Ca2+
concentrations induce sustained increases in Cai2+ [28], while oscillations can vary by up to 50% among isolated hepatocytes
stimulation of hepatocytes with phenylephrine typically evokes stimulated with phenylephrine [27]. However, hepatocytes that
Cai2+ oscillations, but the oscillation frequency is dose‐dependent communicate via gap junctions coordinate their Cai2+ signals.
[28]. The duration of individual Cai2+ spikes also depends upon For example, stimulation of isolated hepatocyte couplets with
the agonist. For example, Cai2+ spikes induced by phenylephrine vasopressin induces a single Cai2+ wave that crosses both of the
are short (around seven seconds) compared to the duration of cells, while stimulation with phenylephrine induces Cai2+ oscil-
spikes induced by vasopressin (around ten seconds) or angio- lations that are synchronized in the two cells [27]. In the isolated
tensin (around fifteen seconds). The frequency of vasopressin‐ perfused rat liver, Cai2+ signaling displays an even higher level
induced Cai2+ oscillations tends to be greater than that of of organization, because vasopressin induces Cai2+ waves that
40: Ca2+ SIGNALING IN THE LIVER 501

(a)
0 sec. 4 sec. 30 sec.

(b)

1.5

[Ca2+]i (F/F0)
1

0 100 200
Time (s)

Figure 40.5  Organization of Ca2+ signals in the intact liver. (a) Serial confocal images of a vasopressin‐induced Ca2+ wave in the isolated perfused
rat liver demonstrates that the increase in Ca2+ begins in the pericentral region, then spreads as a wave toward the periportal regions. Images were
obtained at baseline and after 4 and 30 seconds of stimulation with vasopressin (20 nM). Reproduced from [35] with permission of the American
Physiological Society. (b) Ca2+ signals in individual hepatocytes within an isolated perfused rat liver stimulated with vasopressin. Tracing shows
Ca2+ oscillations detected in three hepatocytes sequentially arranged along the hepatic lobule. The distance between the first (red tracing) and third
(green tracing) cell is 40 μm. The phase difference between the cells is due to the time needed for a Ca2+ wave to spread across the hepatic lobule.
Reproduced from [106] with permission of Elsevier.

cross the entire lobule [35] (Figure 40.5). Cai2+ waves cross indi- signals among adjacent hepatocytes. This also is important for
vidual hepatocytes at the same speed, regardless of whether the associated physiological responses. For example, bile secretion
hepatocytes are isolated or within the liver. Vasopressin‐induced in isolated perfused rat liver is reduced by pharmacological
Cai2+ waves cross the hepatic lobule in a pericentral‐to‐peripor- inhibition of connexin function [42]. Another physiological pro-
tal direction [36], presumably directed by the V1a vasopressin cess that requires gap junction communication among hepato-
receptor gradient present from the pericentral to periportal cytes is glucose output [43] and this is impaired in Cx32
region [36]. In contrast, ATP induces Cai2+ signals in a random knockout mice stimulated with either glucagon or norepineph-
fashion across the hepatic lobule, consistent with the lack of a rine [44]. Cx32 function also is involved in drug‐induced liver
P2Y receptor gradient across the lobule [37]. Thus, sophisti- injury, because mice defective in Cx32 are partially protected
cated patterns of Cai2+ signaling are induced in the intact liver, from acetaminophen‐induced hepatocyte cell death and liver
and these patterns are agonist‐specific. This may permit differ- failure [45].
ent Ca2+ agonists to have distinct effects in liver even though Organization of cell‐to‐cell Cai2+ waves depends on other fac-
Cai2+ signals induced by these agonists appear similar in isolated tors in addition to gap junctions. For example, increases in
hepatocytes. InsP3 are required in each cell across which a Ca2+ wave spreads
The basis for organization of Cai2+ signals in liver has been [46]. Moreover, neither InsP3 nor Ca2+ alone is sufficient to sup-
studied in multicellular systems of hepatocytes. Studies in iso- port the spread of a Cai2+ wave across a hepatocyte [47]. The
lated rat hepatocyte couplets demonstrate that hepatocytes presence of agonist binding to its specific receptor at the surface
communicate via gap junctions, and that both Ca2+ and InsP3 of the cell also is required to support the spread of Cai2+ waves.
can cross these gap junctions [38]. Hormone‐induced Cai2+ sign- This was demonstrated in experiments in which one or both
aling is highly coordinated in such couplets, and this coordi- cells of a hepatocyte couplet were microperfused with norepi-
nation depends upon gap junction conductance as well [27]. nephrine. Stimulation of individual cells evoked Cai2+ oscilla-
Hepatocytes express two gap junction isoforms, connexin 32 tions only in the perfused cell, and perfusion of an entire couplet
(Cx32) and connexin 26 (Cx26) [39]. Expression of both iso- was necessary to evoke Cai2+ oscillations in both cells [47].
forms is dramatically reduced after bile duct ligation, and coor- Thus, the presence of hormone at each cell ensures that suffi-
dination of Cai2+ signals is impaired under this condition as well cient levels of intracellular messengers are generated in order to
[39]. Furthermore, cell‐to‐cell spread of InsP3 and Cai2+ waves reach a level of excitability necessary for supporting the propa-
is markedly impaired in hepatocytes isolated from Cx32 knock- gation of a Cai2+ wave.
out mice [40]. Moreover, the expression of Cx32 or Cx43 in a Other studies have focused on the mechanism by which inter-
liver cell line, where intercellular Ca2+ signals are not normally cellular Cai2+ waves become oriented within the hepatic acinus.
observed, causes propagation of Ca2+ from cell to cell [41]. Vasopressin‐induced waves begin in the region of the central
Thus, gap junctions play an essential role in organizing Cai2+ venule [36], while ATP‐induced Cai2+ waves begin in a
502 THE LIVER:  ORGANIZATION OF Ca2+ SIGNALS

seemingly random pattern across the hepatic acinus [35]. Evidence suggests that InsP3 releases Ca2+ directly from the
Studies in isolated hepatocytes similarly show that pericentral nuclear envelope into the nucleus. InsP3 releases 45Ca2+ into iso-
hepatocytes are more sensitive to vasopressin but not to ATP lated hepatocyte nuclei, and InsP3 increases free nuclear Ca2+ even
[37]. Studies in isolated hepatocyte couplets and triplets stimu- if the nucleus is surrounded by a Ca2+ chelator. In addition, extra-
lated with vasopressin or norepinephrine show that one of the cellular ATP preferentially activates nuclear Ca2+ release in HepG2
cells generally has increased sensitivity to a particular hormone cells, via an InsP3R‐dependent mechanism [61]. The nuclear
and acts as a pacemaker to drive Cai2+ oscillations in neighbor- envelope possesses the machinery necessary to produce InsP3,
ing cells. This increased sensitivity is likely due to increased including PIP2 and PLC [62], and this machinery may be activated
expression of hormone receptor rather than differences in down- selectively through RTK pathways. In one study, IGF‐1 and integ-
stream signaling components such as G proteins or InsP3R [48]. rins caused PIP2 breakdown in the nucleus but not at the plasma
Cells with increased expression of hormone receptor produce membrane, while activation of G protein‐linked receptors caused
higher concentrations of InsP3, so they respond sooner than breakdown of PIP2 in the cytosol but not the nucleus. Similarly,
other cells stimulated with the same concentration of hormone. activation of the HGF receptor c‐Met in a liver cell line caused
As a result, cells with the greatest level of hormone receptor PIP2 breakdown in the nucleus resulting in nuclear Ca2+ signals.
expression act as pacemakers, and different cells can act as the Moreover, activation of this highly localized cascade was depend-
pacemaker for different hormones. Thus, the pattern of Cai2+ ent on translocation of the activated receptor to the nucleus [3].
waves and oscillations in the intact liver depends upon multiple Nuclear Ca2+ signaling also occurs by spread of Cai2+ signals
factors, which include: (i) the establishment of pacemaker cells into the nucleus. The nuclear envelope contains pores that are per-
by virtue of increased expression of hormone receptors, (ii) meable to molecules up to 60 KDa in size. In the absence of a
simultaneous stimulation of both pacemaker and non‐pace- gating mechanism, a pore this size would allow rapid equilibra-
maker cells, and (iii) communication of second messengers tion of Ca2+ between the nucleus and cytosol. Under certain cir-
among these cells via gap junctions [49]. cumstances, free diffusion of Ca2+ through the nuclear pore
The liver also possesses paracrine mechanisms for generating indeed occurs. However, a nuclear‐cytosolic Ca2+ gradient has
and regulating Cai2+ signals. Hepatocytes and bile duct cells been demonstrated in a number of cell types [63], suggesting that
each secrete ATP [50, 51], and both cell types express P2Y ATP the permeability of nuclear pores can be regulated. Moreover,
receptors [52, 53]. Because P2Y receptors are GPCRs that link electrophysiological studies suggest that Ca2+ permeability
to InsP3‐mediated Cai2+ signaling, secretion of ATP by hepato- through nuclear pores is restricted. Atomic force microscopy
cytes stimulates Cai2+ signaling in neighboring hepatocytes and studies similarly suggest that nuclear pore permeability is regu-
bile duct cells [50]. This paracrine signaling mechanism thus lated, and that depletion of Ca2+ within the nuclear envelope
permits increases in Cai2+ to spread among neighboring cells closes the pores. Other work using fluorescent dyes or aequorin
independent of communication via gap junctions. Hepatocytes also demonstrates that depletion of Ca2+ attenuates the permeabil-
and bile duct cells express P2X receptors as well [54], which are ity of the pores to intermediate‐sized molecules that lack a nuclear
plasma membrane ATP‐gated Ca2+ channels, but the physiologi- localization sequence. Studies monitoring diffusion of photoacti-
cal role of these receptors in liver is less clear. Cai2+ signaling in vatable GFP across the nuclear envelope suggest that permeabil-
liver also can be modified rather than initiated by paracrine ity of the nuclear pore may be regulated by cytosolic Ca2+ rather
pathways. For example, bradykinin does not mobilize Cai2+ in than by Ca2+ stores within the nuclear envelope [64]. EF‐hand
isolated hepatocytes, yet in the intact liver it modifies the propa- Ca2+ binding motifs are present in proteins of the nuclear pore, so
gation of Cai2+ waves induced by vasopressin [55], likely via it is possible that these function as Ca2+‐gating sensors for the
nitric oxide (NO) release from endothelial cells, which diffuses pore. There is additional evidence that nuclear Ca2+ sometimes
to hepatocytes where it stimulates generation of cGMP. passively follows Cai2+ [65]. For example, stimulation of hepato-
cytes with vasopressin results in a Ca2+ wave that appears to
spread from the cytosol into the nucleus. Moreover, a mathemati-
Nuclear Ca2+ signaling cal analysis of Ca2+ waves in hepatocytes stimulated with vaso-
The nucleus is separated from the cytosol by the nuclear enve- pressin suggests that nuclear Ca2+ signals can be described simply
lope, which is a specialized region of the ER. Like the ER, the by diffusion of Ca2+ inward from the nuclear envelope.
nuclear envelope is able to store and release Ca2+. The nuclear Nuclear Ca2+ also can contribute to Ca2+ signals in the cyto-
envelope accumulates Ca2+ via a Ca2+‐ATPase pump [56], and sol. For example, localized increases of Ca2+ in the cytosol (Ca2+
releases Ca2+ via channels that are sensitive to InsP3, cADPR, and puffs) can spread across the cell by diffusing across the nucleus.
NAADP [57, 58]. These Ca2+ storage pumps and release channels Ca2+ puffs are highly transient and localized Cai2+ signals that
have distinct distributions within the nuclear envelope. The Ca2+‐ result from the coordinated opening of small clusters of InsP3Rs.
ATPase pump is located only in the outer leaflets of the envelope, Puffs can be triggered by a subthreshold concentration of ago-
while the InsP3R is located only in the inner membrane, and nist and the resulting Cai2+ signal rapidly dissipates by diffusion
cADPR‐sensitive channels are present on both sides of the nuclear in the cytosol and sequestration of Ca2+ into intracellular stores.
envelope. Both InsP3Rs and RyRs are localized along invagina- However, the range of diffusion of Ca2+ in the nucleus can be
tions of the nuclear envelope, denoted the nucleoplasmic reticu- much greater than in cytosol, so Ca2+ puffs generated near the
lum, which work as a regulatory Ca2+ domain within the nucleus nuclear envelope can spread into and across the nucleus in order
[59]. The importance of nuclear Ca2+ is also highlighted by find- to spread to other, more distant regions of the cytosol [65].
ings demonstrating that different transcription factors are directly Thus, the nucleus may function as a tunnel that helps distribute
or indirectly dependent on Ca2+ in the nucleus [60]. Ca2+ to the cytosol.
40: Ca2+ SIGNALING IN THE LIVER 503

Ca2+ signaling in bile duct cells mobilizes glucose mainly from the periportal zone, while nor-
epinephrine and vasopressin preferentially release glucose from
Cai2+ signaling has been examined to a lesser extent in bile duct pericentral hepatocytes. This may in part reflect the fact that
epithelia than in hepatocytes. ATP and UTP both increase Cai2+ pericentral hepatocytes are more sensitive than periportal hepat-
in the Mz‐ChA‐1 cholangiocarcinoma cell line, a model for bile ocytes to vasopressin and norepinephrine [37]. When hepato-
duct epithelium. ATP and UTP also increase Cai2+ in primary cul- cytes are not uniformly sensitive to a particular hormone, then
tures of rat bile duct epithelia, and acetylcholine (ACh) increases intercellular communication via gap junctions enhances glucose
Cai2+ in these cells via M3 muscarinic receptors [52]. As in release. For example, glucose release is impaired in perfused
hepatocytes and other epithelia, the range of patterns of agonist‐ livers from Cx32‐deficient mice upon stimulation with either
induced Cai2+ signals include sustained and transient Cai2+ norepinephrine or glucagon [44]. Similarly, vasopressin‐ or
increases and Cai2+ oscillations. Cai2+ spikes induced by ACh are glucagon‐induced glucose release is impaired in perfused rat
longer in duration and lower in frequency than those induced by livers treated with the gap junction blocker 18α‐glycyrrhetinic
ATP [52]. Cai2+ signaling is mediated by InsP3 in bile duct cells, acid (αGA) [42]. However, glucose release is not altered in
because increases in Cai2+ are blocked by InsP3R antagonists. As αGA‐treated livers stimulated with dibutyryl cAMP or 2,5‐
discussed above, all three InsP3R isoforms are expressed in these di(tert‐butyl)‐1,4‐benzohydroquinone (tBuBHQ), both of which
cells [12]; Ca2+ waves begin in their apical region, where InsP3R3 stimulate glucose release in a receptor‐independent fashion
is concentrated, and then spread basolaterally via the types I and [42]. Hormone‐induced glucose release also is impaired if gap
II InsP3Rs. Bile duct cells are coupled via connexin 43 (Cx43) junctions are blocked in isolated rat hepatocytes, or if hepato-
gap junctions, and expression of Cx43 synchronizes their Ca2+ cytes are dispersed. Therefore, hepatocytes may contribute dif-
oscillations. Moreover, Cx43 permeability is under hormonal ferently to glucose metabolism across the hepatic lobule,
control, and activation of either protein kinase A or C decreases although there is some integration of metabolic activity via gap
permeability and impairs intercellular communication [66]. junctions. This integration of metabolic function is particularly
important in times of stress, because fasting induces hypoglyce-
mia in Cx32‐deficient but not wild‐type mice, and because
endotoxin‐induced hypoglycemia is exacerbated in the knock-
EFFECTS OF Ca2+ SIGNALS IN HEALTH out mice as well [40]. Similarly, liver glucose production upon
AND DISEASE stimulation with glucagon and norepinephrine is reduced in
Cx32 knockout mice [44]. Glucose output by the liver is also
Ca2+ regulates a wide range of functions in liver. The following regulated by Ca2+‐calmodulin kinase II gamma (CamKIIγ), a
sections are meant to provide illustrative examples of the differ- downstream effector of intracellular Ca2+ signaling [71]. In a
ent ways in which Ca2+ regulates normal and abnormal liver physiological context, this kinase is activated by InsP3R‐
function, rather than to provide an exhaustive list of Ca2+‐medi- dependent Ca2+ release during fasting and promotes transloca-
ated functions. tion of the transcription factor FoxO1 to the nucleus of
hepatocytes. This translocation in turn controls a transcriptional
program that potentiates glycogenolysis and gluconeogenesis
Energy metabolism
and leads to increased glucose output by the liver. This mecha-
Storage and release of glucose was among the first functions of nism is potentiated in experimental obesity and contributes to
the liver shown to be regulated by Cai2+. Synthesis of glycogen the high serum levels of glucose found in obese mice.
is regulated by glycogen synthase, while phosphorylase is the Accordingly, genetic deletion of CamKIIγ or deletion of FoxO1
rate‐limiting enzyme for glycogenolysis. Both enzymes are in mice are associated with lower blood concentrations of glu-
regulated by phosphorylation and dephosphorylation, and an cose. A second Ca2+‐dependent transcription program that mod-
increase in Cai2+ is one of the most important signals for regulat- ulates glucose production by the liver is operated by the CREB
ing these events [67]. For example, hormones such as vasopres- regulated transcription coactivator 2 (CRTC2) [72]. Here,
sin and angiotensin increase InsP3 in hepatocytes, which glucagon activates cAMP formation, PKA activation, and phos-
mobilize Ca2+, leading to phosphorylation and activation of gly- phorylation of InsP3R in hepatocytes. This phosphorylation
cogen phosphorylase, and then glycogenolysis. Similarly, extra- renders the receptor more susceptible to activation by InsP3 and
cellular nucleotides activate glycogen phosphorylase and thus thus Ca2+ is more readily released in the cytosol. Free Ca2+ can
stimulate glycogenolysis by binding to P2Y nucleotide recep- then bind to calcineurin which dephosphorylates CRTC2 allow-
tors [68]. Glucagon and beta‐adrenergic agonists also stimulate ing its translocation to the nucleus. The final result is the activa-
glycogen phosphorylase activity in liver, but through an alterna- tion of genes that promote glucose production. Termination of
tive, cAMP‐dependent pathway. Ca2+‐mobilizing bile acids such this mechanism is mainly by insulin‐dependent activation of
as UDCA, TLCA, and LCA activate phosphorylase to the same Akt. In diabetes, insulin resistance thus ensures prolonged acti-
extent as hormones such as vasopressin. These bile acids acti- vation of the CRTC2 and sustained glucose output by the liver.
vate phosphorylase through a Ca2+‐dependent but InsP3‐inde- CREB also increases transcription and thus expression of
pendent mechanism, consistent with the observation that they InsP3R2 [73]. Hepatic InsP3R2 expression also increases dur-
increase Cai2+ in an InsP3‐independent fashion [69]. ing fasting, and it has been postulated that this occurs via gluca-
Gluconeogenic enzymes are preferentially located in the per- gon‐ and cAMP‐mediated activation of CREB [73].
iportal region [70], although other factors also may be involved Ca2+ signaling also plays a role in lipid metabolism and the
in regional differences in glycogenolytic capacity. ATP pathogenesis of non‐alcoholic fatty liver disease (NAFLD).
504 THE LIVER:  EFFECTS OF Ca2+ SIGNALS IN HEALTH AND DISEASE

Relevant studies mainly describe changes in expression of Ca2+ with the plasma membrane [33]. Bile‐salt dependent flow,
handling proteins, such as intracellular channels and pumps, and which relies on the activity of the bile salt export pump (Bsep),
their effect on lipid metabolism. Studies in obese mice showed is similarly decreased in rat hepatocytes with reduced expres-
that expression of Serca2b is specifically downregulated in the sion of InsP3R2 or in cells treated with an intracellular Ca2+
liver [74]. The physiological function of this sarco(endo)plasmic buffering agent [34]. Moreover, disruption of lipid rafts in the
reticulum Ca2+‐ATPase is to maintain high levels of Ca2+ within apical membrane induces InsP3R2 to migrate away from the
the ER, which are required for proper protein folding and secre- periapical region, impairs Ca2+ signals, and reduces secretion
tion. This is achieved by active transport of free Ca2+ from the [13, 34]. Finally, InsP3R2 expression is decreased and the
cytosol to the ER cisternae. In experimental obesity, the down- remaining receptor moves from the apical region in two separate
regulation of Serca2b leads to reduced ER Ca2+ content and ER models of cholestasis induced by estrogen and endotoxin [34].
stress, which collectively activates glucose production and trig- Collectively, these studies suggest that subapical Ca2+ signals in
gers a lipogenic gene expression program. Conversely, re‐expres- hepatocytes potentiate secretion of bile solutes.
sion of Serca2b in obese mice restores ER protein folding and Bile acids such as lithocholic acid (LCA), taurolithocholic acid
improves both glucose and fatty acid metabolism. (TLCA), taurodeoxycholic acid and taurochenodeoxycholic, as
InsP3Rs have also been implicated in lipid processing and well as the therapeutic bile acids ursodeoxycholic acid (UDCA)
droplet formation in hepatocytes. Liver specific InsP3R1 knock- and tauroursodeoxycholic acid (TUDCA) increase intracellular
out (LSKO1) mice are resistant to the development of diet‐ Ca2+ in hepatocytes [69, 79]. LCA and TLCA are cholestatic, but
induced liver steatosis and this effect is due to impaired Ca2+ UDCA and TUDCA are choleretic. In fact, TUDCA can reverse
transfer from ER to mitochondria [18]. Similarly, mice fed a high‐ the cholestatic effect of TLCA, and this depends on InsP3R2 [33].
fat diet (HFD) or genetically obese mice display a reorganization UDCA and TUDCA also induce hepatocytes to secrete ATP,
of ER‐mitochondria contact sites [75]. These contact regions are which may relate to their therapeutic effect [80].
known as mitochondrial‐associated membranes (MAMs) and
facilitate the exchange of phospholipids and transmission of Ca2+
between the two organelles. In obesity, the overall percentage of
Cell proliferation
MAMs is increased, leading to InsP3R1‐dependent Ca2+ overload Ca2+ signals are associated with progression through the cell
in mitochondria and formation of reactive oxygen species (ROS) cycle [81]. In sea urchin embryos, there are two prominent cell
and mitochondrial dysfunction. Defects in SOCE also occur in cycle‐related Ca2+ signals. The first Ca2+ transient occurs just
experimental fatty liver disease. In obese mice, STIM1 transloca- prior to entry into mitosis, and the second one occurs during the
tion to the vicinity of the plasma membrane is impaired, which metaphase–anaphase transition. Intracellular injection of Ca2+
results in markedly diminished SOCE. These alterations in SOCE chelators such as BAPTA or an InsP3R antagonist such as hepa-
in turn are associated with lipid droplet formation and lipid toxic- rin abolish these Ca2+ signals and prevent entry into mitosis.
ity in hepatocytes [76]. Overall, pathological lipid accumulation Introduction of InsP3 or Ca2+ in the cytosol has the opposite
in hepatocytes and alterations in the expression of Ca2+ signaling effect, to accelerate entry into mitosis.
proteins are interrelated, and there is evidence that this occurs in Ca2+ transients are also observed during cell cycle progres-
human disease as well. In particular, there is a progressive sion in somatic cells, although the relationship between these
increase in the colocalization between ER and mitochondria in Ca2+ signals and progression through the cell cycle is less estab-
hepatocytes from normal liver to simple steatosis to non‐alcoholic lished. In Swiss 3T3 cells, serum withdrawal suppresses these
steatohepatitis (NASH). There also is increased expression of Ca2+ transients but does not affect progression though mitosis.
InsP3R1, which is responsible for mitochondrial Ca2+ signals, in However, mitosis is blocked by specific inhibition of Ca2+ tran-
liver biopsies from patients with NASH [18]. sients through injection of BAPTA plus incubation in calcium‐
free media. Moreover, photorelease of caged Ca2+ induces
premature entry into mitosis.
Bile flow Downstream targets of Ca2+ have also been implicated in cell
Cai2+ regulates fluid and electrolyte secretion in many types of cycle progression. Calcineurin is essential for Xenopus laevis
epithelia [25]. This is mediated in part by polarized Cai2+ waves, embryonic development [82]. In addition, pharmacological
which also occur in hepatocytes [77]. Cai2+ has multiple effects inhibition of Ca2+/calmodulin kinase II (CaMKII) arrests cells at
on bile flow. Early studies in isolated perfused rat liver showed the G2/M transition. Moreover, calmodulin overexpression
that the net effect of Cai2+ on bile acid‐independent bile flow is accelerates the cell cycle in mouse C127 cells and its downregu-
inhibitory [78], but more recent studies demonstrate that both lation extends the cell cycle.
bile acid‐independent and bile acid‐dependent flow are potenti- Heterologous expression of the Ca2+ binding protein parval-
ated by Ca2+ signaling in hepatocytes [33, 34]. First, the mem- bumin has also been used to study Ca2+ signaling in the regula-
brane localization and activity of the main apical bile‐acid tion of the cell cycle. This protein is normally expressed in
independent solute transporter, the multidrug resistance‐associ- skeletal muscle and neurons; in myocytes it modulates relaxa-
ated protein 2 (Mrp2) is potentiated by Ca2+ agonists in in vitro tion of fast twitch muscle fibers due to its Ca2+ buffering capac-
systems. Moreover, bile flow is decreased in InsP3R2 knockout ity. The first report using this protein as a molecular tool showed
animals suggesting that Ca2+ release near the apical membrane that buffering Ca2+ slowed progression through the cell cycle in
is important for bile secretion [33]. Total internal reflection flu- mouse C127 cells. More recently, parvalbumin variants targeted
orescence (TIRF) microscopy studies furthermore suggest that to the nucleus or the cytoplasm [60] were used to investigate the
periapical Ca2+ signals induce Mrp2‐containing vesicles to fuse relative importance of Ca2+ signals in each of these cellular
40: Ca2+ SIGNALING IN THE LIVER 505

compartments for regulation of the cell cycle in a liver cell line. exchanger 2 (AE2)‐mediated HCO3−/Cl− exchange that is respon-
It was found that nucleoplasmic rather than cytosolic Ca2+ is sible for secreting bicarbonate into the bile. Ca2+ potentiates ade-
essential for cellular proliferation, and is necessary in particular nylyl cyclase activity and cAMP formation in cholangiocytes via
for progression through early prophase [83]. Recent findings a calcineurin‐dependent pathway [94] and indirectly through
suggest that HGF and insulin, two potent growth factors in liver, cAMP formation induced by SOCE [95]. This crosstalk between
selectively form InsP3 in the nucleus to initiate nuclear Ca2+ sig- the Ca2+ and cAMP signaling pathways has been considered a
nals [3, 84]. These findings suggest that certain growth factors secondary, indirect mechanism for Ca2+‐stimulated secretion.
may stimulate proliferation of hepatocytes by selectively induc- However, cAMP‐ and CFTR‐mediated secretion may rely on Ca2+
ing Ca2+ signals in the nucleus. in a more direct way as well [96]. Activation of CFTR in cholan-
A common model of hepatocyte proliferation in vivo is liver giocytes is linked to exocytosis of vesicles rich in ATP [97], which
regeneration after partial hepatectomy. Following 70% hepatec- releases ATP into the bile, resulting in stimulation of apical P2Y
tomy, the liver undergoes a coordinated regenerative response receptors [51, 96]. Activation of these P2Y receptors links to
involving all hepatic cell types, but this is punctuated by marked GPCRs and InsP3 formation and then Ca2+ release from apical,
proliferation of hepatocytes. In rats, 24 hours after partial hepatec- InsP3R3, which in turn drives Ca2+‐dependent Cl− channels and
tomy there is a decrease in InsP3R2 expression accompanied by a then HCO3− secretion [96]. Conversely, neurotransmitters such
decrease in the frequency of Ca2+ oscillations. This decrease in as acetylcholine stimulate basolateral M3 muscarinic receptors
receptor expression normalizes within 4 days and is then associ- on cholangiocytes, leading to Ca2+ release from InsP3R1 and
ated with an increase in the frequency of Ca2+ oscillations. This InsP3R2. This Ca2+ similarly activates apical Ca2+‐dependent Cl−
remodeling of the Ca2+ signaling machinery is thought to be essen- channels, which ultimately stimulate biliary HCO3− secretion.
tial for the initial regenerative response [85]. However, complete Peptide hormones such as secretin stimulate basolateral secretin
loss of InsP3R2 results in delayed liver regeneration and also receptors, which induce the formation of cAMP, so this pathway
impaired Ca2+ signaling in the hepatocyte nucleus [86]. Fatty liver serves to activate CFTR leading to paracrine, ATP‐mediated
causes a c‐Jun mediated decrease in InsP3R2 expression in hepat- HCO3− secretion. This putative universal role of InsP3R‐mediated
ocytes [86], which may contribute in part to the impaired liver Ca2+ signaling in ductular secretion is in agreement with the
regeneration seen in patients with NAFLD. At least three growth observation that loss of InsP3R3 is a common molecular event
factors, HGF, EGF, and insulin, mobilize Cai2+ in hepatocytes and in cholestatic disorders including primary biliary cholangitis
contribute to liver regeneration [87–89]. Liver regeneration after (PBC), primary sclerosing cholangitis (PSC), extrahepatic biliary
partial hepatectomy is significantly delayed in rats expressing par- obstruction, and biliary atresia [98]. Factors that contribute to this
valbumin in the cytosol of hepatocytes [90]. Buffering of InsP3 in specific loss of InsP3R3 in biliary diseases include the transcrip-
the nucleus of hepatocytes also delays liver regeneration [89], sug- tion factors Nrf2 (nuclear factor erythroid 2‐related factor 2) and
gesting that RTK ‐mediated, InsP3‐dependent Ca2+ signals in the NF‐κB, plus the microRNA miR‐506. Nrf2 is activated during
nucleus are essential for proper hepatocyte proliferation. On the conditions of oxidative stress and specifically downregulates
other hand, liver regeneration is enhanced by expression of parval- InsP3R3 in a human cholangiocyte cell line. Accordingly, it is
bumin in mitochondria, because this mitigates apoptosis that is translocated to the nucleus of bile duct cells in a range of ductular
occurring concomitant with proliferation [20]. Therefore, the rate diseases including PBC and PSC [99]. The NF‐κB pathway is
of liver regeneration is determined by a careful balance between activated by endotoxin‐induced stimulation of TLR4, and this
Ca2+ signals in the cytosol, nucleus, and mitochondria. mechanism has been implicated in loss of InsP3R3 from cholan-
Ca2+‐dependent processes have been associated with progres- giocytes in patients with sepsis or alcoholic hepatitis [100]. The
sion of hepatocellular carcinoma (HCC). For instance, expression specific loss of InsP3R3 in PBC might also be mediated in part by
of Ca2+/calmodulin‐dependent kinase kinase II (CamKKII) is miR‐506, because this microRNA decreases InsP3R3 expression
upregulated in liver cancer and its experimental downregulation in a human cholangiocyte cell line and is increased in bile ducts
or pharmacological inhibition reduce tumor growth in a mouse in liver biopsies from PBC patients [101].
model of HCC [91]. Also, overexpression of hepatitis B virus X Cholangiocytes express primary cilia on their apical mem-
protein (HBx), a known oncogenic viral protein, promotes liver branes, and this sensory organelle also links to Ca2+ signaling
cancer cell invasion and metastasis through a Ca2+‐mediated and secretion. For example, primary cilia can sense and trans-
increase in secretion of HMGB1 [92]. The role of intracellular duce mechanical stimuli from within the bile duct lumen, so that
Ca2+ channels and downstream targets of Ca2+ signaling in the increases in bile flow activate both cAMP production and Ca2+
pathogenesis of liver cancer is a topic of active investigation. release [102]. Cholangiocyte cilia also sense increases in osmo-
larity within the bile duct lumen, and this links to activation of
TRPV4 channels, which allow entry of extracellular Ca2+ into
Ductular secretion the cytoplasm [103]. The role of cilia and ciliary Ca2+ signaling
Cai2+ directly regulates fluid and electrolyte secretion in bile in cholestatic disorders remains an area of investigation.
duct epithelia. Cholangiocytes express the apical, Ca2+‐activated The GPCR for bile acids (TGR5) also is expressed at the api-
Cl− channel TMEM16a [93], which is the primary mechanism cal membrane of cholangiocytes and it is enriched in cilia of
for Ca2+‐stimulated secretion in these cells. The other principal rodent and human bile duct cells [104]. TGR5 activation cou-
mechanism for regulating cholangiocyte secretion is via the api- ples to the formation of cAMP in cholangiocytes, so pharmaco-
cal, cAMP‐activated, cystic fibrosis transmembrane conductance logical agonists of this receptor are currently under study as
regulator (CFTR) Cl− channel [51]. Together, these two mecha- choleretic agents in a variety of acquired and genetic cholestatic
nisms create the Cl− gradient responsible for driving anion disorders.
506 THE LIVER:  REFERENCES

ACKNOWLEDGMENTS 22. Csordas, G., Renken, C., Varnai, P. et al. Structural and functional features
and significance of the physical linkage between ER and mitochondria. J
Cell Biol, 2006;174(7):915–21.
This work was supported by NIH grants DK57751, DK34989, 23. Hajn¢czky, G., Hager, R., and Thomas, A.P. Mitochondria suppress local
DK112797, and DK114041, and by grants from CNPq and feedback activation of inositol 1,4,5‐trisphosphate receptors by Ca2+. J Biol
FAPEMIG. Chem, 1999;274:14157–62.
24. Robb‐Gaspers, L.D., Burnett, P., Rutter, G.A., Denton, R.M., Rizzuto, R.,
and Thomas, A.P. Integrating cytosolic calcium signals into mitochondrial
metabolic responses. EMBO J, 1998;17:4987–5000.
25. Nathanson, M.H. Cellular and subcellular calcium signaling in gastrointesti-
REFERENCES nal epithelium. Gastroenterology, 1994;106:1349–64.
26. Zhao, Y., Araki, S., Wu, J. et al. An expanded palette of genetically encoded
1. Berridge, M.J., Bootman, M.D., and Roderick, H.L. Calcium signalling: Ca(2)(+) indicators. Science, 2011;333(6051):1888–91.
dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol, 27. Nathanson, M.H. and Burgstahler, A.D. Coordination of hormone‐induced
2003;4(7):517–29. calcium signals in isolated rat hepatocyte couplets: demonstration with con-
2. Rhee, S.G. and Choi, K.D. Regulation of inositol phospholipid‐specific focal microscopy. Mol Biol Cell, 1992;3:113–21.
phospholipase C isozymes. J Biol Chem, 1992;267:12393–6. 28. Rooney, T.A., Sass, E.J., and Thomas, A.P. Characterization of cytosolic cal-
3. Gomes, D.A., Rodrigues, M.A., Leite, M.F. et al. c‐Met must translocate to cium oscillations induced by phenylephrine and vasopressin in single Fura‐2
the nucleus to initiate calcium signals. J Biol Chem, 2008;283(7):4344–51. loaded hepatocytes. J Biol Chem, 1989;264:17131–41.
4. Futatsugi, A., Nakamura, T., Yamada, M.K. et  al. IP3 receptor types 2 29. Dolmetsch, R.E., Xu, K., and Lewis, R.S. Calcium oscillations increase the
and  3  mediate exocrine secretion underlying energy metabolism. Science, efficiency and specificity of gene expression. Nature, 1998;392(6679):933–6.
2005;309(5744):2232–4. 30. Jones, B.F., Boyles, R.R., Hwang, S.Y., Bird, G.S., and Putney, J.W. Calcium
5. Chandrasekhar, R., Alzayady, K.J., Wagner, L.E., 2nd, and Yule, D.I. Unique influx mechanisms underlying calcium oscillations in rat hepatocytes.
regulatory properties of heterotetrameric inositol 1,4,5‐trisphosphate recep- Hepatology, 2008;48(4):1273–81.
tors revealed by studying concatenated receptor constructs. J Biol Chem, 31. Feske, S., Skolnik, E.Y., and Prakriya, M. Ion channels and transporters in
2016;291(10):4846–60. lymphocyte function and immunity. Nat Rev Immunol, 2012;12(7):532–47.
6. Choe, C.U. and Ehrlich, B.E. The inositol 1,4,5‐trisphosphate receptor 32. Hernandez, E., Leite, M.F., Guerra, M.T. et  al. The spatial distribution of
(IP3R) and its regulators: sometimes good and sometimes bad teamwork. Sci inositol 1,4,5‐trisphosphate receptor isoforms shapes Ca2+ waves. J Biol
STKE, 2006;2006(363):re15. Chem, 2007;282(13):10057–67.
7. Bosanac, I., Alattia, J.R., Mal, T.K. et al. Structure of the inositol 1,4,5‐tris- 33. Cruz, L.N., Guerra, M.T., Kruglov, E. et al. Regulation of multidrug resist-
phosphate receptor binding core in complex with its ligand. Nature, ance‐associated protein 2 by calcium signaling in mouse liver. Hepatology,
2002;420(6916):696–700. 2010;52(1):327–37.
8. Hagar, R.E., Burgstahler, A.D., Nathanson, M.H., and Ehrlich, B.E. Type III 34. Kruglov, E.A., Gautam, S., Guerra, M.T., and Nathanson, M.H. Type 2 ino-
InsP3 receptor channel stays open in the presence of increased calcium. sitol 1,4,5‐trisphosphate receptor modulates bile salt export pump activity in
Nature, 1998;396(6706):81–4. rat hepatocytes. Hepatology, 2011;54(5):1790–9.
9. Bimboese, P., Gibson, C.J., Schmidt, S., Xiang, W., and Ehrlich, B.E. 35. Motoyama, K., Karl, I.E., Flye, M.W., Osborne, D.F., and Hotchkiss, R.S.
Isoform‐specific regulation of the inositol 1,4,5‐trisphosphate receptor by Effect of Ca2+ agonists in the perfused liver: determination via laser scan-
O‐linked glycosylation. J Biol Chem, 2011;286(18):15688–97. ning confocal microscopy. Am J Physiol Regul Integr Comp Physiol,
10. Dufour J‐F., Luthi, M., Forestier, M., and Magnino, F. Expression of inositol 1999;276:R575–85.
1,4,5‐trisphosphate receptor isoforms in rat cirrhosis. Hepatology, 1999;30: 36. Nathanson, M.H., Burgstahler, A.D., Mennone, A., Fallon, M.B., Gonzalez,
1018–26. C.B., and Saez, J.C. Ca2+ waves are organized among hepatocytes in the
11. Hirata, K., Pusl, T., O’Neill A.F., Dranoff, J.A., and Nathanson, M.H. The intact organ. Am J Physiol, 1995;269(1 Pt 1):G167–71.
type II inositol 1,4,5‐trisphosphate receptor can trigger Ca2+ waves in rat 37. Tordjmann, T., Berthon, B., Combettes, L., and Claret, M. The location of
hepatocytes. Gastroenterology, 2002;122(4):1088–100. hepatocytes in the rat liver acinus determines their sensitivity to calcium‐
12. Hirata, K., Dufour, J.F., Shibao, K. et al. Regulation of Ca(2+) signaling in mobilizing hormones. Gastroenterology, 1996;111:1343–52.
rat bile duct epithelia by inositol 1,4,5‐trisphosphate receptor isoforms. 38. Saez, J.C., Connor, J.A., Spray, D.C., and Bennett, M.V.L. Hepatocyte gap
Hepatology, 2002;36(2):284–96. junctions are permeable to the second messenger, inositol 1,4,5‐triphosphate,
13. Nagata, J., Guerra, M.T., Shugrue, C.A., Gomes, D.A., Nagata, N., and and to calcium ions. Proc Natl Acad Sci USA, 1989;86:2708–12.
Nathanson, M.H. Lipid rafts establish calcium waves in hepatocytes. 39. Fallon, M.B., Nathanson, M.H., Mennone, A., Sáez J.C., Burgstahler, A.D.,
Gastroenterology, 2007;133(1):256–67. and Anderson, J.M. Altered expression and function of hepatocyte gap junc-
14. Pierobon, N., Renard‐Rooney, D.C., Gaspers, L.D., and Thomas, A.P. tions after common bile duct ligation in the rat. Am J Physiol Cell Physiol,
Ryanodine receptors in liver. J Biol Chem, 2006;281(45):34086–95. 1995;268:C1186–94.
15. Zhang, F. and Li, P.L. Reconstitution and characterization of a nicotinic acid 40. Correa, P.R., Guerra, M.T., Leite, M.F., Spray, D.C., and Nathanson, M.H.
adenine dinucleotide phosphate (NAADP)‐sensitive Ca2+ release channel Endotoxin unmasks the role of gap junctions in the liver. Biochem Biophys
from liver lysosomes of rats. J Biol Chem, 2007;282(35):25259–69. Res Commun, 2004;322(3):718–26.
16. De Stefani, D., Rizzuto, R., and Pozzan, T. Enjoy the trip: calcium in mito- 41. Leite, M.F., Hirata, K., Pusl, T. et al. Molecular basis for pacemaker cells in
chondria back and forth. Annu Rev Biochem, 2016;85:161–92. epithelia. J Biol Chem, 2002;277:16313–23.
17. Antony, A.N., Paillard, M., Moffat, C. et al. MICU1 regulation of mitochon- 42. Nathanson, M.H., Rios‐Velez, L., Burgstahler, A.D., and Mennone, A.
drial Ca(2+) uptake dictates survival and tissue regeneration. Nat Comm, Communication via gap junctions modulates bile secretion in the isolated
2016;7:10955. perfused rat liver. Gastroenterology, 1999;116:1176–83.
18. Feriod, C.N., Oliveira, A.G., Guerra, M.T. et al. Hepatic inositol 1,4,5 tris- 43. Bartlett, P.J., Gaspers, L.D., Pierobon, N., and Thomas, A.P. Calcium‐
phosphate receptor type 1 mediates fatty liver. Hepatol Commun, dependent regulation of glucose homeostasis in the liver. Cell Calcium,
2017;1(1):23–35. 2014;55(6):306–16.
19. White, C., Li, C., Yang, J. et al. The endoplasmic reticulum gateway to apopto- 44. Stumpel, F., Ott, T., Willecke, K., and Jungermann, K. Connexin 32 gap
sis by Bcl‐X(L) modulation of the InsP3R. Nat Cell Biol, 2005;7(10):1021–8. junctions enhance stimulation of glucose output by glucagon and noradrena-
20. Guerra, M.T., Fonseca, E.A., Melo, F.M. et al. Mitochondrial calcium regu- line in mouse liver. Hepatology, 1998;28:1616–20.
lates rat liver regeneration through the modulation of apoptosis. Hepatology, 45. Patel, S.J., Milwid, J.M., King, K.R. et al. Gap junction inhibition prevents
2011;54(1):296–306. drug‐induced liver toxicity and fulminant hepatic failure. Nat Biotechnol,
21. Minagawa, N., Kruglov, E.A., Dranoff, J.A., Robert, M.E., Gores, G.J., and 2012;30(2):179–83.
Nathanson, M.H. The anti‐apoptotic protein Mcl‐1 inhibits mitochondrial 46. Boitano, S., Dirksen, E.R., and Sanderson, M.J. Intercellular propagation of
Ca2+ signals. J Biol Chem, 2005;280(39):33637–44. calcium waves mediated by inositol trisphosphate. Science, 1992;258:292–5.
40: Ca2+ SIGNALING IN THE LIVER 507

47. Tordjmann, T., Berthon, B., Claret, M., and Combettes, L. Coordinated 71. Ozcan, L., Wong, C.C., Li, G. et  al. Calcium signaling through CaMKII
­intercellular calcium waves induced by noradrenaline in rat hepatocytes: regulates hepatic glucose production in fasting and obesity. Cell Metabol,
dual control by gap junction permeability and agonist. EMBO J, 1997;16: 2012;15(5):739–51.
5398–407. 72. Wang, Y., Li, G., Goode, J. et al. Inositol‐1,4,5‐trisphosphate receptor regu-
48. Tordjmann, T., Berthon, B., Jacquemin, E. et al. Receptor‐oriented intercel- lates hepatic gluconeogenesis in fasting and diabetes. Nature, 2012;485(7396):
lular calcium waves evoked by vasopressin in rat hepatocytes. EMBO J, 128–32.
1998;17:4695–703. 73. Kruglov, E., Ananthanarayanan, M., Sousa, P., Weerachayaphorn, J., Guerra,
49. Burgstahler, A.D. and Nathanson, M.H. Coordination of calcium waves among M.T., and Nathanson, M.H. Type 2 inositol trisphosphate receptor gene
hepatocytes: teamwork gets the job done. Hepatology, 1998;27:634–5. expression in hepatocytes is regulated by cyclic AMP. Biochem Biophys Res
50. Schlosser, S.F., Burgstahler, A.D., and Nathanson, M.H. Isolated rat hepato- Commun, 2017;486(3):659–64.
cytes can signal to other hepatocytes and bile duct cells by release of nucleo- 74. Park, S.W., Zhou, Y., Lee, J., Lee, J., and Ozcan, U. Sarco(endo)plasmic
tides. Proc Natl Acad Sci USA, 1996;93:9948–53. reticulum Ca2+‐ATPase 2b is a major regulator of endoplasmic reticulum
51. Fiorotto, R., Spirli, C., Fabris, L., Cadamuro, M., Okolicsanyi, L., and stress and glucose homeostasis in obesity. Proc Natl Acad Sci USA,
Strazzabosco, M. Ursodeoxycholic acid stimulates cholangiocyte fluid 2010;107(45):19320–5.
secretion in mice via CFTR‐dependent ATP secretion. Gastroenterology, 75. Arruda, A.P., Pers, B.M., Parlakgul, G., Guney, E., Inouye, K., and
2007;133(5):1603–13. Hotamisligil, G.S. Chronic enrichment of hepatic endoplasmic reticulum‐
52. Nathanson, M.H., Burgstahler, A.D., Mennone, A., and Boyer, J.L. mitochondria contact leads to mitochondrial dysfunction in obesity. Nat
Characterization of cytosolic Ca2+ signaling in rat bile duct epithelia. Am J Med, 2014;20(12):1427–35.
Physiol Gastrointest Liver Physiol, 1996;271:G86–96. 76. Arruda, A.P., Pers, B.M., Parlakgul, G. et al. Defective STIM‐mediated store
53. Kitamura, T., Brauneis, U., Gatmaitan, Z., and Arias, I.M. Extracellular ATP, operated Ca(2+) entry in hepatocytes leads to metabolic dysfunction in
intracellular calcium and canalicular contraction in rat hepatocyte doublets. ­obesity. Elife, 2017;6.
Hepatology, 1991;14:640–7. 77. Nathanson, M.H., Burgstahler, A.D., and Fallon, M.B. Multi‐step mecha-
54. Gonzales, E., Prigent, S., Abou‐Lovergne, A. et al. Rat hepatocytes express nism of polarized Ca2+ wave patterns in hepatocytes. Am J Physiol
functional P2X receptors. FEBS Lett, 2007;581(17):3260–6. Gastrointest Liver Physiol, 1994;267:G338–49.
55. Patel, S., Robb‐Gaspers, L.D., Stellato, K.A., Shon, M., and Thomas, A.P. 78. Nathanson, M.H., Gautam, A., Bruck, R., Isales, C.M., and Boyer, J.L.
Coordination of calcium signalling by endothelial‐derived nitric oxide in the Effects of Ca2+ agonists on cytosolic Ca2+ in isolated hepatocytes and
intact liver. Nat Cell Biol, 1999;1:467–71. on  bile secretion in the isolated perfused rat liver. Hepatology,
56. Lanini, L., Bachs, O., and Carafoli, E. The calcium pump of the liver nuclear 1992;15:107–16.
membrane is identical to that of endoplasmic reticulum. J Biol Chem, 79. Beuers, U., Nathanson, M.H., and Boyer, J.L. Effects of tauroursodeoxy-
1992;267(16):11548–52. cholic acid on cytosolic Ca2+ signals in isolated rat hepatocytes.
57. Gerasimenko, J.V., Maruyama, Y., Yano, K. et al. NAADP mobilizes Ca2+ Gastroenterology, 1993;104:604–12.
from a thapsigargin‐sensitive store in the nuclear envelope by activating 80. Nathanson, M.H., Burgstahler, A.D., Masyuk, A., and Larusso, N.F.
ryanodine receptors. J Cell Biol, 2003;163(2):271–82. Stimulation of ATP secretion in the liver by therapeutic bile acids.
58. Gerasimenko, O.V., Gerasimenko, J.V., Tepikin, A.V., and Petersen, O.H. Biochem, J., 2001;358(1):1–5.
ATP‐dependent accumulation and inositol trisphosphate‐ or cyclic ADP‐ 81. Poenie, M., Alderton, J., Steinhardt, R., and Tsien, R. Calcium rises abruptly
ribose‐mediated release of Ca2+ from the nuclear envelope. Cell, and briefly throughout the cell at the onset of anaphase. Science, 1986;233(4766):
1995;80:439–44. 886–9.
59. Malhas, A., Goulbourne, C., and Vaux, D.J. The nucleoplasmic reticulum: 82. Nishiyama, T., Yoshizaki, N., Kishimoto, T., and Ohsumi, K. Transient acti-
form and function. Trends Cell Biol, 2011;21(6):362–73. vation of calcineurin is essential to initiate embryonic development in
60. Pusl, T., Wu, J.J., Zimmerman, T.L. et al. Epidermal growth factor‐mediated Xenopus laevis. Nature, 2007;449(7160):341–5.
activation of the ETS domain transcription factor Elk‐1 requires nuclear cal- 83. Rodrigues, M.A., Gomes, D.A., Leite, M.F. et al. Nucleoplasmic calcium is
cium. J Biol Chem, 2002;277(30):27517–27. required for cell proliferation. J Biol Chem, 2007;282(23):17061–8.
61. Leite, M.F., Thrower, E.C., Echevarria, W. et al. Nuclear and cytosolic cal- 84. Rodrigues, M.A., Gomes, D.A., Andrade, V.A., Leite, M.F., and Nathanson,
cium are regulated independently. Proc Natl Acad Sci USA, M.H. Insulin induces calcium signals in the nucleus of rat hepatocytes.
2003;100(5):2975–80. Hepatology, 2008;48(5)1621–31.
62. Divecha, N., Rhee S‐G., Letcher, A.J., and Irvine, R.F. Phosphoinositide sig- 85. Nicou, A., Serriere, V., Hilly, M. et al. Remodelling of calcium signalling
nalling enzymes in rat liver nuclei: phosphoinositidase C isoform ·1 is spe- during liver regeneration in the rat. J Hepatol, 2007;46(2):247–56.
cifically, but not predominantly, located in the nucleus. Biochem J, 86. Khamphaya, T., Chukijrungroat, N., Saengsirisuwan, V. et al. Nonalcoholic
1993;289:617–20. fatty liver disease impairs expression of the type II inositol 1,4,5‐trisphos-
63. Waybill, M.M., Yelamarty, R.V., Zhang, Y. et al. Nuclear calcium gradients phate receptor. Hepatology, 2018;67(2):560–74.
in cultured rat hepatocytes. Am J Physiol Endocrinol Metab, 87. Paranjpe, S., Bowen, W.C., Bell, A.W., Nejak‐Bowen, K., Luo, J.H., and
1990;261:E49–57. Michalopoulos, G.K. Cell cycle effects resulting from inhibition of hepato-
64. O’Brien E.M., Gomes, D.A., Sehgal, S., and Nathanson, M.H. Hormonal cyte growth factor and its receptor c‐Met in regenerating rat livers by RNA
regulation of nuclear permeability. J Biol Chem, 2007;282(6):4210–7. interference. Hepatology, 2007;45(6):1471–7.
65. Lipp, P., Thomas, D., Berridge, M.J., and Bootman, M.D. Nuclear calcium 88. Rodrigues, M.A., Gomes, D.A., Andrade, V.A., Leite, M.F., and Nathanson,
signalling by individual cytoplasmic calcium puffs. EMBO J, M.H. Insulin induces calcium signals in the nucleus of rat hepatocytes.
1997;16:7166–73. Hepatology, 2008.
66. Bode, H.P., Wang, L., Cassio, D. et  al. Expression and regulation of gap 89. Amaya, M.J., Oliveira, A.G., Guimaraes, E.S. et  al. The insulin receptor
junctions in rat cholangiocytes. Hepatology, 2002;36(3):631–40. translocates to the nucleus to regulate cell proliferation in liver. Hepatology,
67. Blackmore, P.F., Strickland, W.G., Bocckino, S.B., and Exton, J.H. 2014;59(1):274–83.
Mechanism of hepatic glycogen synthase in activation induced by Ca2+‐ 90. Lagoudakis, L., Garcin, I., Julien, B. et al. Cytosolic calcium regulates liver
mobilizing hormones. Biochem J, 1986;237:235–42. regeneration in the rat. Hepatology, 2010;52(2):602–11.
68. Keppens, S. and DeWulf, H. Characterization of the liver P2‐purinoceptor 91. Lin, F., Marcelo, K.L., Rajapakshe, K. et al. The camKK2/camKIV relay
involved in the activation of glycogen phosphorylase. Biochem J, is  an essential regulator of hepatic cancer. Hepatology, 2015;62(2):
1986;240:367–71. 505–20.
69. Combettes, L., Dumont, M., Berthon, B., Erlinger, S., and Claret, M. Release 92. Chen, S., Dong, Z., Yang, P. et al. Hepatitis B virus X protein stimulates high
of calcium from the endoplasmic reticulum by bile acids in rat liver cells. J mobility group box 1 secretion and enhances hepatocellular carcinoma
Biol Chem, 1988;263:2299–303. metastasis. Cancer Lett, 2017;394:22–32.
70. Halpern, K.B., Shenhav, R., Matcovitch‐Natan, O. et al. Single‐cell spatial 93. Li, Q., Dutta, A., Kresge, C., Bugde, A., and Feranchak, A.P. Bile acids stim-
reconstruction reveals global division of labour in the mammalian liver. ulate cholangiocyte fluid secretion by activation of transmembrane member
Nature, 2017;542(7641):352–6. 16A Cl(‐) channels. Hepatology, 2018;68(1):187–99.
508 THE LIVER:  REFERENCES

94. Alvaro, D., Alpini, G., Jezequel, A.M. et al. Role and mechanisms of action 100. Franca, A., Filho, A., Guerra, M.T. et  al. Effects of endotoxin on type 3
of acetylcholine in the regulation of rat cholangiocyte secretory function. J inositol 1,4,5‐trisphosphate receptor in human cholangiocytes. Hepatology,
Clin Invest, 1997;100:1349–62. 2019;69(2):817–830.
95. Spirli, C., Locatelli, L., Fiorotto, R. et  al. Altered store operated calcium 101. Ananthanarayanan, M., Banales, J.M., Guerra, M.T. et  al. Post‐transla-
entry increases cyclic 3’,5’‐adenosine monophosphate production and extra- tional regulation of the type III inositol 1,4,5‐trisphosphate receptor by
cellular signal‐regulated kinases 1 and 2 phosphorylation in polycystin‐2‐ miRNA‐506. J Biol Chem, 2015;290(1):184–96.
defective cholangiocytes. Hepatology, 2012;55(3):856–68. 102. Masyuk, A.I., Masyuk, T.V., Splinter, P.L., Huang, B.Q., Stroope, A.J.,
96. Minagawa, N., Nagata, J., Shibao, K. et al. Cyclic AMP regulates b­ icarbonate and  LaRusso, N.F. Cholangiocyte cilia detect changes in luminal fluid
secretion in cholangiocytes through release of ATP into bile. Gastroenterology, flow  and transmit them into intracellular Ca2+ and cAMP signaling.
2007;133(5):1592–602. Gastroenterology, 2006;131(3):911–20.
97. Sathe, M.N., Woo, K., Kresge, C. et al. Regulation of purinergic signaling in 103. Gradilone, S.A., Masyuk, A.I., Splinter, P.L. et  al. Cholangiocyte cilia
biliary epithelial cells by exocytosis of SLC17A9‐dependent ATP‐enriched express TRPV4 and detect changes in luminal tonicity inducing bicarbo-
vesicles. J Biol Chem, 2011;286(28):25363–76. nate secretion. Proc Natl Acad Sci USA, 2007;104(48):19138–43.
98. Shibao, K., Hirata, K., Robert, M.E., and Nathanson, M.H. Loss of inositol 104. Keitel, V., Ullmer, C., and Haussinger, D. The membrane‐bound bile acid
1,4,5‐trisphosphate receptors from bile duct epithelia is a common event in receptor TGR5 (Gpbar‐1) is localized in the primary cilium of cholangio-
cholestasis. Gastroenterology, 2003;125(4):1175–87. cytes. Biol Chem, 2010;391(7):785–9.
99. Weerachayaphorn, J., Amaya, M.J., Spirli, C. et  al. Nuclear factor, eryth- 105. Pusl, T. and Nathanson, M.H. The role of inositol 1,4,5‐trisphosphate
roid  2‐like 2 regulates expression of type 3 inositol 1,4,5‐trisphosphate receptors in the regulation of bile secretion in health and disease. Biochem
receptor and calcium signaling in cholangiocytes. Gastroenterology, Biophys Res Commun, 2004;322(4):1318–25.
2015;149(1):211–22. 106. Gaspers, L.D. and Thomas, A.P. Calcium signaling in liver. Cell Calcium,
2005;38(3–4):329–42.
Clinical Genomics
41 of NAFLD
Frank Lammert
Department of Medicine II, Saarland University Medical Center, Homburg, Germany

INTRODUCTION States, prevalent NAFLD cases are forecasted to increase 21%


from 83 million in 2015 to 101 million 2030, while prevalent
To date fatty liver is a common liver disease worldwide and is to NASH cases will increase 63% from 17 million to 27 million
be included in the differential diagnosis of elevated liver cases. The overall NAFLD prevalence among the population
enzymes, in particular in the setting of indicators of the meta- aged greater than or equal to 15 years is projected at 34% in
bolic syndrome, such as central obesity, dyslipidemia, hyperten- 2030. In 2015, approximately 20% of NAFLD cases were clas-
sion, or increased fasting plasma glucose. The prevalence of sified as NASH, increasing to 27% by 2030. Incidence of
non‐alcoholic fatty liver disease (NAFLD) depends on the defi- decompensated cirrhosis increases 170% to 105 000 cases by
nition, since liver fat content represents a continuous parameter. 2030, while incidence of hepatocellular cancer (HCC) will
In liver biopsies, steatosis (NAFL) is diagnosed when hepatic increase by 140% to 12 000 cases. Liver deaths will increase
steatosis exceeds 5%, and non‐alcoholic steatohepatitis (NASH) 180% to an estimated 78 000 deaths in 2030, and during 2015–
is defined as the presence of at least 5% hepatic steatosis plus 2030, there are projected to be nearly 800 000 excess liver
inflammation with hepatocyte injury [1]. In the Dallas heart deaths [5].
study (a multi‐ethnic population‐based probability sample of Family studies show that first degree relatives of patients with
Dallas County residents), the upper 95th percentile of liver fat NAFLD display a markedly higher risk to develop NAFLD than
measured by proton magnetic spectroscopy (1H‐MRS) in the general population, independent of obesity. Recently,
healthy subjects was 5.6%, corresponding to 15% histological ­magnetic resonance imaging was used to quantify hepatic stea-
liver fat, and according to this definition, 31% of the cohort tosis (PDFF) and liver fibrosis in 60 pairs of twins residing in
­displayed hepatic steatosis [2]. When measured by magnetic Southern California [6]. The presence of hepatic steatosis and
resonance imaging‐based techniques such as the proton density the stage of liver fibrosis correlated between monozygotic but
fat fraction (PDFF), 5% steatosis corresponds to a PDFF of 6.0 not dizygotic twins. In multivariate models adjusted for age,
to 6.4% [3]. sex, and ethnicity, the heritability of hepatic steatosis (based on
The presence of cirrhosis with current or previous histopatho- PDFF) and liver fibrosis was 50%.
logical evidence of steatosis or NASH defines NASH‐cirrhosis.
These definitions of fatty liver diseases imply that there is a gradi-
ent from common benign NAFL to defined subgroups of patients
with more advanced disease such as NASH or cirrhosis. Given PNPLA3 (ADIPONUTRIN) VARIANT
the lack of reliable non‐invasive markers, a liver biopsy is still AND FATTY LIVER DISEASE
formally required to detect or exclude the presence of NASH [4].
Modeling the epidemic of NAFLD until 2030 in a Markov Adiponutrin, the enzyme encoded by the PNPLA3 gene (aka
model based on historical and projected changes in adult preva- calcium‐independent phospholipase A2ε), is a 481‐amino acid
lence of obesity and type 2 diabetes mellitus demonstrates an member of the patatin‐like phospholipase domain‐containing
exponential increase in burden of NAFLD [5]. In the United family (PNPLA). This domain was originally discovered in

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
510 THE LIVER:  PNPLA3 (ADIPONUTRIN) VARIANT AND FATTY LIVER DISEASE

lipid hydrolases of potato and named after the most abundant that the PNPLA3 variant is associated with steatosis, portal and
protein of the potato tuber, patatin. However, since several fam- lobular inflammation, NAFLD activity score (NAS), and fibro-
ily members are not phospholipases, a more appropriate gene sis. The first meta‐analysis by Sookoian and Pirola [9] included
designation has been called for [7]. PNPLA3 is expressed data from 16 studies and concluded that the PNPLA3 p.I148M
­predominantly in the liver, retina, skin, and adipose tissue [8]. In variant is associated with fatty liver (odds ratio [OR] for
a series of seminal genetic studies, the common non‐synony- homozygous carriers 3.3 and heterozygous carriers 1.9), NASH
mous variant p.I148M (c.444C>G, rs738409) of the PNPLA3 (OR 3.3 and 2.7, respectively), necroinflammation (OR 3.2 and
gene has emerged as the key genetic determinant of fatty liver 2.6, respectively), and fibrosis (OR 3.3 and 2.1, respectively);
disease in pediatric and adult patients [9] (Table  41.1). the association with necroinflammation and fibrosis was inde-
Figure  41.1 summarizes the geographical distribution of pendent of the severity of steatosis. The recent meta‐analysis by
PNPLA3 p.I148M genotypes in patients with NAFLD. Of note, Xu et al. [22] confirmed these results, with subgroup and sensi-
the risk allele frequency is 25% in Europeans [10]. The tivity analyses showing that the results were neither influenced
­prevalence differs among ethnic groups, and these differences by ethnicities nor the age of subjects or the source of controls.
generally parallel those for NASH and its sequelae. The variant Using two histopathologically characterized cohorts encom-
is most prevalent in Hispanics (49%), is less common in non‐ passing steatosis, steatohepatitis, fibrosis, and cirrhosis
Hispanic Europeans (23%), and is least common in African (N = 1074), Liu et al. [23] confirmed that the PNPLA3 variant is
Americans (17%) [2, 11]. associated with advanced fibrosis/cirrhosis, which is independ-
The first genome‐wide association study (GWAS) in the ent of age, BMI, and type 2 diabetes as confounders (Table 41.2).
­population‐based Dallas heart study comprising 2111 individu- Availing of transient elastography to quantify liver fibrosis in
als with different ethnic backgrounds demonstrated the p.I148M 899 patients with chronic liver diseases, an association between
variant to be associated (P = 5.9 × 10−10) with increased liver fat the PNPLA3 mutation and enhanced liver stiffness was identi-
content, as determined by 1H‐magnetic resonance spectrometry fied in a wide spectrum of viral and nonviral chronic liver
[2]. There was a stepwise increase of hepatic triglyceride ­content ­diseases [24]. Sensitivity analysis showed that the association
with the number of p.I148M risk alleles. The association was was present across a broad range of stiffness values (12–40 kPa)
most prominent in Hispanic patients, who are in general at a [24], indicating that the variant affects not only fibrogenesis but
greater risk of developing fatty liver as compared to European also cirrhosis severity.
and African Americans. Non‐obese NAFLD is defined as NAFLD that develops in
Of note, the PNPLA3 variant p.I148M is also associated with patients with a BMI less than 25 kg m−2 [25]. In patients with
serum activities of liver enzymes. The analyses of two large non‐obese NAFLD, the PNPLA3 p.I148M allele is more
population‐based cohorts demonstrated that PNPLA3 polymor- ­frequent than in other NAFLD patients [11, 26–28] and inde-
phisms are associated with serum alanine aminotransferase pendently associated with both NASH and fibrosis stage greater
(ALT) and γ‐glutamyl transpeptidase activities in healthy indi- than or equal to F2 [29].
viduals [12, 13]. The genetic association between the PNPLA3 Carriers of the PNPLA3 mutation who are obese [30] or pre-
mutation p.I148M and fatty liver disease was subsequently sent with NASH [31] are predisposed to HCC development. In
­replicated in many studies [14–17]. multivariate analysis adjusted for age, sex, BMI, diabetes, and
Further studies showed that the PNPLA3 variant not only cirrhosis, carriage of each copy of the PNPLA3 risk allele con-
increases the odds of developing fatty liver itself but it also ferred an additive risk for HCC (OR 2.3), with homozygotes
determines the degree of hepatic injury and the full spectrum of exhibiting a fivefold increased risk over the wild‐type genotype.
histopathological consequences of NAFLD [18]. Valenti et al. When compared to the UK general population, the risk‐effect
[19, 20] reported that the PNPLA3 variant was not only associ- was more pronounced (OR 12.2) [31]. Further analysis of the
ated with hepatic steatosis but with NAFLD severity as deter- data found the positive predictive value of p.I148M to be weak,
mined by liver biopsy, in particular the presence of NASH, but the negative predictive value of the absence of p.I148M is
steatosis grade greater than S1, and fibrosis stage greater than strong (97%). Hence, prospective studies are warranted to
F1, independent of age, body mass index (BMI), or diabetes. ­validate the clinical utility of PNPLA3 genotyping to select the
Another detailed analysis of histopathological severity of patients who are least likely to develop HCC and therefore least
NAFLD was performed by Rotman et al. [21], demonstrating likely to benefit from surveillance [32]. In a series of newly

Table 41.1  Key examples of liver diseases associated with variant PNPLA3


Disease Study Year Reference
Non‐alcoholic fatty liver disease Romeo et al. 2008 [2]
Alcoholic liver cirrhosis Tian et al. 2010 [98]
Stickel et al. 2011 [99]
Liver fibrosis Krawczyk et al. 2011 [24]
HBV steatosis Vigano et al. 2013 [100]
HCV steatosis Cai et al. 2011 [101]
Alcohol and HCV cirrhosis Müller et al. 2011 [102]
HCV cirrhosis Valenti and Fargion 2011 [103]
Hepatocellular cancer Nischalke et al. 2011 [104]
HBV, hepatitis B virus; HCV, hepatitis C virus; PNPLA3, patatin‐like phospholipase domain‐containing protein 3.
41:  Clinical Genomics of NAFLD 511

Figure 41.1  Worldwide estimated prevalence of NAFLD and distribution of PNPLA3 genotypes. PNPLA3 genotype is presented as minor risk
allele frequency (light blue section of the pie chart). Reproduced with permission of Springer Nature from [105].

Table 41.2  Multivariate analysis of association between PNPLA3 and TM6SF2 genotypes and fibrosis stages F0–1 (mild) versus F2–4
(advanced)
Discovery cohort (n = 349) Validation cohort (n = 725) Combined cohort (n = 1047)
Variables OR (95% CI) P‐value OR (95% CI) P‐value OR (95% CI) P‐value
TM6SF2 genotype 2.94 (1.76–4.89) 3.44 × 10−5 1.46 (1.03–2.09) 0.0362 1.88 (1.41–2.5) 1.63 × 10−5
PNPLA3 genotype 1.57 (1.21–2.19) 0.0086 1.32 (1.05–1.66) 0.0183 1.40 (1.16–1.69) 4.84 × 10−4
Age 1.03 (1.01–1.06) 0.0045 1.02 (1.01–1.04) 0.0041 1.03 (1.01–1.04) 1.57 × 10−5
Gender (female) 0.94 (0.57–1.56) 0.8297 1.81 (1.30–2.50) 4.50 × 10−4 1.43 (1.09–1.89) 0.0096
BMI 1.05 (1.00–1.10) 0.0368 1.03 (1.01–1.05) 9.80 × 10−4 1.04 (1.02–1.05) 3.78 × 10−5
T2DM 2.39 (1.49–3.84) 0.0003 2.73 (1.93–3.88) 1.68 × 10−8 2.57 (1.95–3.39) 1.78 × ×10−11
BMI, body mass index; CI, confidence interval; OR, odds ratio; T2DM, type 2 diabetes mellitus.
Additive model including age, gender, BMI, T2DM, and PNPLA3 rs738409 genotype as covariates. Discovery/validation/combined cohorts: stage F0–1 (mild) n = 198/439/637,
stage F2–4 (advanced) n = 151/286/437.

diagnosed HCC cases  [33], homozygosity for the genotype Variant PNPLA3 and metabolic traits
p.I148M was even an independent risk factor for death; HCC
patients with this PNPLA3 genotype displayed reduced median In general, patients with fatty liver disease often present with
survival (16.8 months) in comparison to carriers of the wild‐ dyslipidemia and insulin resistance [35]. However, the associ-
type allele (25.9 months). ation of PNPLA3 with metabolic traits in humans is more com-
Overall the above studies document that the PNPLA3 mutation plex. Several studies did not detect a relationship between the
increases the risk of developing severe hepatic fat accumulation, p.I148M polymorphism and serum glucose or lipid concentra-
progressive inflammation, advanced fibrosis, and HCC across dif- tions [17, 19, 36, 37]. In contrast, a Danish analysis of more
ferent ethnicities worldwide (Figure  41.2). Quantitative analyses than 4000 individuals with normal glucose tolerance demon-
showed that adiposity was shown to amplify the genetic risk con- strated an association of the risk allele with increased fasting
ferred by PNPLA3 across the full spectrum of NAFLD, ranging glucose levels, but the same allele was associated with lower
from increased hepatic triglyceride content to cirrhosis (Figure 41.3) serum concentrations of triglycerides and cholesterol in
[34]. Additional studies are needed to determine the cost effective- patients with impaired glucose intolerance [38]. Lower fasting
ness and utility of multifactorial risk stratification that incorporates triglyceride levels were also observed in severely overweight
PNPLA3 genotype and other risk factors for HCC. patients carrying the risk variant [39]. In line with these results,
512 THE LIVER:  PNPLA3 (ADIPONUTRIN) VARIANT AND FATTY LIVER DISEASE

Figure 41.2  Model illustrating the regulation of PNPLA3 gene expression in liver and the link between the p.I148M variant and progressive liver
disease. The PNPLA3 variant drives the full phenotypic spectrum from steatosis to steatohepatitis, fibrosis, cirrhosis, and HCC. Note that ChREBP
controls Pnpla3 mRNA levels in mouse hepatocytes, whereas the specific ChREBP response element is missing in the human PNPLA3 promoter.
Abbreviations: ChREBP, carbohydrate response element binding protein; HCC, hepatocellular cancer; SREBP1c, sterol regulatory element binding
protein 1c. Reproduced with permission of Elsevier from [106].

(a) (b)
25 M variant
II M variant
IM 10 II
20 MM IM
MM
OR (95%CI)
HTGC (%)

15
5
10

2
5
1
<25 25–30 30–35 >35
0 BMI (kg/m2)
<25 25–30 30–35 >35 M variant
BMI (kg/m2) n total II 24,787 22,159 7,034 2,026
IM 14,528 13,079 4,156 1,232
n
MM 2,086 1,851 608 173
PNPLA3 II 360 526 396 342 24
n events II 77 67
M variant IM 90 309 199 171 34 10
IM 48 69
MM 33 60 57 31 7 8
MM 20
16 4

Figure 41.3  (a) Hepatic triglyceride content (HTGC) measured by magnetic resonance spectroscopy, stratified PNPLA3 genotype, and BMI in
the Dallas heart study. The triglyceride increasing effect of the variant is amplified by increasing obesity (Pinteraction = 4 × 10−5). Data are presented
as median ± interquartile range. (b) Risk of cirrhosis by PNPLA3 genotype and BMI in the Copenhagen cohort. The risk‐increasing effect of the
variant is amplified by increasing obesity (Pinteraction = 0.026). Data are OR ± 95% confidence interval (CI). Reproduced with permission of Springer
Nature from [34].
41:  Clinical Genomics of NAFLD 513

Krawczyk et  al. [10] and other groups [40, 41] identified a [43]. Furthermore, Pirazzi et  al. [44] demonstrated that
possible association between higher fasting glucose levels and PNPLA3 has retinyl‐palmitate lipase activity in hepatic stel-
the PNPLA3 p.I148M mutation. These results contradict the late cells. The expression of human PNPLA3 is induced by
general paradigm that insulin resistance represents the main carbohydrates and fatty acids via the sterol response element
driver of common NAFLD. Indeed, a dissociation between the binding protein 1c (SREBP1c, Figure  41.3) as transcription
presence of fatty liver and insulin resistance appears to be pre- factor [43, 45–47].
sent among carriers of the PNPLA3 risk variant [36]. Specific Livers from NAFLD patients are characterized by the
circulating triacylglycerol signatures were observed in carriers increased number and size of lipid droplets within hepatocytes.
of the variant, which resembled those of obese subjects with Of note, PNPLA3 is predominantly localized in the endoplas-
NAFLD [42]. PNPLA3‐associated NAFLD is associated with mic reticulum (ER) and on the surface of lipid droplets [48]
a relative deficiency of triacylglycerols, supporting the idea (Figure  41.4). The rs738409 variant of PNPLA3 results in an
that the variant impedes intrahepatocellular lipolysis rather isoleucine to methionine substitution at amino acid residue 148.
than stimulating lipid synthesis. NAFLD in obese patients is Structural modeling indicates that this substitution occludes
associated with multiple changes in triacylglycerols, which access of substrates to the catalytic dyad, resulting in loss of
can be attributed to obesity and insulin resistance but not function [49, 50]. In carriers of the p.I148M mutation, the vari-
increased liver fat content per se. ant PNPLA3 evades ubiquitylation and proteosomal degrada-
tion but accumulates on lipid droplets, which increase in number
and median size and display impaired triglyceride mobilization
Functional analyses of the PNPLA3 variant
[49, 51–53].
The close similarity to adipose triglyceride lipase and the Studies performed in transgenic mice overexpressing variant
presence of typical structural motifs (α–β–α sandwich struc- PNPLA3148M but not wild‐type PNPLA3 in liver demonstrated
ture, consensus serine lipase GXSXG motif, catalytic dyad that these animals develop steatosis due to triacylglycerol accu-
S‐D) indicate a lipase function of PNPLA3 [7]. Alternatively, mulation as well as several alterations of hepatic lipid metabo-
lysophosphatidic acyltransferase activity has been proposed lism, including increased synthesis of fatty acids and

Figure 41.4  (a) PNPLA3 is induced with increased free fatty acid availability during hyperinsulinemia and expressed in the ER and on the surface
of lipid droplets. It is rapidly ubiquitylated and degraded upon fasting. (b) PNPLA3 p.I148M on lipid droplets escapes degradation and its accumu-
lation favors triglyceride accumulation and steatohepatitis. (c) Reduced expression of PNPLA3 p.I148M due to the copresence of the p.E434K vari-
ant reduces liver damage. Wild‐type PNPLA3 (p.148I) with intact protective enzymatic activity is illustrated as green curves, and variant PNPLA3
p.148M causing lipid droplet formation as red curves. The number is proportional to protein levels. Red arrows indicate damaging pathways, green
arrows protective pathways, and dashed lines denote suppressed pathways. Abbreviations: TAG, triglyceride; Ub, ubiquitin. Reproduced with per-
mission of John Wiley and Sons from [48].
514 THE LIVER:  TM6SF2 AS SECOND NAFLD RISK GENE

triacylglycerol, impaired hydrolysis of triacylglycerol, and TM6SF2 AS SECOND NAFLD RISK GENE
depletion of triacylglycerol long‐chain polyunsaturated fatty
acids [54]. To determine whether the mutant p.148M allele A GWAS with a denser panel of exonic polymorphisms in three
causes fat accumulation in the liver when expressed at physio- independent populations (N > 80 000) identified the TM6SF2
logical levels, Pnpla3148M knockin gene were generated, which variant p.E167K (c.449C>T, rs58542926) to confer susceptibil-
displayed normal levels of hepatic fat on chow diet, but when ity to NAFLD in addition to PNPLA3 [23]. The TM6SF2 variant
challenged with a high‐sucrose diet the liver fat content is markedly less frequent than the PNPLA3 polymorphism in all
increased two‐ to threefold compared to wild‐type controls ethnic groups (3–7%). TM6SF2 encodes a protein of 351 amino
without changes in glucose homeostasis. The increased liver fat acids with 7–10 predicted transmembrane domains. TM6SF2 is
in the knockin mice was accompanied by a fortyfold increase in localized in the ER and the Golgi complex of hepatocytes and
PNPLA3 on hepatic lipid droplets with no increase in hepatic enterocytes (Figure  41.5), which synthesize apolipoprotein B‐
Pnpla3 mRNA [55]. containing lipoproteins [52]. In contrast to PNPLA3, the expres-
sion of TM6SF2 is not altered by dietary challenges [52]. The
Genotype‐guided treatment of NAFLD polymorphism is associated with higher liver fat content but
lower serum levels of total and low‐density lipoprotein (LDL)
The in vitro and in vivo studies indicate that suppression of cholesterol, and triglycerides. The TM6SF2 p.E167K variant
mutant PNPLA3 would have beneficial effects in NAFLD does not exert its effect on hepatic fat content or circulating lipid
and represent a novel therapeutic target for NAFLD. Until profiles by causing hepatic insulin resistance [62]. In patients
then, lifestyle modification represents the cornerstone of with histologically proven NAFLD, an association of this vari-
NAFLD treatment. Although to date large prospective stud- ant with the degree of steatosis, necroinflammation, ballooning,
ies concerning long‐term effects of the common PNPLA3 and advanced fibrosis was observed, after adjustment for age,
p.I148M variant on liver phenotypes are lacking, the first sex, BMI, and diabetes [23, 63] (Table 41.2). Accordingly, the
pilot studies indicate that weight loss has positive effects in TM6SF2 variant was independently associated with advanced
carriers of the risk allele [56, 57]. To explore whether the fibrosis, when assessed noninvasively by transient elastography
PNPLA3 variant modulates the effects of weight loss on liver in 890 individuals [64].
fat and insulin sensitivity, eight homozygous carriers were TM6SF2 protein expression was decreased markedly in
placed on a hypocaloric low‐carbohydrate diet for six days liver  of NAFLD patients, compared to controls [65]. siRNA
[57]. Liver fat content (measured by PDFF), whole‐body knockdown of Tm6sf2 in mice decreases very‐low‐density
insulin sensitivity of glucose metabolism (euglycemic clamp ­lipoprotein (VLDL) secretion and increases cellular triglycer-
technique), and lipolysis ([2H5]glycerol infusion) were meas- ide concentration and lipid droplet content, whereas overex-
ured before and after the diet. At baseline, fasting serum pression reduces liver cell steatosis [66, 67]. Chronic
insulin and C‐peptide concentrations were lower in mutation inactivation of Tm6sf2 in mice is associated with hepatic stea-
carriers. However after a mean weight loss of 3.1 kg, liver fat tosis and inflammation as well as decreased plasma levels of
decreased by 45% in patients with PNPLA3148M but by 18% total and LDL cholesterol, thus recapitulating the phenotypes
only in controls. The PNPLA3 variant also modulated the observed in humans [52, 68]. These observations indicate that
changes in metabolic profile and intrahepatic triglycerides TM6SF2 activity is required for normal VLDL assembly and
(IHTG, measured by proton magnetic resonance spectros- that carriers of the TM6SF2 p.E167K variant have fatty liver as
copy) in 154 NAFLD patients [58]. The presence of the a result of reduced VLDL lipidation (Figure 41.5). As a result,
mutation and BMI were independently associated with these patients display lower circulating lipids and reduced risk
greater reduction in IHTG (genotype II: 3.7 ± 5.2%, IM: 6.5 of developing carotid plaques and cardiovascular events [63].
± 3.6%, MM: 11.3 ± 8.8%). Although the PNPLA3 risk allele In addition, the effect of insulin on glucose production and
confers a higher risk of NAFLD, these patients are more sen- lipolysis is better in patients carrying the TM6SF2 p.E167K
sitive to the beneficial effects of lifestyle modification. In variant. Hence, they display a distinct subtype of NAFLD,
the  same line, obese patients carrying the prosteatotic which is characterized by preserved insulin sensitivity with
PNPLA3148M allele lose more weight and liver fat one year regard to lipolysis, hepatic glucose production, and lack of
after bariatric surgery, as compared to carriers of PNPLA3 hypertriglyceridemia despite clearly increased hepatic fat
wild‐type alleles [59]. The PNPLA3 genotype and the initial ­content [62].
grade of steatosis were independent predictors of improve- The dual and opposite role of the TM6SF2 variant in pro-
ment of steatosis after surgery. tecting against cardiovascular disease and conferring risk for
In the WELCOME trial, 103 patients with NAFLD were NAFLD was confirmed in a meta‐analysis of 10 studies.
randomized to ω3 fatty acids or placebo for 15–18 months in Pooled estimates of random effects in more than 100 000 indi-
a randomized controlled trial. Adjusting for baseline meas- viduals showed that homozygous or heterozygous carriers of
urement and covariates, the PNPLA3 p.I148M variant was the minor T allele are protected from cardiovascular disease,
independently associated with percentage of docosahexae- showing lower levels of total and LDL cholesterol as well as
noic acid enrichment and end of study liver fat percentage, but triglycerides, while hepatic lipid fat content is about 2% higher
did not influence the change in serum triglyceride concentra- [69]. Accordingly, there is a moderate overall effect on the risk
tions [60]. The PNPLA3 p.I148M variant has also been of NAFLD (OR 2.1). Hence, the variant confers protection
reported to reduce the beneficial effects of statins on steato- against cardiovascular disease at the expense of an increased
hepatitis [61]. risk of NAFLD.
41:  Clinical Genomics of NAFLD 515

Figure 41.5  Role of TM6SF2 in VLDL lipidation. VLDL synthesis is initiated in the ER with the co‐translational addition of phospholipids to
apolipoprotein (Apo) B. The addition of triglycerides to the particle also begins in the ER, a process that requires MTTP. The partially lipidated
VLDL particle is packaged into vesicles and exported to the Golgi, where they appear to undergo further bulk phase lipidation. TM6SF2 promotes
this bulk phase lipidation, either by transporting neutral lipids from lipid droplets to the particle by transferring lipid to MTTP (➀), to neutral lipid
droplets in the ER lumen (➁), or directly to the nascent VLDL particle (➂); alternatively, TM6SF2 could participate in the transfer of lipid to the
particle en route to or within the Golgi complex (➃ and ➄). Abbreviations: ER, endoplasmic reticulum; LD, lipid droplet; MTTP, microsomal tri-
glyceride transfer protein; PL, phospholipid; PLTP, phospholipid transfer protein; TG, triglycerides; VLDL, very‐low‐density lipoprotein.
Reproduced with permission of Smagris from [52].

Interestingly, the PNPLA3 and TM6SF2 variants are not only MBOAT7: THE THIRD MAN
associated with clinical phenotypes but also with health services
utilization in the general population [70]. In 3759 participants After a GWAS had reported that the rs641738 C>T variant linked
from the study of health in Pomerania (SHIP), homozygous carri- to the 3’ untranslated region of the gene encoding membrane‐
ers of the PNPLA3 risk allele had an increased odds of hospitali- bound O‐acyltransferase domain‐containing 7 gene (MBOAT7,
zation (OR 1.5) as compared to major allele homozygous subjects, aka lysophosphatidylinositol‐acyltransferase 1, LPIAT1) increases
and carriers of the TM6SF2 risk allele had higher outpatient utili- the risk for alcoholic cirrhosis [73], 2736 participants from the
zation (+68%) and inpatient days than major allele homozygous Dallas heart study who had undergone proton magnetic resonance
subjects. These findings highlight the strong genetic effects spectroscopy to measure IHTG and 1149 European individuals
that can even be detected in health economic analysis. Recently from the liver biopsy cross‐sectional cohort were genotyped for
the PNPLA3 p.I148M polymorphism was studied in four large rs641738 [74]. This variant, which encodes p.G17E in the trans-
cohorts with extensive health information (23andMe, UK membrane channel‐like 4 (TMC4), is associated with suppression
biobank, FINRISK, CHOP) for association with 1683 binary end- of MBOAT7 at mRNA and protein levels. It was also associated
points in up to 700 000 individuals [71]. This study design has with increased hepatic fat content, more severe liver damage, and
been termed phenome‐wide association study (PheWAS), which increased stage of fibrosis compared with subjects without the
is an unbiased approach to test for associations between genetic variant. MBOAT7 transfers polyunsaturated fatty acids (PUFA)
variants and a wide range of phenotypes in large population such as arachidonic acid to lysophospholipids. The MBOAT7
cohorts [72]. Of note, the PNPLA3 variant, which predisposes to rs641738 risk allele was associated with altered plasma phosphati-
fatty liver disease, was associated with diabetes and lower serum dylinositol (PI) species, consistent with decreased MBOAT7 func-
cholesterol levels, and an inverse association was also detected tion. Metabolic profiling indicates changes of PI side‐chain
for gallstones, acne, and gout. Beyond that, the analysis also indi- remodeling, in particular a lack of transfer of arachidonoyl‐CoA to
cated that carriers of the risk allele are prone to develop drug‐ lyso‐PI (Lands cycle) [75]. Interestingly, PNPLA3 promotes the
induced liver injury (OR 1.5) when treated with non‐steroidal transfer of PUFA, especially arachidonic acid, from triglycerides
anti‐inflammatory drugs (NSAID) such as ibuprofen or aspirin. to phospholipids in hepatic lipid droplets [76]. These findings
These associations were replicated in the UK biobank cohort and together indicate a potential role of arachidonic acid incorporation
remained when adjusting for elevated liver tests [71]. in phospholipids during NAFLD pathogenesis.
516 THE LIVER:  PROTECTIVE GENE VARIATION

ADDITIONAL RISK GENES [83]. In line with these changes, the expression of key enzymes
in fatty acid synthesis, such as fatty acid synthase, acetyl‐CoA
GWAS and their meta‐analyses have identified additional vari- carboxylase 1, and stearoyl‐CoA desaturase, is increased in liv-
ants that might be associated with NAFLD. For example, ers from Hsd17b3 knockout mice, while glucose tolerance does
Speliotes and coworkers [17, 77] reported that variants associ- not differ [83].
ated at genome‐wide significant levels in or near the glucoki-
nase regulatory protein (GCKR), lysophospholipase‐like 1 Studies in pediatric cohorts
(LYPLAL1) and protein phosphatase 1‐regulatory subunit 3b
(PPP1R3B) might be associated with liver fat content and/or NAFLD is becoming an emerging health problem in pediatric
histopathological NAFLD phenotypes. The common missense populations. In line with the studies in adult patients, an associa-
loss‐of‐function GCKR mutation p.P446L (c.1337T>C, tion between fatty liver and the PNPLA3 mutation p.I148M was
rs1260326) reduces the ability of GCKR to inhibit glucokinase observed in children with NAFLD, too. In a study by Valenti
in response to fructose‐6‐phosphate, thereby increasing glu- et al. [20], the PNPLA3 variant was associated with the steatosis
cokinase activity and hepatic glucose uptake. The unrestricted grade in 149 children with biopsy‐proven NAFLD. In this cohort,
hepatic glycolysis reduces fasting glucose and insulin levels but all 23 children who were homozygous carriers of the PNPLA3
increases the production of malonyl‐CoA, which in turn favors mutation were diagnosed with NASH [20]. The association
hepatic fat accumulation by serving as a substrate for de novo between the PNPLA3 mutation and pediatric NAFLD has also
lipogenesis and by disruption of mitochondrial fatty acid β‐oxi- been detected in African American [84] and Chinese children
dation [78, 79]. The variants at these loci exhibit distinct pat- [85]. Interestingly, children carrying the PNPLA3 risk allele
terns of association with serum lipids as well as glycemic and seem to be predisposed to the early development of NAFLD
anthropometric traits, which also differ with ethnicity. Again [21]. The analysis of 6–12‐year‐old Mexican children showed
NAFLD‐associated variants are not uniformly associated with that already at this age the PNPLA3 mutation is associated with
abnormalities in serum lipids or glycemic and anthropometric increased serum ALT activities [86], and we detected the same
traits, suggesting genetic heterogeneity in the pathways influ- association in a cohort of German children aged 5–9 years [87].
encing these traits. The TM6SF2 variant p.E167K was also shown to be associated
with hepatic steatosis and increased aminotransferase activities
but lower serum cholesterol and triglyceride concentrations in
obese children [88, 89]. In multiethnic cohorts of obese children
PROTECTIVE GENE VARIATION and adolescents [90, 91], the MBOAT7 risk allele was associated
with hepatic fat content (as determined by magnetic resonance
Abul‐Husn et  al. [80] reported that a truncated variant of imaging), fasting insulin and plasma glucose levels, and reduced
hydroxysteroid 17β dehydrogenase 13 (HSD17B13) is associ- whole‐body insulin sensitivity, independent of age, sex, and BMI
ated with a reduced risk of chronic liver disease and of progres- z‐score. Moreover, these studies detected joint effects among
sion from steatosis to steatohepatitis. It has to be noted that the TM6SF2 p.E167K, PNPLA3 p.I148M, and the MBOAT7 rs626283
term “protection” represents the flipside of “susceptibility”, and GCKR rs1260326 polymorphisms in determining IHTG. The
hence it depends on the population frequencies of the alleles, four polymorphisms combined explain about 20% of the hepatic
with the minor allele defining the direction of the effect (pre- fat fraction in Caucasian obese children.
disposition vs. protection). Using exome sequence data and A practical consideration is that physical activity and weight
electronic health records from 46 544 participants in the reduction might substantially improve the liver status in chil-
DiscovEHR human genetics study, gene variants that are asso- dren carrying the risk variants and therefore rescue their delete-
ciated with serum activities of aminotransferases were identi- rious liver phenotypes [56]. This possibility points to the need of
fied. Replicated variants were evaluated for association with early detection of children carrying risk mutations who may
clinical diagnoses of chronic liver disease in DiscovEHR study require more careful follow‐up and tailored therapies aiming to
participants as well as two independent cohorts (N = 37 173) intercept NAFLD progression.
and with histopathological severity of liver disease in 2391
human liver biopsies. In homozygous and heterozygous carri-
Combined effects of gene variants
ers of the variant, the risks of NAFLD and NASH‐cirrhosis
were reduced by 30% and 17% and 49% and 26%, respectively. Mancina et  al. [74] analyzed the combined genetic effects of
Of note, the HSD17B13 variant ameliorated liver injury associ- PNPLA3, TM6SF2, and MBOAT7 effects on NAFLD histopathol-
ated with the PNPLA3 p.I148M risk allele. In a second detailed ogy (Table 41.3). They did not observe an interaction, instead the
histopathological study of 356 patients with biopsy‐proven dis- three gene variants appear to act in an additive fashion, with a
ease the variant protected against lobular inflammation, bal- stepwise increase in mean hepatic fat content per each additional
looning degeneration, and fibrosis [81]. risk allele (Figure 41.6). The combined effects of the PNPLA3,
Hydroxysteroid17β dehydrogenases (HSD17B) form an TM6SF2, and MBOAT7 variants on NAFLD severity were stud-
enzyme family characterized by their ability to catalyze reac- ied in a German multicenter biopsy‐based study that recruited
tions in steroid and lipid metabolism. Recently Ma et al. [82] 515 patients with NAFLD [92]. In the multivariate model, the
reported that HSD17B13 is a hepatic retinol dehydrogenase that PNPLA3 and TM6SF2 variants were associated with steatosis,
is targeted to lipid droplets. Mice deficient in Hsd17b13 develop and fibrosis stages were affected by the PNPLA3 and MBOAT7
hepatic steatosis, while serum steroid concentrations are normal polymorphism. Of note, a significant increase of serum AST
41:  Clinical Genomics of NAFLD 517

Table 41.3  Population attributable fractionsa of NAFLD genes


Steatosis Fibrosis
(confidence intervals) NASH F2–F4
PNPLA3 23% (11–36) 19% (12–26) 18% (10–26)
TM6SF2 4% (0–14) 4% (1–8) 3% (0–7)
MBOAT7 15% (3–28) 7% (0–14) 11% (2–23)
a
 The contribution of a risk factor to a disease or a death is quantified using PAF. PAF is the
proportional reduction in population disease or mortality that would occur if exposure to a risk
factor were reduced to an alternative ideal exposure scenario.

(b) P < 0.0001


500
300
100
100

AST [U/I]
(a)
20 50
P-trend < .0001
Mean hepatic TG%

15 0

le

s
le

le

le

le

le
le
le

le

le

le

le
al
al

al

al

al

al
k
k

ris

k
ris

ris

ris

ris

ris
10

1
0

5
Number of risk alleles

5 P = 0.08
400
300
200
0 150
0 1 2 3 4 5
ALT [U/I]

n= 627 1066 747 248 41 6 100

Number of risk alleles


50

s
s
s

le

s
le
le
le

le

le
n = 515
le

le
le
le

le

le
al

al
al
al

al

al
k

k
k
k

ris

k
ris
ris
ris

ris

ris
1

4
3
0

5
Number of risk alleles

Figure 41.6  Combined genetic effects in NAFLD. (a) Association between number of PNPLA3, TM6SF2, and MBOAT7 risk alleles and hepatic
triglyceride (TG) content in the Dallas heart study. The graph shows mean hepatic TG content by the number of risk alleles (error bars: SEM). From
[74]. (b) Association between number of PNPLA3, TM6SF2, and MBOAT7 risk alleles and liver function tests in the NAFLD CSG study. The graphs
illustrate median aspartate aminotransferase (AST) and alanine aminotransferase (ALT) activities by the number of risk alleles (error bars: range).
Analyses were performed using trend tests. The following frequencies of carriers of risk alleles were detected: 0 risk alleles, N = 56; one risk allele,
N = 142; two risk alleles, N = 170; three risk alleles, N = 117; four risk alleles, N = 27; five risk alleles, N = 3. Reproduced with permission of
Elsevier from [92].

activities was associated with the increment of risk alleles of over hepatic steatosis. However, they provide insight into the
either of the genotypes, and similar trends existed for ALT and pathogenesis of hepatic steatosis and can be considered in rare
γ‐glutamyl transpeptidase levels. The same observation was made cases when the etiology is unclear. Examples include abetali-
in Korea: the PNPLA3 and TM6SF2 variants, but not the MBOAT7 poproteinemia (microsomal triglyceride transfer protein,
variant were associated with both NASH and significant fibrosis MTTP), hypobetalipoproteinemia (apolipoprotein B, APOB),
(≥ F2), even after adjustment for insulin resistance [93]. citrullinemia type 2 (solute carrier SLC25A13), Wilson disease
(ATPase ATP7B), neutral lipid storage disease (PNPLA2), and
cholesteryl ester storage disease (lysosomal acid lipase, LIPA).
Rare inherited mitochondriopathies have also been associated
MONOGENIC DISEASES CAUSING NAFLD with NASH and point to the role of impaired fatty acid oxida-
tion in the pathogenesis of NASH. Genetic testing is helpful
Monogenic diseases that are associated with NAFLD are rare where a rare monogenic inherited disease of lipid metabolism
and usually result in extrahepatic manifestations that dominate or mitochondrial function is suspected. More recently exome
518 THE LIVER:  REFERENCES

and whole genome sequencing is applied to identify novel gene 13. Yuan, X., Waterworth, D., Perry, J.R. et al. Population‐based genome‐wide
variants causing liver diseases, including severe types of association studies reveal six loci influencing plasma levels of liver enzymes.
Am J Hum Genet, 2008;83(4):520–8.
NAFLD. 14. Kawaguchi, T., Sumida, Y., Umemura, A. et al. Genetic polymorphisms of
the human PNPLA3 gene are strongly associated with severity of non‐alco-
holic fatty liver disease in Japanese. PLoS One, 2012;7(6), e38322.
15. Kitamoto, T., Kitamoto, A., Yoneda, M. et al. Genome‐wide scan revealed
PNPLA3‐ AND TM6SF2‐ASSOCIATED that polymorphisms in the PNPLA3, SAMM50, and PARVB genes are asso-
STEATOHEPATITIS AND FUTURE ciated with development and progression of nonalcoholic fatty liver disease
in Japan. Hum Genet, 2013;132(7):783–92.
DIRECTIONS 16. Kotronen, A., Johansson, L.E., Johansson, L.M. et al. A common variant in
PNPLA3, which encodes adiponutrin, is associated with liver fat content in
PNPLA3 (and TM6SF2 plus MBOAT7) genotyping may be used humans. Diabetologia, 2009;52(6):1056–60.
17. Speliotes, E.K., Yerges‐Armstrong, L.M., Wu, J. et al. Genome‐wide asso-
as novel non‐invasive indicator for an increased risk of progres- ciation analysis identifies variants associated with nonalcoholic fatty liver
sive fatty liver disease and could be included in the clinical deci- disease that have distinct effects on metabolic traits. PLoS Genet, 2011;7(3),
sion‐making in patients with NAFLD. Because in patients who e1001324.
carry the PNPLA3 risk variant increased lipid content and inflam- 18. Sookoian, S., Castano, G.O., Burgueno, A.L. et al. A nonsynonymous gene
variant in the adiponutrin gene is associated with nonalcoholic fatty liver
mation in liver can be driven predominantly by PNPLA3 together
disease severity. J Lipid Res, 2009;50(10):2111–6.
with environmental risk factors, the name PNPLA3‐NAFLD or 19. Valenti, L., Al‐Serri, A., Daly, A.K. et al. Homozygosity for the patatin‐like
PASH (i.e. PNPLA3‐associated steatohepatitis) might be used as phospholipase‐3/adiponutrin I148M polymorphism influences liver fibrosis
a novel gene‐based liver disease entity [94, 95]. PNPLA3‐ in patients with nonalcoholic fatty liver disease. Hepatology, 2010;51(4):
NAFLD/PASH represents an example of how to reclassify dis- 1209–17.
20. Pelusi, S., Cespiati, A., Rametta, R. et al. Prevalence and risk factors of sig-
ease according to molecular pathways and pathophysiological
nificant fibrosis in patients with nonalcoholic fatty liver without steatohepa-
changes in the era of personalized medicine [96]. Carriers of the titis. Clin Gastroenterol Hepatol, 2019;epub.
genetic risk factors could benefit from a more systematic, early, 21. Rotman, Y., Koh, C., Zmuda, J.M. et al. The association of genetic variability
and careful surveillance of complications of progressive fatty in patatin‐like phospholipase domain‐containing protein 3 (PNPLA3) with
liver disease, including HCC in the absence of cirrhosis [97]. histological severity of nonalcoholic fatty liver disease. Hepatology,
2010;52(3):894–903.
22. Xu, R., Tao, A., Zhang, S. et al. Association between patatin‐like phospholi-
pase domain containing 3 gene (PNPLA3) polymorphisms and nonalcoholic
fatty liver disease: a HuGE review and meta‐analysis. Sci Rep, 2015;5, 9284.
REFERENCES 23. Liu, Y.L., Reeves, H.L., Burt, A.D. et  al. TM6SF2 rs58542926 influences
hepatic fibrosis progression in patients with non‐alcoholic fatty liver disease.
  1. Chalasani, N., Younossi, Z., Lavine, J.E. et al. The diagnosis and management Nat Commun, 2014;5, 4309.
of nonalcoholic fatty liver disease: practice guidance from the American 24. Krawczyk, M., Grünhage, F., Zimmer, V. et al. Variant adiponutrin (PNPLA3)
Association for the Study of Liver Diseases. Hepatology, 2018;67(1):328–57. represents a common fibrosis risk gene: non‐invasive elastography‐based
  2. Romeo, S., Kozlitina, J., Xing, C. et al. Genetic variation in PNPLA3 confers study in chronic liver disease. J Hepatol, 2011;55(2):299–306.
susceptibility to nonalcoholic fatty liver disease. Nat Genet, 2008;40(12): 25. Kim, D. and Kim, W.R. Nonobese fatty liver disease. Clin Gastroenterol
1461–5. Hepatol, 2017;15(4):474–85.
  3. Petaja, E.M. and Yki‐Järvinen, H. Definitions of normal liver fat and the 26. Wei, J.L., Leung, J.C., Loong, T.C. et al. Prevalence and severity of nonalco-
association of insulin sensitivity with acquired and genetic NAFLD‐A sys- holic fatty liver disease in non‐obese patients: a population study using
tematic review. Int J Mol Sci, 2016;17(5). proton‐magnetic resonance spectroscopy. Am J Gastroenterol, 2015;110(9):
  4. Verhaegh, P., Bavalia, R., Winkens, B. et al. Noninvasive tests do not accurately 1306–14.
differentiate nonalcoholic steatohepatitis from simple steatosis: a systematic 27. Leung, J.C., Loong, T.C., Wei, J.L. et al. Histological severity and clinical
review and meta‐analysis. Clin Gastroenterol Hepatol, 2018;16(6):837–61. outcomes of nonalcoholic fatty liver disease in nonobese patients.
  5. Estes, C., Razavi, H., Loomba, R. et al. Modeling the epidemic of nonalco- Hepatology, 2017;65(1):54–64.
holic fatty liver disease demonstrates an exponential increase in burden of 28. Krawczyk, M., Bantel, H., Rau, M. et  al. Could inherited predisposition
disease. Hepatology, 2018;67(1):123–33. drive non‐obese fatty liver disease? Results from German tertiary referral
  6. Loomba, R., Schork, N., Chen, C.H. et al. Heritability of hepatic fibrosis and centers. J Hum Genet, 2018;63(5):621–6.
steatosis based on a prospective twin study. Gastroenterology, 2015;149(7): 29. Fracanzani, A.L., Petta, S., Lombardi, R. et  al. Liver and cardiovascular
1784–93. damage in patients with lean nonalcoholic fatty liver disease, and association
  7. Zechner, R., Zimmermann, R., Eichmann, T.O. et al. Fat signals – lipases with visceral obesity. Clin Gastroenterol Hepatol, 2017;15(10):1604–11.
and lipolysis in lipid metabolism and signaling. Cell Metab, 30. Burza, M.A., Pirazzi, C., Maglio, C. et  al. PNPLA3 I148M (rs738409)
2012;15(3):279–91. genetic variant is associated with hepatocellular carcinoma in obese indi-
  8. Huang, Y., He, S., Li, J.Z. et al. A feed‐forward loop amplifies nutritional viduals. Dig Liver Dis, 2012;44(12):1037–41.
regulation of PNPLA3. Proc Natl Acad Sci USA, 2010;107(17):7892–7. 31. Liu, Y.L., Patman, G.L., Leathart, J.B. et  al. Carriage of the PNPLA3
  9. Sookoian, S. and Pirola, C.J. Meta‐analysis of the influence of I148M variant rs738409 C >G polymorphism confers an increased risk of non‐alcoholic
of patatin‐like phospholipase domain containing 3 gene (PNPLA3) on the fatty liver disease associated hepatocellular carcinoma. J Hepatol,
susceptibility and histological severity of nonalcoholic fatty liver disease. 2014;61(1):75–81.
Hepatology, 2011;53(6):1883–94. 32. Anstee, Q.M., Seth, D., Day, C.P. Genetic factors that affect risk of
10. Krawczyk, M., Grünhage, F., Mahler, M. et  al. The common adiponutrin alcoholic  and nonalcoholic fatty liver disease. Gastroenterology,
­
variant p.I148M does not confer gallstone risk but affects fasting glucose and 2016;150(8):1728–44.
triglyceride levels. J Physiol Pharmacol, 2011;62(3):369–75. 33. Hassan, M.M., Kaseb, A., Etzel, C.J. et al. Genetic variation in the PNPLA3
11. Diehl, A.M. and Day, C. Nonalcoholic steatohepatitis. N Engl J Med, gene and hepatocellular carcinoma in USA: risk and prognosis prediction.
2018;378(8):781. Mol Carcinog, 2013;52(1), E139–47.
12. Chambers, J.C., Zhang, W., Sehmi, J. et al. Genome‐wide association study 34. Stender, S., Kozlitina, J., Nordestgaard, B.G. et al. Adiposity amplifies the
identifies loci influencing concentrations of liver enzymes in plasma. Nat genetic risk of fatty liver disease conferred by multiple loci. Nat Genet,
Genet, 2011;43(11):1131–8. 2017;49(6):842–7.
41:  Clinical Genomics of NAFLD 519

35. Palasciano, G., Moschetta, A., Palmieri, V.O. et al. Non‐alcoholic fatty liver 60. Scorletti, E., West, A.L., Bhatia, L. et al. Treating liver fat and serum triglyc-
disease in the metabolic syndrome. Curr Pharm Des, 2007;13(21):2193–8. eride levels in NAFLD, effects of PNPLA3 and TM6SF2 genotypes: Results
36. Kantartzis, K., Peter, A., Machicao, F. et al. Dissociation between fatty liver from the WELCOME trial. J Hepatol, 2015;63(6):1476–83.
and insulin resistance in humans carrying a variant of the patatin‐like phos- 61. Dongiovanni, P., Petta, S., Mannisto, V. et al. Statin use and non‐alcoholic
pholipase 3 gene. Diabetes, 2009;58(11):2616–23. steatohepatitis in at risk individuals. J Hepatol, 2015;63(3):705–12.
37. Speliotes, E.K., Butler, J.L., Palmer, C.D. et al. PNPLA3 variants specifi- 62. Zhou, Y., Llaurado, G., Oresic, M. et al. Circulating triacylglycerol signa-
cally confer increased risk for histologic nonalcoholic fatty liver disease but tures and insulin sensitivity in NAFLD associated with the E167K variant in
not metabolic disease. Hepatology, 2010;52(3):904–12. TM6SF2. J Hepatol, 2015;62(3):657–63.
38. Krarup, N.T., Grarup, N., Banasik, K. et al. The PNPLA3 rs738409 G‐allele 63. Dongiovanni, P., Petta, S., Maglio, C. et al. Transmembrane 6 superfamily
associates with reduced fasting serum triglyceride and serum cholesterol in member 2 gene variant disentangles nonalcoholic steatohepatitis from car-
Danes with impaired glucose regulation. PLoS One, 2012;7(7), e40376. diovascular disease. Hepatology, 2015;61(2):506–14.
39. Stojkovic, I.A., Ericson, U., Rukh, G. et al. The PNPLA3 Ile148Met inter- 64. Petta, S., Di Marco, V., Pipitone, R.M. et al. Prevalence and severity of non-
acts with overweight and dietary intakes on fasting triglyceride levels. Genes alcoholic fatty liver disease by transient elastography: genetic and metabolic
Nutr, 2014;9(2):388. risk factors in a general population. Liver Int, 2018;38(11):2060–8.
40. Palmer, C.N., Maglio, C., Pirazzi, C. et al. Paradoxical lower serum triglyc- 65. Sookoian, S., Castano, G.O., Scian, R. et al. Genetic variation in transmem-
eride levels and higher type 2 diabetes mellitus susceptibility in obese indi- brane 6 superfamily member 2 and the risk of nonalcoholic fatty liver disease
viduals with the PNPLA3 148M variant. PLoS One, 2012;7(6), e39362. and histological disease severity. Hepatology, 2015;61(2):515–25.
41. Rembeck, K., Maglio, C., Lagging, M. et al. PNPLA 3 I148M genetic vari- 66. Kozlitina, J., Smagris, E., Stender, S. et al. Exome‐wide association study
ant associates with insulin resistance and baseline viral load in HCV geno- identifies a TM6SF2 variant that confers susceptibility to nonalcoholic fatty
type 2 but not in genotype 3 infection. BMC Med Genet, 2012;13, 82. liver disease. Nat Genet, 2014;46(4):352–6.
42. Hyysalo, J., Gopalacharyulu, P., Bian, H. et  al. Circulating triacylglycerol 67. Mahdessian, H., Taxiarchis, A., Popov, S. et al. TM6SF2 is a regulator of
signatures in nonalcoholic fatty liver disease associated with the I148M vari- liver fat metabolism influencing triglyceride secretion and hepatic lipid
ant in PNPLA3 and with obesity. Diabetes, 2014;63(1):312–22. droplet content. Proc Natl Acad Sci USA, 2014;111(24):8913–8.
43. Kumari, M., Schoiswohl, G., Chitraju, C. et al. Adiponutrin functions as a 68. Fan, Y., Lu, H., Guo, Y. et al. Hepatic transmembrane 6 superfamily member
nutritionally regulated lysophosphatidic acid acyltransferase. Cell Metab, 2 regulates cholesterol metabolism in mice. Gastroenterology,
2012;15(5):691–702. 2016;150(5):1208–18.
44. Pirazzi, C., Valenti, L., Motta, B.M. et  al. PNPLA3 has retinyl‐palmitate 69. Pirola, C.J. and Sookoian, S. The dual and opposite role of the TM6SF2‐
lipase activity in human hepatic stellate cells. Hum Mol Genet, rs58542926 variant in protecting against cardiovascular disease and confer-
2014;23(15):4077–85. ring risk for nonalcoholic fatty liver: a meta‐analysis. Hepatology,
45. Huang, Y., Cohen, J.C., Hobbs, H.H. Expression and characterization of a 2015;62(6):1742–56.
PNPLA3 protein isoform (I148M) associated with nonalcoholic fatty liver 70. Kopp, J., Flessa, S., Lieb, W. et al. Association of PNPLA3 rs738409 and
disease. J Biol Chem, 2011;286(43):37085–93. TM6SF2 rs58542926 with health services utilization in a population‐based
46. Dubuquoy, C., Robichon, C., Lasnier, F. et al. Distinct regulation of adiponu- study. BMC Health Serv Res, 2016;16, 41.
trin/PNPLA3 gene expression by the transcription factors ChREBP and 71. Diogo, D., Tian, C., Franklin, C.S. et al. Phenome‐wide association studies
SREBP1c in mouse and human hepatocytes. J Hepatol, 2011;55(1):145–53. across large population cohorts support drug target validation. Nat Commun,
47. Perttila, J., Huaman‐Samanez, C., Caron, S. et al. PNPLA3 is regulated by 2018;9(1):4285.
glucose in human hepatocytes, and its I148M mutant slows down triglycer- 72. Bush, W.S., Oetjens, M.T., Crawford, D.C. Unravelling the human genome‐
ide hydrolysis. Am J Physiol Endocrinol Metab, 2012;302(9), E1063–9. phenome relationship using phenome‐wide association studies. Nat Rev
48. Valenti, L. and Dongiovanni, P. Mutant PNPLA3 I148M protein as pharma- Genet, 2016;17(3):129–45.
cological target for liver disease. Hepatology, 2017;66(4):1026–8. 73. Buch, S., Stickel, F., Trepo, E. et al. A genome‐wide association study con-
49. He, S., McPhaul, C., Li, J.Z. et al. A sequence variation (I148M) in PNPLA3 firms PNPLA3 and identifies TM6SF2 and MBOAT7 as risk loci for alcohol‐
associated with nonalcoholic fatty liver disease disrupts triglyceride hydroly- related cirrhosis. Nat Genet, 2015;47(12):1443–8.
sis. J Biol Chem, 2010;285(9):6706–15. 74. Mancina, R.M., Dongiovanni, P., Petta, S. et al. The MBOAT7‐TMC4 vari-
50. Xin, Y.N., Zhao, Y., Lin, Z.H. et  al. Molecular dynamics simulation of ant rs641738 increases risk of nonalcoholic fatty liver disease in individuals
PNPLA3 I148M polymorphism reveals reduced substrate access to the cata- of european descent. Gastroenterology, 2016;150(5):1219–30.
lytic cavity. Proteins, 2013;81(3):406–14. 75. Wang, B. and Tontonoz, P. Phospholipid remodeling in physiology and dis-
51. Chamoun, Z., Vacca, F., Parton, R.G. et al. PNPLA3/adiponutrin functions in ease. Annu Rev Physiol, 2018;81:165–88.
lipid droplet formation. Biol Cell, 2013;105(5):219–33. 76. Mitsche, M.A., Hobbs, H.H., and Cohen, J.C. Patatin‐like phospholipase
52. Smagris, E., Gilyard, S., BasuRay, S. et al. Inactivation of Tm6sf2, a gene domain‐containing protein 3 promotes transfer of essential fatty acids from
defective in fatty liver disease, impairs lipidation but not secretion of very triglycerides to phospholipids in hepatic lipid droplets. J Biol Chem,
low density lipoproteins. J Biol Chem, 2016;291(20):10659–76. 2018;293(18):6958–68.
53. BasuRay, S., Smagris, E., Cohen, J.C. et al. The PNPLA3 variant associated 77. Hernaez, R., McLean, J., Lazo, M. et al. Association between variants in or
with fatty liver disease (I148M) accumulates on lipid droplets by evading near PNPLA3, GCKR, and PPP1R3B with ultrasound‐defined steatosis
ubiquitylation. Hepatology, 2017;66(4):1111–24. based on data from the third National Health and Nutrition Examination
54. Li, J.Z., Huang, Y., Karaman, R. et al. Chronic overexpression of PNPLA3I148M Survey. Clin Gastroenterol Hepatol, 2013;11(9):1183–90.
in mouse liver causes hepatic steatosis. J Clin Invest, 2012;122(11):4130–44. 78. Beer, N.L., Tribble, N.D., McCulloch L.J. et al. The P446L variant in GCKR
55. Smagris, E., BasuRay, S., Li, J. et al. Pnpla3I148M knockin mice accumu- associated with fasting plasma glucose and triglyceride levels exerts its
late PNPLA3 on lipid droplets and develop hepatic steatosis. Hepatology, effect  through increased glucokinase activity in liver. Hum Mol Genet,
2015;61(1):108–18. 2009;18(21):4081–8.
56. Marzuillo, P., Grandone, A., Perrone, L. et al. Weight loss allows the dissec- 79. Eslam, M., Valenti, L., and Romeo, S. Genetics and epigenetics of NAFLD
tion of the interaction between abdominal fat and PNPLA3 (adiponutrin) in and NASH: clinical impact. J Hepatol, 2018;68(2):268–79.
the liver damage of obese children. J Hepatol, 2013;59(5):1143–4. 80. Abul‐Husn, N.S., Cheng, X., Li, A.H. et al. A protein‐truncating HSD17B13
57. Sevastianova, K., Kotronen, A., Gastaldelli, A. et  al. Genetic variation in variant and protection from chronic liver disease. N Engl J Med,
PNPLA3 (adiponutrin) confers sensitivity to weight loss‐induced decrease in 2018;378(12):1096–106.
liver fat in humans. Am J Clin Nutr, 2011;94(1):104–11. 81. Pirola, C.J., Garaycoechea, M., Flichman, D. et al. Splice variant rs72613567
58. Shen, J., Wong, G.L., Chan, H.L. et al. PNPLA3 gene polymorphism and prevents worst histologic outcomes in patients with nonalcoholic fatty liver
response to lifestyle modification in patients with nonalcoholic fatty liver disease. J Lipid Res, 2019;60(1):176–85.
disease. J Gastroenterol Hepatol, 2015;30(1):139–46. 82. Ma, Y., Belyaeva, O.V., Brown, P.M. et  al. 17‐Beta hydroxysteroid
59. Krawczyk, M., Jimenez‐Aguero, R., Alustiza, J.M. et al. PNPLA3 p.I148M dehydrogenase 13 is a hepatic retinol dehydrogenase associated with
­
variant is associated with greater reduction of liver fat content after bariatric histological features of nonalcoholic fatty liver disease. Hepatology,
­
surgery. Surg Obes Relat Dis, 2016;12(10):1838–46. 2019;69(4):1504–19.
520 THE LIVER:  REFERENCES

83. Adam, M., Heikela, H., Sobolewski, C. et al. Hydroxysteroid (17beta) dehy-   94. Krawczyk, M., Portincasa, P., Lammert, F. PNPLA3‐associated steatohepa-
drogenase 13 deficiency triggers hepatic steatosis and inflammation in mice. titis: toward a gene‐based classification of fatty liver disease. Semin Liver
FASEB J, 2018;32(6):3434–47. Dis, 2013;33(4):369–79.
84. Santoro, N., Kursawe, R., D’Adamo E. et al. A common variant in the pata-   95. Yki‐Järvinen, H. Non‐alcoholic fatty liver disease as a cause and a consequence
tin‐like phospholipase 3 gene (PNPLA3) is associated with fatty liver disease of metabolic syndrome. Lancet Diabetes Endocrinol, 2014;2(11):901–10.
in obese children and adolescents. Hepatology, 2010;52(4):1281–90.   96. Mirnezami, R., Nicholson, J., Darzi, A. Preparing for precision medicine. N
85. Lin, Y.C., Chang, P.F., Hu, F.C. et al. A common variant in the PNPLA3 gene Engl J Med, 2012;366(6):489–91.
is a risk factor for non‐alcoholic fatty liver disease in obese Taiwanese chil-   97. Valenti, L., Dongiovanni, P., Ginanni Corradini, S. et al. PNPLA3 I148M
dren. J Pediatr, 2011;158(5):740–4. variant and hepatocellular carcinoma: a common genetic variant for a rare
86. Larrieta‐Carrasco, E., Leon‐Mimila, P., Villarreal‐Molina, T. et  al. disease. Dig Liver Dis, 2013;45(8):619–24.
Association of the I148M/PNPLA3 variant with elevated alanine transami-   98. Tian, C., Stokowski, R.P., Kershenobich, D. et  al. Variant in PNPLA3 is
nase levels in normal‐weight and overweight/obese Mexican children. Gene, associated with alcoholic liver disease. Nat Genet, 2010;42(1):21–3.
2013;520(2):185–8.   99. Stickel, F., Buch, S., Lau, K. et al. Genetic variation in the PNPLA3 gene is
87. Krawczyk, M., Liebe, R., Maier, I.B., et  al. The frequent adiponutrin associated with alcoholic liver injury in caucasians. Hepatology,
(PNPLA3) variant p.Ile148Met is associated with early liver injury: 2011;53(1):86–95.
analysis of a German pediatric cohort. Gastroenterol Res Pract, 100. Vigano, M., Valenti, L., Lampertico, P. et  al. Patatin‐like phospholipase
2015;205079. domain‐containing 3 I148M affects liver steatosis in patients with chronic
88. Grandone, A., Cozzolino, D., Marzuillo, P. et al. TM6SF2 Glu167Lys poly- hepatitis B. Hepatology, 2013;58(4):1245–52.
morphism is associated with low levels of LDL‐cholesterol and increased 101. Cai, T., Dufour, J.F., Müllhaupt, B. et  al. Viral genotype‐specific role of
liver injury in obese children. Pediatr Obes; 2016;11(2):115–9. PNPLA3, PPARG, MTTP, and IL28B in hepatitis C virus‐associated stea-
89. Goffredo, M., Caprio, S., Feldstein, A.E. et al. Role of TM6SF2 rs58542926 tosis. J Hepatol, 2011;55(3):529–35.
in the pathogenesis of nonalcoholic pediatric fatty liver disease: a multieth- 102. Müller, T., Buch, S., Berg, T. et al. Distinct, alcohol‐modulated effects of
nic study. Hepatology, 2016;63(1):117–25. PNPLA3 genotype on progression of chronic hepatitis C. J Hepatol,
90. Di Sessa, A., Umano, G.R., Cirillo, G. et al. The membrane‐bound O‐acyl- 2011;55(3):732–3.
transferase7 rs641738 variant in pediatric nonalcoholic fatty liver fisease. J 103. Valenti, L. and Fargion, S. Patatin‐like phospholipase domain containing‐3
Pediatr Gastroenterol Nutr, 2018;67(1):69–74. Ile148Met and fibrosis progression after liver transplantation. Hepatology,
91. Umano, G.R., Caprio, S., Di Sessa, A. et  al. The rs626283 variant in the 2011;54(4):1484.
MBOAT7 gene is associated with insulin resistance and fatty liver in 104. Nischalke, H.D., Berger, C., Luda, C. et al. The PNPLA3 rs738409 148M/M
Caucasian obese youth. Am J Gastroenterol, 2018;113(3):376–83. genotype is a risk factor for liver cancer in alcoholic cirrhosis but shows no
92. Krawczyk, M., Rau, M., Schattenberg, J.M. et al. Combined effects of the or weak association in hepatitis C cirrhosis. PLoS One, 2011;6(11):e27087.
PNPLA3 rs738409, TM6SF2 rs58542926, and MBOAT7 rs641738 variants 105. Younossi, Z., Anstee, Q.M., Marietti, M. et al. Global burden of NAFLD
on NAFLD severity: a multicenter biopsy‐based study. J Lipid Res, and NASH: trends, predictions, risk factors and prevention. Nat Rev
2017;58(1):247–55. Gastroenterol Hepatol, 2018;15(1):11–20.
93. Koo, B.K., Joo, S.K., Kim, D. et  al. Additive effects of PNPLA3 and 106. Dubuquoy, C., Burnol, A.F., Moldes, M. PNPLA3, a genetic marker of pro-
TM6SF2 on the histological severity of non‐alcoholic fatty liver disease. J gressive liver disease, still hiding its metabolic function? Clin Res Hepatol
Gastroenterol Hepatol, 2018;33(6):1277–85. Gastroenterol, 2013;37(1):30–5.
SECTION B:
LIVER GROWTH
AND REGENERATION
Stem Cell‐Fueled Maturational
42 Lineages in Hepatic and
Pancreatic Organogenesis
Wencheng Zhang1, Amanda Allen1, Eliane Wauthier1, Xianwen Yi2,
Homayoun Hani1, Praveen Sethupathy3,10, David Gerber2, Vincenzo
Cardinale5, Guido Carpino6,7, Juan Dominguez‐Bendala8,9, Giacomo
Lanzoni9, Domenico Alvaro5,8,*, Eugenio Gaudio7,*, and Lola Reid1,4,*
1 
Department of Cell Biology and Physiology, University of North Carolina School of Medicine,
Chapel Hill, NC, USA
2 
Department of Surgery, University of North Carolina School of Medicine, Chapel Hill, NC, USA
3 
Department of Genetics, University of North Carolina School of Medicine, Chapel Hill, NC, USA
4 
Program in Molecular Biology and Biotechnology, University of North Carolina School of Medicine,
Chapel Hill, NC, USA
5 
Department of Medico‐Surgical Sciences and Biotechnologies, Sapienza University of Rome, Latina, Italy
6 
Department of Movement, Human and Health Sciences, Division of Health Sciences, University of Rome
“Foro Italico”, Rome, Italy
7 
Department of Anatomical, Histological, Forensic Medicine and Orthopedics Sciences, Sapienza University
of Rome, Rome, Italy
8 
Department of Medicine and Medical Specialties, Sapienza University of Rome, Rome, Italy
9 
Diabetes Research Institute, Miller School of Medicine, University of Miami, Miami, FL, USA
10 
Department of Biomedical Sciences, Cornell University, Ithaca, NY, USA

INTRODUCTION of the body depends heavily on liver, biliary tree, and pancreatic
functions, and failure in any of them results in rapid death.
Liver, biliary tree, and pancreas are mid‐gut endodermal organs The focus of this review is on current knowledge of the newly
central to handling glycogen and lipid metabolism, detoxifica- discovered network of multiple stem/progenitor cell populations
tion of xenobiotics, processing of nutrients for optimal utiliza- present primarily in the biliary tree and giving rise to maturational
tion, regulation of energy needs, and synthesis of diverse factors lineages of cells forming liver and pancreas throughout life. In addi-
ranging from coagulation proteins to carrier proteins (e.g. tion, the chapter summarizes information regarding key paracrine
alpha‐fetoprotein (AFP), albumin, transferrin) [1]. The integrity signals from lineage‐dependent epithelial–mesenchymal cell inter-
actions in the maturational processes and found useful for ex vivo
maintenance and differentiation of hepato/biliary and pancreatic
* L.M. Reid was the only one who wrote most of the chapter with the qualifier cells at particular lineage stages. In prior reviews and previous edi-
that the last segment, that on plasticity was written by Domenico Alvaro and Lola tions of this book are given more details on hepatic stem cells and
Reid and with input also from Eugenio Gaudio. Amanda Allen, an artist and hepatoblasts, on early studies on the biliary tree stem cells, and on
animator, did the figures. All of the other authors have done the experiments facets of extracellular matrix chemistry and biology, key require-
resulting in the discoveries that are summarized in the review. The senior authors
in the management of the experiments comprise G. Lanzoni and J. Dominguez‐
ments for ex vivo maintenance of the cells [1–9].
Bendala at the Diabetes Research Institute, D. Alvaro and E. Gaudio at Sapienza For the sake of brevity, we will not discuss findings regarding
Medical Center, P. Sethupathy at Cornell, and L. Reid at UNC. the lineage restriction of embryonic stem (ES) cells or induced

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
524 THE LIVER:  THE BILIARY TREE AS A ROOT SYSTEM FOR HEPATIC/PANCREATIC ORGANOGENESIS

pluripotent stem (iPS) cells to an hepatic or pancreatic fate. mesenchyme, connections associated with rapid vascularization
Information from these studies is provided in representative of the forming liver tissue. This results in the liver and the
recent articles and reviews [10–13]. ­ventral pancreas sharing the hepato‐pancreatic common duct
This chapter focuses primarily on studies in human tissues connecting them to the duodenum.
and with some references to similar investigations in mice, rats,
and/or pigs. The phenomena associated with stem/progenitor
cell maturational lineages in hepatic and pancreatic organogen-
esis are evident in all mammals studied (mice, rats, pigs, and THE BILIARY TREE AS A ROOT SYSTEM
humans) but with a few distinctive variations in some species as
noted below.
FOR HEPATIC/PANCREATIC
The stem cell or progenitor cell populations are indicated by ORGANOGENESIS (GENERAL ANATOMY
an acronym which is preceded by a small letter indicating the AND IN SITU STUDIES)
species: m, murine; r, rat; p, porcine; h, human.
Newly discovered within the last decade is that hepatic and pan-
creatic organogenesis are ongoing throughout life “fueled” by a
network of stem cells within the biliary tree, the ramifying ducts
EMBRYONIC DEVELOPMENT connecting the liver and the pancreas to the duodenum [5, 9, 27,
30–33] (Figure 42.1). The biliary tree, long assumed to be the
During early development, definitive endoderm derives from conduits managing removal of bile from the liver and digestive
embryonic stem cells through the effects of a number of key enzymes from the pancreas, has proven also to be a “root”
transcription factors including Goosecoid, MIXL1, SMAD2/3, ­system, a reservoir of stem cells giving rise to maturational
SOX7, and SOX17 (a transcription factor essential for differen- ­lineages of cells contributing to hepatic and pancreatic organo-
tiation of liver) [14]. Endoderm subsequently segregates into genesis and to regenerative processes in these organs. The most
foregut (lung, thyroid), midgut (pancreas, biliary tree, and liver), primitive of the stem cells are located within extramural peribil-
and both foregut and hindgut (intestine) through the effects of iary glands (PBGs, stem cell niches for biliary tree stem cells)
specific mixes of transcription factors. Those dictating the mid- tethered to the outside of large ducts (i.e. in humans: greater
gut organs include SOX9 (transcription factor associated with than 300 μm) [34]. Their functions are not known but are
endodermal tissues), SOX17, FOXA1/FOXA2 (forkhead box hypothesized to be “seeds” for organogenesis given that they
a1/2), Onecut2/OC‐2, and others [15–18]. The liver, biliary tree, sprout into ducts with regenerative demands (Reid and associ-
and pancreas derive from midgut endoderm established at the ates, unpublished observations).
gastrulation stage of early embryonic development [19]. Among The network of intramural (within the duct walls) niches
the other organs of endodermal origin, endogenous adult stem begins with the Brunner’s glands, submucosal glands found in the
cells have been identified in most, including the small and large duodenum [35, 36] (Cardinale et  al. personal communication).
intestines [20], the stomach [21], and the lungs [22, 23]. These are located between the major papilla (the papilla of Vater),
The pancreas is distinct in that lineage tracing experiments the entranceway to the hepato‐pancreatic duct, and the minor
indicate that there are only very rare stem cells postnatally [24– papilla, the port connecting the duodenum to the dorsal pancreatic
26]. Clarification of this paradox has emerged with the many duct. The Brunner’s glands are not found elsewhere within the
studies on the biliary tree: the stem cells for the pancreas are intestinal tract. Indeed, they are used to define the transition from
not  in the pancreas proper but rather within the biliary tree, the duodenum to the beginning of the small intestine.
especially within the peribiliary glands (PBGs) of the hepato‐ The intramural network of stem cells extends throughout the
pancreatic common duct. These pancreatic stem cells are biliary tree in the PBGs found in high numbers behind branch
anatomically linked and give rise to committed progenitors
­ points in the biliary tree, in the cystic duct connecting the gall-
located in pancreatic duct glands (PDGs) within the pancreas bladder to the common duct, in the large intrahepatic bile ducts,
proper [27, 28]. and in the hepato‐pancreatic common duct [9, 30, 34, 37]. The
During fetal development, the formation of the liver and pan- gallbladder does not have PBGs, but its stem cells are found in
creas occurs with outgrowths on either side of the duodenum crypts at the base of villi in the gallbladder [29]. The second
that extend and ramify into a branching biliary tree structure largest numbers of PBGs are found within the large, intrahe-
that, at its ends, is influenced by the cardiac mesenchyme to patic bile ducts and connect anatomically to the canals of
form liver and by the retroperitoneal mesenchymal to form pan- Hering and thence to the sinusoidal plates of parenchymal cells
creas [28]. The gallbladder is a major branch point connected to within liver acini [34]. The largest numbers of PBGs are found
the common duct via the cystic duct [29]. The pancreas derives within the hepato‐pancreatic common duct and connect to the
from two separate anlage: the dorsal pancreas connected to the PDGs within the ventral pancreas [27, 34]. Although not yet
duodenum via the accessory duct. The ventral pancreas begins well defined, they are found also in the accessory duct connect-
as a branch from the common duct and nearer to the duodenum. ing to the dorsal pancreas.
The formation of the intestine involves a twisting motion that The network of stem cells transition to niches of bipotent,
swings the ventral pancreas anlage to the other side where it transit amplifying cells, cells with considerable proliferative
subsequently merges with the dorsal pancreas anlage to form potential but questionable self‐replicative ability. These com-
the complete pancreatic organ. The liver cannot swing to the prise hepatoblasts, located next to the canals of Hering [38–41]
opposite side, given its size and its connections into the cardiac and to pancreatic ductal progenitors in the PDGs [42–44].
42:  Stem Cell‐Fueled Maturational Lineages in Hepatic and Pancreatic Organogenesis 525

Figure 42.1  The anatomy of the connections between duodenum, biliary tree, gallbladder, liver, and pancreas in humans. Most of the anatomi-
cal features are self‐evident from the figure. The connections of the biliary tree to the duodenum are via two ports, the ampulla of Vater for the
hepato‐pancreatic common duct, and the minor duodenal papilla for the accessory duct (also called the duct of Santorini) connecting to the
dorsal pancreas.

These, in turn give rise to unipotent committed progenitors that blood and interstitial fluid. The net sum of the activities of cells
link to mature hepatic or pancreatic cell populations. at the sequential maturational lineage stages yields that for the
composite tissue.
The phenomena associated with stem cell networks in
hepatic/pancreatic organogenesis are common to all mammals.
LINEAGE‐DEPENDENT EPITHELIAL– However, there are some species‐associated distinctions.
MESENCHYMAL CELL PARTNERSHIPS Examples include the species‐specific variations with respect to
ploidy profiles in the liver (note: these have not been analyzed
The cells throughout the biliary tree, liver, and pancreas are in yet in pancreas). The lineages in all mammals begin in fetal and
epithelial–mesenchymal cell partnerships that undergo coordi- neonatal tissues with entirely diploid cells that transition to
nate maturation resulting in lineage‐stage‐dependent paracrine increasing proportions of polyploid cells in adulthood and espe-
signaling influencing the stage‐dependent phenotypic traits as cially in geriatric hosts. Within the liver acinus, diploid cells are
summarized in more detail in prior reviews [1, 45, 46]. In brief, always found periportally; the polyploid cells are always found
the beginning stages are comprised of epithelial stem cells pericentrally. What varies is the proportion of the cells of the
­partnered with angioblasts (CD117+, prominin 1 [CD133]+, liver plates that are polyploid. In humans, by around 20 years of
vascular endothelial cell growth factor [VEGF]‐receptor+, Van age, the shift is to a minor fraction of tetraploid cells, all in zone
Willebrand Factor+, CD31-) [47]. These give rise to cellular three of the liver acinus; in rats, by four weeks of age, it transi-
descendants that mature coordinately and then split into two tions to predominantly tetraploid cells (80%), all in zone two
branches: epithelia partnered with endothelia (hepatocytes, islets) and with a small fraction (10%) of octoploid cells in zone three;
and epithelia partnered with stellate cells and their descendants, and in mice the shift occurs by three to four weeks of age and
stroma and myofibroblasts (cholangiocytes, acinar cells). The results in 97% polyploid cells comprised mostly of tetraploid
endothelial cell precursors (CD133+, VEGF‐receptor+, Van and octoploid in a part of zone one, all of zone two, and with a
Willebrand Factor+, CD31+) and the stellate cell precursors minor fraction of cells that are 16N and 32N in zone three. The
(CD146+, intercellular adhesion molecule [ICAM]‐1+, exceptions to this ploidy profile are the newly discovered dip-
desmin+, alpha‐smooth muscle actin [ASMA]+, glial fibrillary loid parenchymal cells linked on their lateral borders to endothe-
acidic protein [GFAP]-) give rise to descendants with distinct lia encompassing the central vein; these constitute a reservoir of
phenotypic traits [1] The lineage‐dependent traits comprise committed (unipotent) hepatocytic progenitors that replace
morphology, cell size, DNA content (ploidy), growth potential, apoptotic (senescing) hepatocytes [48].
antigenic profile, gene expression, and lineage‐dependent par- Another notable variation is that pigs and oxen do not have an
acrine signaling by the synergistic effects of extracellular matrix hepato‐pancreatic common duct, the location of the stem cell
components, matrix‐bound signals (growth factors and cytokines niches for the ventral pancreas for all other mammals. Ongoing
linked primarily to glycosaminoglycan [GAG] chains bound to studies are exploring possible locations for these stem cell
core proteins in proteoglycans) and soluble signals present in niches for the ventral pancreas in pigs.
526 THE LIVER:  RADIAL AXIS OF MATURATION

MICROENVIRONMENT OF THE STEM b­iliary tree, with the greatest numbers being present in the
CELL AND PROGENITOR CELL NICHES hepato‐pancreatic common duct and the large intrahepatic
bile ducts [9, 34]. Other than findings from the pioneering stud-
Stem and progenitor cells reside in discrete locations called niches, ies of Nakanuma and associates [64–66], almost nothing used to
with unique microenvironments that are poorly defined for the be known of the possible roles of the PBGs until the recent stud-
hepato/biliary/pancreatic network (Table  42.1 and Figure  42.2). ies within the last decade analyzing them as reservoirs, crypts,
The niches include PBGs in the extrahepatic and intrahepatic containing stem cell populations. Each PBG contains a ring of
­biliary tree [4, 27, 30, 34] and in the cystic duct connecting the cells at its perimeter and is replete with mucous production and
gallbladder to the biliary tree [29]; the ductal plates in fetal and periodic acid‐Schiff (PAS)‐positive material in its center. The
neonatal livers and transitioning into the canals of Hering in pedi- cells in the ring are phenotypically quite homogeneous in
atric and adult livers [49–53]; and the PDGs found throughout the the PBGs in some sites (e.g. near the fibromuscular layers of the
pancreas but with an especial concentration in the head of the pan- hepato‐pancreatic common duct and the large intrahepatic bile
creas [28, 54–56]. The gallbladder does not have PBGs, but does ducts), but are quite heterogeneous in other sites (e.g. neared to
have crypts at the base of villi and contains stem cells similar to the lumens of all ducts, especially the cystic duct, hilum, com-
late stage biliary tree stem cells (BTSCs). Collectively, these mon duct). The pattern of the variations implicate maturational
niches form a network that is continuous throughout the biliary lineages for which there are two axes [27, 34].
tree and anatomically connects directly through to the canals of
Hering to the sinusoidal plates within the liver and through to the
PDGs, the reservoirs of committed progenitors within the pan-
creas, and then to the acinar cells and islets.
RADIAL AXIS OF MATURATION
As noted above, a key paradigm to the tissue organization is the
A radial axis of maturation is evident within the duct walls (the
epithelial–mesenchymal cell partnerships. It can be mimicked in
intramural PBGs), beginning with the cells in PBGs near the fibro-
vitro by use of feeder cells of the relevant mesenchymal type or by
muscular layer (center of the duct walls) and ending with cells at
use of defined mixes of matrix components and soluble signals [47].
the ducts’ lumens (Figure 42.3). The most primitive of the stem
Matrix and soluble signals in the stem cell and progenitor cell niches
cells, those in the PBGs at the fibromuscular layer, have no mark-
have been only partially defined [40, 41, 47, 57, 58]. That which is
ers of adult liver or pancreas but instead have moderately high lev-
known is summarized in Table 42.1 and in Figure 42.2. The known
els of expression of pluripotency genes (e.g. SOX2 [a transcription
matrix chemistry of these stem/progenitor cell niches is domi-
factor that is essential for maintaining self‐renewal, or pluripo-
nated by hyaluronans (HA) [59], non‐sulfated glycosaminoglycans
tency in embryonic and determined stem cells], SALL4 [Sal‐like
(GAGs), and minimally sulfated forms of chondroitin sulfate prote-
protein 4 found to be important for self‐replication of stem cells],
oglycans (CS‐PGs) [60]. In association with the transit amplifying
BMI-1, NANOG, KLF4/5, octamer‐binding transcription factor 4
cells, there are also fetal collagens (e.g. type IV), and fetal adhesion
[OCT4, also known as POU5F1 {POU domain, class 5, transcrip-
molecules (e.g. fetal laminins) along with more sulfated proteogly-
tion factor 1}, a gene expressed by stem cells]); co‐expression of
cans (e.g. heparan sulfate proteoglycans [HS‐PGs] and CS‐PGs)
transcription factors for both liver and pancreas (e.g. SOX17,
[61, 62]. Receptors for one or another of the known hyaluronan
PDX1 [pancreatic and duodenal homeobox 1, a transcription fac-
receptors (e.g. CD44) are common features of the stem cells [63].
tor critical for pancreatic development]); express classic markers
With maturation to adult cell types, there is a branching to separate
indicative of active proliferation such as Ki‐67; have high levels of
lineages (hepatocytes, islets) associated with endothelia versus chol-
CD44 (hyaluronan receptors); and express sodium iodide sym-
angiocytes and acinar cells associated with stellate cells. The matrix
porter (NIS). NIS is hypothesized to transport anionic intermedi-
chemistry changes parallel the split: those in epithelial–endothelial
ates needed in the synthesis of hyaluronans [67].
relations contain matrix chemistry dominated by network collagens
As one progresses toward the lumen, the cells in the PBGs
(type IV, type VI), by forms of laminin, and by HS‐PGS with
lose expression of the pluripotency traits, diminish the evidence
increasing sulfation resulting in the expression of heparin proteogly-
for proliferation, and acquire intermediate markers associated
cans (HP‐PGs) associated with the most mature cells (polyploid
with stemness such as LGR5 and epithelial cell adhesion mole-
hepatocytes, mature islets). The branch of the epithelia‐stellate cells
cule (EpCAM). At the lumens, all traits of stemness have faded
gives rise to cholangiocytes and acinar cells and is associated with a
and been replaced with hepatic traits in the vicinity of the car-
matrix chemistry comprised of fibrillar collagens (e.g. type III, type
diac mesenchyme or pancreatic traits for ducts in the vicinity of
I), adhesion molecules (forms of fibronectins, entactin), and with
the retroperitoneal mesenchyme.
CS‐PGs that transition in late stages of cells to highly sulfated forms,
We hypothesize that a new nomenclature will be required
dermatan sulfate proteoglycans (DS‐PGs).
for the PBGs and PDGs, since they have traits that are less
those of glands and instead have features of crypts. The phe-
nomenon parallels that observed in the intestinal crypts con-
PERIBILIARY GLANDS taining stem cells that mature in a radial axis along the villi to
yield adult intestinal cells (e.g. enterocytes and goblet cells).
The PBGs occur throughout the biliary tree as intramural glands, Temporarily we will abstain from converting to a new nomen-
found within the bile duct walls, and as extramural glands that clature to enable further studies to be done that should help
are tethered by extensions to the large bile ducts [64]. They with establishing a logical new nomenclature for the compo-
occur in highest frequencies at the branching points of the nents of the network.
Table 42.1  Representative biomarkers of major lineage stages of precursors in organogenesis of liver and pancreas

Periportal region of PBGs in large intrahepatic PBGs in extrahepatic PBGs in hepato‐


liver acini Canals of Hering bile ducts biliary tree pancreatic common duct PDGs in pancreas
Lineage stage of cells Hepatoblasts, hepatic HpSCs Stage 1–3 BTSCs Pancreatic ductal progenitors
committed progenitors HpSCs PSCs
Endodermal transcription Sox9 SOX 9, SOX 9, SOX 17, PDX1 in Stages 1 and 2 BTSCs SOX 9, PDX1
factors SOX 17 SOX 9 and SOX 17 in Stage 3 SOX9 and PDX in
Stage 3
Pluripotency genes Low to none Intermediate level of Moderate to strong expression in stages 1 and 2 BTSCs; Intermediate level of None
expression intermediate level of expression in stage 3 BTSCs expression
OCT4, SOX2, NANOG, SALL4, BMi‐1, KLF4/KLF5
Cell adhesion molecules EpCAM, EpCAM, NCAM NCAM in stage 1 and 2 BTSCs, EpCAM and NCAM in stage 3 BTSCs EpCAM, ICAM‐1
ICAM‐1
Other stem cell markers Weak CXCR4 CD133, LGR5 CXCR4 in stages 1 and 2 BTSCs, CD133 in all of the stages of BTSCS; LGR5 in CXCR4, CD133, CD24
CD133, LGR5 stages 2 and 3 BTSCs
Hedgehog proteins Weak Indian and Sonic Strong Indian and Not yet studied Weak Sonic
Sonic
Matrix proteins Laminin, type IV collagen Laminin 5, type III Not yet studied Islets: network collagens; acinar
collagen cells: fibrillar collagens
GAGs/PGS HA, CD44, syndecans HA, CD44, minimally HA, CD44, others not yet studied Islets: syndecans and glypicans;
(HS‐PGs and CS‐PGs) sulfated CS‐PGs acinar cells: CS‐PGs, DS‐PGs
Multidrug resistance MDRI‐negative; MDR1 moderate, Not yet studied None
genes ABCG2‐moderate ABCG2 very strong
Hepatic traits Albumin +, Albumin +/‐ Negative for hepatic traits
AFP +++, P450A7, AFP‐, P450A7‐
Glycogen
Pancreatic traits Negative for pancreatic traits ISL1, PROX1, NGN3, MAFA, MUC6,
NeuroD, PAX4 Nkxx6.1/NKx6.2, Ptf1a, Glut2
HA, hyaluronans; HS‐PGs, heparan sulfate proteoglycan; CS‐PGs, chondroitin sulfate proteoglycan; DS‐PGs, dermatan sulfate proteoglycan; syndecans, proteoglycans with a transmembrane core protein;
glypicans, proteoglycans linked to plasma membrane by PI linkages; MDR1, multidrug resistance genes; HpSCs, hepatic stem cells; HBs, hepatoblasts; BTSCs, biliary tree stem cells; PSCs, pancreatic
stem cells; PBGs, peribiliary glands; PDGs, pancreatic duct glands.
All of the known stages of the BTSCs are in the large intrahepatic bile ducts including in the extrahepatic biliary tree and in the hepato‐pancreatic common duct. The precursors to these are in the
extramural peribiliary glands and in the Brunner’s glands in the submucosa of the duodenum.

0004435634.INDD 527 11/5/2019 6:46:31 PM


528 THE LIVER:  RADIAL AXIS OF MATURATION

A. Stem Cell Niches relevant to Hepatic and Pancreatic Organogenesis


Stem Cells = self-replicate; multipotent; all express moderate levels of pluripotency genes; all express one or more of the isoforms of
CD44 (hyaluronan receptors) and the most primitive ones found also express NIS (sodium iodide symporter). NIS is hypothesized
to be a transporter of anionic intermediates in the hyaluronan biosynthesis pathway.

Determined Endodermal Stem Cell subpopulations identified to date:


Extramural Peribiliary Gland Stem Cells: extremely primitive. With regenerative demands, produce ductules.
Intramural Stem Cell Populations:
Brunner’s Glands’ Stem Cells --submucosa of the duodenum; extremely primitive
Biliary Tree Stem Cells (BTSCs). Three phenotypically distinct stages within bile duct walls
Stages: 1) NIS+, Negative for LGR5 and EpCAM 2) NIS+, LGR5+ EpCAM- 3) NIS-, LGR5+, EpCAM+
Gallbladder stem cells --similar to stage 3-BTSCs
Hepatic stem cells --primarily within canals of Hering and in PBGs of biliary tree
Pancreatic stem cells --primarily within PBGs of hepato-pancreatic common duct
Note: Only committed progenitors found within pancreas proper (so, no true stem cells within pancreas)

Epithelial-mesenchymal Cell Partnerships. All are comprised of epithelial stem cells coupled with mesenchymal cell partners;
there is lineage-stage-specific paracrine signaling and coordinate maturation. Stem Cell stages= epithelial stem cells partnered with
angioblasts

Stem Cell Niche Microenvironment = known components comprise hyaluronans, non-sulfated glycosaminoglycans, plus minimally
sulfated chondroitin sulfate proteoglycans (no heparan sulfate proteoglycans).

B. Intermediates In Hepatic and Pancreatic Organogenesis

Transit Amplifying Cells = highly proliferative but debatable if they self-replicate. Weak to negligible levels of pluripotency genes.

Hepatoblasts --bipotent (yield hepatocytes and cholangiocytes); signature feature= alpha-fetoprotein (AFP); located adjacent
to canals of Hering; anatomically connected through the hepatic stem cells to network of stem/progenitor cells
within peribiliary glands (PBGs) of the large intrahepatic bile ducts.

Pancreatic Ductal Progenitor Cells – bipotent (yield acinar cells and islets); found within pancreatic duct
glands (PDGs) within the pancreas proper; are anatomically connected to the network of stem cells within PBGs of
the hepato-pancreatic common duct of the biliary tree.

Committed Progenitor Cells = do not self-replicate; highly proliferative; unipotent; no expression of pluripotency
genes. Anatomically linked to the mature cells

Epithelial-mesenchymal cell partnerships:


Hepatocytic or islet progenitors and descendants coupled to endothelial cell lineage stages
Cholangiocytic or acinar progenitors and descendants coupled to stellate cell lineage stages; late
lineage stages are associated with myofibroblasts

Niche Microenvironment = known components include hyaluronans + minimally sulfated sulfated chondroitin
sulfate proteoglycans and heparan sulfate proteoglycans + laminin + type IV collagen

Figure 42.2  The stem/progenitor cell niches in the biliary tree, liver, and pancreas. (a) A chart shows the known connections between the niches
throughout the biliary tree and within the organs. (b) Summarizes the features of the stem cell niches. (c) A summary of the transit amplifying cells.

The maturational progression of the cells is always within the


Proximal‐to‐distal axis in maturation radial axes in the duct walls, but that in or near the liver is within
A proximal‐to‐distal axis in the maturation occurs beginning at the domain of the cardiac mesenchyme and results in mature
the duodenum with the stem cells in the Brunner’s glands; con- cells of an hepatic fate. The radial axes within ducts near the
nects into the stem cells within the PBGs within the hepato‐­ pancreas are in the domain of the retroperitoneal mesenchyme
pancreatic common duct; and then transitions into the pancreas and result in mature cells of pancreatic fates. Those in between
to connect with the PDGs. The branch for the liver is via the liver and pancreas yield cells with mature bile duct markers. The
common duct and once past the cystic duct leading to the gall- implications are that isolated BTSCs, especially the most primi-
bladder connects into the PBGs of the large intrahepatic bile tive of them, should be able ex vivo to give rise both to liver and
ducts. These give rise to links with the canals of Hering that pancreas if provided the requisite paracrine signals from either
transition into the sinusoidal plates of the liver acini. cardiac mesenchyme versus retroperitoneal mesenchyme. This
42:  Stem Cell‐Fueled Maturational Lineages in Hepatic and Pancreatic Organogenesis 529

C. Hepato/Pancreatic Stem/Progenitor Cell Niches


Extramural PBGs Brunner’s Glands
Tethered to outside of large Bile Ducts (Submucosa of the Duodenum)
Very primitive Endodermal Stem Cells BGSCs, CPs

Common Duct Intramural PBGs Hepato-Pancreatic Common Duct


(Extrahepatic Biliary Tree)
BTSCs, HpSCs, PSCs, CPs

PBGs (Large Intrahepatic Bile Ducts) PBGs (Cystic Duct) PBGs (Hepato-pancreatic
BTSCs, HpSCs, HBs, CPs BTSCs, CPs Common Duct)
BTSCs, PSCs, CPs

Ductal Plates Crypts-bottoms of villi


Pancreatic Duct Glands
Canals of Hering (Gallbladder) (Pancreas)
(Liver) Stage 3-BTSCs, CPs Pancreatic Ductal Progenitors (CPs)
HpSCs, HBs, CPs

Hepatocytes and Cholangiocytes Cholangiocytes Islets and Acinar Cells


BGSCs = Brunner’s glands stem cells
BTSCs = biliary tree stem cells (3 stages)
PBG= Peribiliary Gland HpSCs = hepatic stem cells; HBs = Hepatoblasts
Extramural PBGs = tethered to duct wall PSCs = pancreatic stem cells
Intramural PBGs = inside the duct wall CPs = committed progenitors (unipotent)

Figure 42.2  (Continued)

(a)

Figure 42.3  The radial axis of maturation in (a) liver, (b) pancreas, and (c) gallbladder. The peribiliary glands (PBGs) near to the fibromuscular layer
within the bile duct walls are crypts containing the most primitive of the stem cells, ones that can give rise both to liver and to pancreas. There are hints
that the cells move along the walls of the PBGs and progress upwards towards the lumens of the ducts. With that progression, the phenotypic traits
transition from those for stem cells to those for mature cells. Some of the traits for the cells in the radial axis maturation are noted. The gallbladder does
not have PBGs, but does have late stage BTSCs located at the base of villi in the gallbladder. These move upwards toward the tops of the villi.

is indeed what has been found [27, 30]. It also implicates the These phenomena parallel the well‐described, intestinal mat-
potential for using the BTSCs in cell therapies for both liver and urational lineage system. The radial axis of maturation in the
for pancreas, an hypothesis that is currently under investigation. intestine progresses from stem cells in the crypts to fully dif-
The network provides a biological framework for ongoing ferentiated cells at the tops of the villi. The proximal‐to‐distal
organogenesis of liver, biliary tree and pancreas throughout life. axis follows the length of the intestine and results in distinct
530 THE LIVER:  EX VIVO STUDIES ON THE STEM/PROGENITORS

(b)

(c)

Figure 42.3  (Continued)

mature cells depending on whether the radial axis is located in The culture conditions required for the different lineage
the esophagus, stomach, duodenum, small or large intestine. stages are distinct. The stem cell stages require conditions with
soluble forms of hyaluronans for organoids and with hyaluronan
hydrogel substrata for monolayers or for embedded organoids;
these hydrogel must be very soft, under 100 Pa to maintain
EX VIVO STUDIES ON THE STEM/ stemness traits [68]. The medium required for the stem cells is
PROGENITORS serum‐free, devoid of growth factors and cytokines, and tailored
for the cells at this stage. One of the best ones was that estab-
The in situ and anatomical findings summarized above are com- lished by Hiroshi Kubota [38]. It is comprised of any rich basal
plemented by ex vivo (monolayer and organoid) cultures of the medium with low calcium (around 0.3 mM), no copper, sele-
various lineage stages including those of Brunner’s glands nium (10−10 M), zinc (10−12 M), insulin (around 5 μg mL−1),
(Cardinale et al. personal communication), of two of the BTSC transferrin/Fe (around 5 μg mL−1), high density lipoprotein
stages [30], of the hepatic stem cells (HpSCs) and hepatoblasts (around 10 μg mL−1), and a defined mixture of purified free fatty
(HBs) [1, 39, 40, 68], pancreatic ductal progenitors and islets acids bound to highly purified albumin. Mature cells do not
[28], and adult hepatocytes and adult cholangiocytes [47, 69]. ­survive in Kubota’s medium, only the stem cells from both
42:  Stem Cell‐Fueled Maturational Lineages in Hepatic and Pancreatic Organogenesis 531

epithelial and mesenchymal cell lineages. However, more rapid The gene expression profiles of cells in 3D hydrogels com-
expansion of the mesenchymal partners of the epithelial stem plement the morphological observations. For example, cells
cells, such as the angioblasts, and conditions permissive for self‐ cultured under conditions for hepatocytes produce albumin,
replication of both are facilitated by the addition of leukemia transferrin, and P450s. Cells in conditions for cholangiocytes
inhibitory factor (LIF) and by VEGF [30, 38, 41, 47, 69]. Thus, express anion exchanger 2 (AE2), cystic fibrosis transmem-
the conditions co‐select for any of the true stem cell subpopula- brane conductancy regulator (CFTR), gamma glutamyl trans-
tions and their mesenchymal cell partners, angioblasts and their peptidase (GGT), and secretin receptor. Cells in conditions
descendants, precursors of stellate cells, or endothelia [38, 47] for pancreatic islets express transcription factor PDX1 and
Under these conditions, we observed two major types of the hormones glucagon, somatostatin, and insulin. Specific
BTSC colonies in cultures that correlate with stage two and staining for human C‐peptide confirms de novo synthesis of
stage three BTSCs from the in situ studies; we have yet to iden- proinsulin, and its secretion can be regulated in response to
tify conditions for the stage one BTSCs. The stage two category the level of glucose. In vivo studies in which the cultured cells
consist of cells that undulate (“dancing cells”), are very motile, are transplanted provide further evidence for the multipotency
and initially do not express EpCAM (CD326) but acquire it at of the BTSCs for hepatic, biliary tree, and pancreatic fates
the edges (the perimeters) of the colonies, corresponding to [4, 27, 30, 70].
slight cellular differentiation [30]. These are precursors to stage To confirm endocrine pancreatic differentiation, pre‐
three BTSCs that show uniform expression of EpCAM from the induced neo‐islet structures were implanted into mouse fat
outset and display a carpet‐like appearance with cells of uni- pads, and the animals were treated with a toxin (streptozo-
form morphology [30]. The stage three BTSCs found through- tocin) at a dose sufficient to destroy their own pancreatic beta
out the biliary tree are also the same or similar to the stem cell cells, but not human beta cells. Those mice transplanted with
subpopulation found in the gallbladder [29]. the human neo‐islets showed significant resistance to hyper-
The expansion potential of the cells in culture in Kubota’s glycemia compared to controls that did not receive cell
medium is considerable: two to three cells can grow to colonies ­therapy. The presence of functional beta‐like cells derived
of more than 500 000 cells in around eight weeks [30]. The cells from the biliary tree stem cells produced serum‐levels of
retain a stable stem cell phenotype (i.e. self‐renew) throughout human C‐peptide, that was regulated appropriately in response
months of culture, and may be subcultured (“passaged”). Initially to a glucose challenge [30]. Further studies have confirmed
cells show a typical division time of about one to two days, but and expanded upon these initial findings, leading us to con-
within a week, they slow to a division every two to three days. At clude that the hepato‐pancreatic common duct is the major
eight weeks the colonies contain cells in the centers that are mor- reservoir of stem cells giving rise to committed progenitors
phologically uniform, small (7–9 μm) and express high levels of found in pancreatic duct glands, and thence to pancreatic
stem cell markers. Cells at the edges of the large colonies are islets throughout life [27].
slightly larger (around 10–12 μm), have weak expression of
EpCAM and expression of markers intermediate in the differen-
tiation pathways, indicating potential loss of stemness and transi-
tion to more mature progenitors. HEPATIC STEM CELLS
Using three‐dimensional (3D) hyaluronan hydrogels contain-
ing appropriate signaling molecules, the BTSCs can be induced Those familiar with the myth of Prometheus will recall that the
to differentiate to hepatocytes, cholangiocytes, or pancreatic liver possesses a remarkable capacity for regeneration [2, 3, 71].
neo‐islets [30]. We have not succeeded yet to lineage restrict the Yet liver diseases, potentially leading to organ failure due to
BTSCs to acinar cells. The differentiation is achieved by embed- hepatitis viruses, alcohol consumption, diet and metabolic dis-
ding the BTSCs in specific mixes of extracellular matrix orders, and other causes, constitute a major medical burden
­components (hyaluronans and type I collagen for bile ducts; [72–74].
hyaluronans and type IV collagen and laminin for hepatocytes Cell‐based therapies and tissue engineering represent possi-
or islets) and providing a serum‐free, hormonally defined ble approaches to address these needs [1, 75–78]. Sourcing of
medium (HDM) tailored for each specific transit amplifying cells for such applications is a significant challenge. In some
cells or mature cell type. countries it is possible to obtain fetal tissues. In others neonatal
The HDM are prepared by supplementing Kubota’s medium or adult tissues can be used. Given the newly discovered source
with copper (10−12 M), higher calcium (0.6 mM), bFGF (basic of stem cell populations in the biliary tree and precursors for
fibroblast growth factor, 10 ng mL−1) and then adding a unique both liver and pancreas, the need for liver or pancreatic tissue as
set of hormones and growth factors for hepatoblasts (EGF, a source is reduced. There is likely to be a transition to biliary
hepatocyte growth factor [HGF], T3, soluble [or substrata] tree tissue as a primary source for clinical programs, and it is
forms of type IV collagen and laminin), hepatocytes (substrata considerably easier to obtain.
of type IV collagen and laminin plus glucagon, galactose, T3, The role(s) of hepatic stem cells in the normal maintenance
oncostatin M, HGF, EGF, glucocorticoids), cholangiocytes of the liver and in regeneration from various insults remains a
(substrata of type I and III collagen + HGF, EGF, VEGF, gluco- subject of active research and debate [52, 53, 71, 78–85].
corticoids), versus for pancreatic islets (soluble [or substrata] Although there is now general acceptance that hepatic stem
forms of type IV collagen and laminin) plus B27, ascorbic acid, cells exist postnatally, their relevance is compared and debated
cyclopamine, retinoic acid, HGF, and, after four days, replace- with that of the plasticity of the postnatal parenchymal cells
ment of bFGF with Exendin‐4] [30]. [86]. Further discussion of this is given below.
532 THE LIVER:  HEPATIC STEM CELL ISOLATION AND EXPANSION

HEPATIC STEM CELL ISOLATION More recently, we have observed that Sal‐like protein 4
AND EXPANSION (SALL4, found to be important for self‐replication of stem cells) is
strongly expressed in the hBTSCs and hHpSCs but not in commit-
Information on the findings on HpSCs is important, given that ted progenitors of either liver or pancreas [93]. SALL4 is a mem-
human hepatic stem cells (hHpSCs) have already been used in ber of a family of zinc finger transcription factors and a regulator
clinical trials in India and in Italy for patients with diverse liver of embryogenesis, organogenesis, and pluripotency. It can elicit
diseases and conditions [70, 87–89]. The isolation of hHpSCs reprogramming of somatic cells and is a marker of stem cells.
from fetal, neonatal, and adult human livers was achieved by A crucial prerequisite for successful expansion of hHpSCs is
immunoselection with a monoclonal antibody for the surface to mimic an appropriate microenvironment. When selected for
marker EpCAM [41]. These cells constitute approximately one growth in vitro on tissue culture plastic and in Kubota’s medium,
percent (0.5–1.5%) of the total liver population from early the hHpSCs grow as colonies with feeders of angioblasts
childhood onwards. Unlike mature hepatocytes, they survive (CD117+, VEGF‐receptor+, CD133+, Von Willebrand Factor+)
extended periods of ischemia, allowing collection even several [41, 47, 58]; the feeders can be replaced with soft forms of hya-
days after cardiac arrest [90]. The hHpSCs express additional luronans (under 100 Pa) and type III collagen [47, 68, 91, 94].
surface markers often found on stem/progenitor cells, such as The cells expand for months under these conditions. By con-
CD133 (prominin), CD56 (neural cell adhesion molecule, trast, hHBs survive for only about a week under these same con-
NCAM), and CD44 (the hyaluronan receptor); they also express ditions, but they can survive if they are co‐cultured with stellate
characteristic endodermal transcription factors SOX9, SOX17, cell precursors (CD146+, alpha‐smooth muscle actin+, desmin+,
and HES1. They are small (diameter 7–9 μm, which is less than VCAM+, ICAM‐1+, GFAP‐negative) or feeders of mesenchy-
half that of mature parenchymal cells) and express weak or neg- mal stem cells (MSCs). The stellate [95] feeder cells (or feeders
ligible levels of adult liver‐specific functions such as albumin, of MSCs) can be replaced with hyaluronans, type IV collagen,
cytochrome P450s, and transferrin. The hHpSCs display far and/or laminin [47, 96, 97]. The medium and matrix conditions
greater capacity to proliferate in culture than hepatocytes or described above allow for flow cytometrically‐purified hHpSCs
cholangiocytes, and can continue to expand for months with a or hHBs to survive and proliferate in culture and without the
doubling time of 36–40 hours. The colonies that form look need for feeders. Both type III collagen and hyaluronans are
remarkably similar to those of embryonic stem (ES) cells or constituents of the normal liver stem cell niche [47, 57, 91].
induced pluripotent stem (iPS) cells [5, 1, 47, 91]. A systematic study of hHpSC behavior in 3D cultures using
The hHpSCs serve as immediate precursors of human hepa- hyaluronan hydrogels of differing stiffness indicated that rigid-
toblasts (hHBs). The hHBs are readily distinguished from hHp- ity of the microenvironment is an important parameter in regu-
SCs by the expression of AFP and intercellular adhesion lating maintenance of stemness versus differentiation to more
molecule‐1 (ICAM‐1), for which the hHpSCs are entirely nega- restricted progenitors [68]. This was studied previously in dif-
tive [41, 47, 51]. The hHBs, in turn, are precursors of committed ferentiation of progenitors for bone and other hard tissues, but
unipotent progenitors for hepatocytes and cholangiocytes. When this report is the first for internal organs such as the liver.
injected into livers of immune‐deficient mice, the ­hHpSCs give The hHpSCs, like human ES cells, grow in tight colonies.
rise to cells expressing characteristic human liver and bile duct Dissociating either type of stem cells has proven to be an impor-
proteins, especially after the host’s liver has been damaged by tant practical problem for their efficient expansion ex vivo and
treatment with carbon tetrachloride [41]. for cryopreservation [98]. When treated enzymatically to gener-
Whereas there has been limited success to achieve ex vivo ate a single cell suspension, both of these stem cell types undergo
expansion of hematopoietic stem cells, the hHpSCs proliferate a high level of cell death. Ding’s laboratory screened for chemi-
for a sustained period in Kubota’s medium [38] which, as stated cals that would enable ES cells to survive enzymatic dissociation
previously, contains no additional growth factor or cytokine. and remain pluripotent. They identified two compounds, a
Pathways important for hHpSC survival in vivo, such as 2,4‐disubstituted thiazole (Thiazovivin) and a 2,4‐disubstituted
Hedgehog (Hh) signaling [40], are activated through autocrine pyrimidine (Tyrintegin), that met these criteria [99]. They found
loops. The expanded hHpSCs maintain a stable marker pheno- that Thiazovivin inhibits the Rho‐associated kinase (ROCK), a
type and also express the enzyme telomerase. The telomerase key component of the pathway that controls cytoskeleton remod-
mRNA and the protein encoded are localized to the nucleus eling, and a likely regulator of cell–ECM and cell–cell interac-
of the hHpSCs and the hHBs and telomeric enzymatic activity tions. Tyrintegin enhances attachment of dissociated ES cells to
correlate well with both the mRNA and protein levels [92]. ECM and stabilizes E‐cadherin. The investigators concluded that
However, later lineage stages (committed progenitors to late ES cell interactions in the normal niche generate signals essen-
­lineage stage mature cells) have no evidence of synthesis of tel- tial to survival, and that small molecules modulating those sig-
omerase but have large amounts of telomerase protein localized nals can maintain viability of dissociated cells.
cytoplasmically. Telomeric enzymatic activity does not corre- Interestingly, we have observed that hyaluronans, a normal
late with total telomerase protein levels; we hypothesize that it component of stem cell niches, can protect hHpSCs during dis-
correlates with the nuclear levels of the protein. Therefore, we sociation and cryopreservation [98]. Thus, hyaluronans, a natural
further hypothesize that regenerative demands will result in molecule, can mediate the needed protection of the cells as well
small amounts of the cytoplasmic reserves of telomerase relo- as the artificial ones described above. The addition of hyaluro-
cating to the nucleus. If we are correct, the enzymatic activity nans was found to protect cell adhesion mechanisms including
levels should correlate with the amount of telomerase (protein) the hyaluronan receptor, E‐cadherin, and certain integrins, mark-
in the nucleus [92]. ers shared by many other stem cell subpopulations [100].
42:  Stem Cell‐Fueled Maturational Lineages in Hepatic and Pancreatic Organogenesis 533

THE NEED FOR GRAFTING STRATEGIES microenvironments. Cytokines and other soluble factors neces-
IN TRANSPLANTATION OF CELLS sary for liver development and for the maintenance of differenti-
ated hepatocytes have been known for some time [46, 103, 104].
FROM SOLID ORGANS However, the specific and efficiently directed differentiation of
stem or progenitor cells to fully mature hepatocytes and cholan-
Cell therapies attempted to date are usually by delivery of the
giocytes ex vivo has remained a difficult challenge. This, in fact,
cells by a vascular route. This is logical for hemopoietic cell
is a general problem in much of stem cell biology, whether start-
types. Long‐lived hemopoietic cells have evolved to be able to
ing with lineage‐restricted adult stem cells or pluripotent ES and
flip between splice forms of matrix molecules (e.g. fibronec-
iPS cells. The most promising strategies are to make use of com-
tins) from ones with no cell binding domains (resulting in cells
plex extracellular matrix scaffolds, effectively solid‐state sign-
floating freely in blood or interstitial fluid) versus ones with
aling apparatuses that can guide the differentiation of the cells.
cell binding domains (resulting in attachment  –  a process
referred to as “homing”). By contrast, transplantation of cells
from solid organs involves cell types in which their attachment
proteins always have cell binding domains and so they rapidly EXTRACELLULAR MATRIX SCAFFOLDS
(within seconds) aggregate. When delivered to a target organ/
tissue by a vascular route, the aggregates cause an embolus In recent years, complex matrix scaffolds are being utilized for
essential for engraftment but if too large resulting potentially in optimal differentiation of cells [5–8]. However, all of the scaf-
lethal consequences for the host. Moreover, even when suc- fold types reported are limited and inefficient in their effects due
cessful, there is inefficient engraftment (typically around to their methods for isolation, ones resulting in the loss of criti-
10–20%) and with the remainder of the cells either dying or cal matrix components such as the proteoglycans. The only
dispersing to ectopic sites [94]. known method by which to isolate a matrix scaffold with reten-
Our studies and those conducted by many others have found tion of these critical factors is one developed by the Reid lab in
that infusion of mature hepatocytes achieves only around 20% partnership with collagen chemists [105]. They designed a
engraftment if injected into the portal vein of the liver [73, 101, method tailored to the known solubility constants of given col-
102]. Stem cells are even more challenging, with approximately lagens using the strategy to isolate a matrix complex with a salt
only 3% of the cells engrafting if administered via the portal buffer at a concentration to keep insoluble all of the known
vein (or via the spleen that connects directly to the portal vein). types of collagens in a given tissue. The insoluble complex so
This can be improved to around 20–25% engraftment in the isolated was termed “biomatrices” [106]. Frozen sections or
liver if the hHpSCs are injected into the hepatic artery [89]. The pulverized liver biomatrices used as cell culture substrata ena-
remaining majority of the hHpSCs either die or engraft in bled the long‐term survival of highly functional hepatocytes, far
ectopic sites, most commonly the lung. Cells that lodge in the beyond what could be achieved on plastic or with simple type I
vascular beds of ectopic sites can survive for months [100], a collagen gels. Recently, we have revisited the method and estab-
finding of unknown significance at this time, but of potential lished an improved protocol, one involving perfusion strategies
clinical concerns. and an improved delipidation method along with the high
We have devised grafting strategies for transplantation of salt  strategies, to prepare decellularized organs/tissues called
hHpSCs embedded into a mix of soluble signals and extracel- “­biomatrix scaffolds” [5]. They are tissue‐specific but mini-
lular matrix biomaterials (e.g. hyaluronans) found in stem cell mally (if at all) species‐specific, and they potently induce cell
niches [100]. The hHpSCs maintain a stable stem cell pheno- differentiation [5]. The biomatrix scaffolds contain more than
type under the graft conditions. The grafts were transplanted by 98% of the collagens and known collagen‐bound matrix compo-
injection grafting into the livers of immuno‐compromised nents, including most of the fibronectins, laminins, nidogen,
murine hosts, with and without carbon tetrachloride treatment, entactin, elastin, and so on, and essentially all the proteoglycans
to assess the effects of quiescent versus injured liver conditions. (PGs). They retain physiological levels of the known cytokines
Grafted cells remained localized to the livers, resulting in a and growth factors found in the tissue. Mature parenchymal
larger bolus of engrafted cells in the host livers under quiescent cells plated on biomatrix scaffolds in a serum‐free HDM
conditions and demonstrated more rapid expansion upon liver remained stable for many weeks and continued to express liver‐
injury. We therefore have proposed grafting as a preferred strat- specific functions equivalent to that achieved by freshly isolated
egy for cell therapies for solid organs such as liver [94, 100]. cells [107, 108].
Ongoing studies are resulting in assessment of other forms of
grafting (e.g. patch grafting) that enable transplantation of large
numbers of cells.
LIVER REGENERATION
The renowned regenerative capacity of the liver has inspired
DIFFERENTIATION countless studies on mechanisms associated with the process
[71]. We will not summarize that considerable literature but
The pharmacology of stem cell differentiation also must encom- refer the readers to some recent reviews [3, 109, 110]. Here we
pass both soluble signals (i.e. conventional biologics and/or will note only the known responses of the stem cells and pro-
drugs) and matrix components corresponding to the cells’ 3D genitors in two distinct forms of liver regeneration: that after
534 THE LIVER:  PLASTICITY ISSUES AND REGENERATIVE PHENOMENA

partial hepatectomy, and that after selective loss of cells in aci- Although cellular reprogramming is achievable under conditions
nar zone three (the pericentral zone). (We assume that a parallel ex vivo or in extreme artificial conditions in vivo such as hosts with
process occurs in pancreatic regeneration, though it has been suicidal transgenes, it has yet to be demonstrated in transplanted
studied in far less detail.) adult cells in common diseases. By contrast, there is evidence that
A key to understanding the responses is recognition of “feed- stem/progenitors give rise to maturational lineages of cells sup-
back loop signals,” factors produced by the most mature liver porting tissue regeneration [2, 32].
cells, those in zone three of the liver acinus, and secreted into Stem cell functions of BTSCs and HpSCs are manifested
the bile. The bile flows from pericentral zone to periportal zone both by their genetic signatures and by their clonogenic, self‐
and then into the biliary tree and finally into the gut. The signal- replicative capacity ex vivo for months when under specific
ing molecules include bile acids and salts that affect differentia- serum‐free, wholly defined conditions. In addition, these cell
tion [111]; acetylcholinesterase [112], which is produced by populations are multipotent, the other trait required for proof of
mature hepatocytes and serves to inactivate acetylcholine pro- stemness, as revealed by their ability to lineage restrict in vitro
duced by periportal cells [113, 114]; and heparins, which are or in vivo to hepatocytes, cholangiocytes, or islets depending on
produced by mature hepatocytes [115] (J. Esko, A. Cadwallader, their microenvironment before or after culture expansion [30,
and L. Reid, unpublished observations) and are relevant in 41]. This is consistent with the data of Dorrell et al. [126] who
­control of stem cells and of tissue‐specific gene expression showed phenotypic and functional similarity of organoid‐initi-
[116, 117]. In addition the flow of the bile mechanically affects ating cells in mouse pancreas and liver.
­primary cilia on periportal cells and thereby influences signal Another issue contributing to misunderstandings is that liver
transduction processes mediated by these organelles [118–120]. regenerative phenomena in murine versus human tissues can be
In the presence of feedback signals, the stem cells remain in a distinct [2, 32]. In long‐lasting chronic human liver diseases, a
quiescent state. Diminution or loss of these signals results in severe and progressive impairment of hepatocyte proliferative
dis‐inhibition of the stem/progenitor cell compartments. This capabilities is common. Indeed, specific insults exhaust hepato-
leads to hyperplasia of the stem cells and other early lineage cyte proliferation, induce cellular senescence, and/or arrest the
stage cells. Factors that may release the stem cell compartment hepatocyte cell cycle [32]. In human pathologic tissue, HpSCs
from the normal feedback signaling control loops include activate to produce the so‐called ductular reactions that give rise
viruses, toxins, or radiation that selectively kill cells in zone to nascent hepatocytes repopulating cirrhotic livers through for-
three, the pericentral zone of the acinus. The hyperplasia transi- mation of hepatocyte buds [2, 32]. By contrast, murine models
tions into differentiation of the cells. The resulting fully mature of liver injury typically do not result in a severe blockade of
cells produce bile, and the restoration or enhancement of feed- hepatocyte proliferation [2]. In a novel mouse model [32] in
back loop signals then inactivates the proliferative response. which apoptosis, necrosis, and senescence are induced in nearly
Regeneration of the liver after partial hepatectomy is distinct all hepatocytes, HpSC activation proved crucial for survival and
from that described above [71, 110, 121]. The tissue remaining functional liver reconstitution.
after surgical removal of a portion of the liver (e.g. two‐thirds of Evidence promoting plasticity in the liver is based on studies in
its mass) continues to have feedback loop signals, and the early a single publication by Tarlow et al. [128] These studies made use
lineage stage cells remain competent to respond to these signals. of a highly artificial model of liver regeneration: a murine model
The depletion below threshold levels of various liver functions of Type I tyrosinemia, caused by a shortage of the enzyme fuma-
and secreted products triggers DNA synthesis as a wave across rylacetoacetate hydrolase (FAH). Moreover, Tarlow et al. [128]
the liver plates [110]. However, the DNA synthesis in most of the subjected the FAH mice to a second insult: the administration of
cells of the liver (especially those in zones two and three) is not DDC (2′‐3′‐dideoxycytidine). DDC causes alterations of the bil-
accompanied by cytokinesis [122]. So these cells increase their iary tree mimicking the spectra of primary sclerosing cholangitis;
level of ploidy and demonstrate hypertrophic growth [123]. The the consequent secondary cholestasis heavily impacts the hepatic
polyploidy triggers an increased rate of apoptosis resulting in lineages, leading to the hyperactivation of ductular reactions
turnover of the liver. With the loss of the apoptotic cells, there is a shown in Tarlow et  al. [128]. The findings in such an extreme
low level of proliferation of the stem cells and early lineage stage model should not be used to make generic statements about stem/
cells to replace those cells eliminated during apoptotic processes. progenitors under either normal or ­typical disease conditions.
In mammalian species examined, this turnover occurs in weeks. Additional evidence used to promote the concept of plasticity
is based on experimental findings that cholangiocarcinomas can
originate from dedifferentiated hepatocytes. This provocative
assumption is based on observations by genetic lineage tracing
PLASTICITY ISSUES studies that cholangiocarcinomas can arise from cells expressing
AND REGENERATIVE PHENOMENA albumin or transthyretin [129], markers erroneously ascribed
only to mature hepatocytes. Albumin and transthyretin are
Kopp et al. [86] and others [3, 124] have hypothesized that plastic- expressed also in HpSCs and HBs subpopulations [30]. Similarly,
ity is dominant or even the sole mechanism mediating regenerative low levels of insulin are expressed in BTSC subpopulations in
responses for liver and pancreas. It is an hypothesis emanating the hepato‐pancreatic common duct and in multipotent progeni-
from discoveries that somatic cells can be reprogrammed by artifi- tors within the PDGs in the pancreas [32, 130]. Therefore, claims
cial means to dedifferentiate or to transdifferentiate to other cell that new beta cells derive exclusively from pre‐existing beta cells
types by transfection of cells with multiple transcription factors and based on insulin expression [24] ignore the fact that insulin
such as those identified by Takahashi and Yamanaka [125]. lineage tracing also labels stem/progenitors.
42:  Stem Cell‐Fueled Maturational Lineages in Hepatic and Pancreatic Organogenesis 535

The claim that mature liver cells are capable of extensive, Professor Alvaro was supported by FIRB grant #RBAP10Z7FS_
complete cell division is also not true except when these cells are 004 and by PRIN grant #2009X84L84_002. The study was also
transplanted into livers with the FAH mutation [131] or ones supported by Consorzio Interuniversitario Trapianti d’Organo,
expressing suicide transgenes [132], all of them having highly Rome, Italy.
artificial microenvironments that, theoretically, could be result-
Cornell University (Ithaca, NY). Dr. Sethupathy is funded in part
ing in reprogramming of the cells. Transplantation of mature
by a UNC genetics and molecular biology curriculum T32 training
cells into the livers of normal animals results in ­negligible cell
grant (T32‐GM‐007092–41); a grant (SRA‐60486) from Vesta
division. Transplantation into hosts with classic disease or regen-
Therapeutics (awarded to L.M.R); a grant (A11–0552) from Vesta
eration conditions (e.g. carbon tetrachloride or partial hepatec-
Therapeutics (awarded to P.S); and a grant (A16–0311) from the
tomy) results in transient, moderate amounts of cell division.
Fibrolamellar Cancer Foundation (awarded to P.S.).
These findings in experimental systems are in line with
those from clinical trials of mature hepatocyte transplantation
into patients with diverse liver dysfunctions or of islets into Intellectual property
patients with diabetes [32]. Transplantation of adult hepato-
cytes provides effects lasting only a few months. Transplanted Findings from the studies summarized in this review have been
islets tend to be insufficient to maintain long‐term glucose included in patent applications belonging to one or more institu-
homeostasis and exhibit limited proliferation in vivo. By con- tions including the following: University of North Carolina
trast, early findings of stem cell therapies for liver diseases (UNC) at Chapel Hill, Sapienza University in Rome, Italy, and
indicate that transplantation with BTSCs or HpSCs into the the Diabetes Research Institute (DRI) of the  University of
livers of patients with diverse liver dysfunctions results in Miami, Florida. The IP has been licensed to Vesta Therapeutics
long‐term effects over years with a steady improvement in (Bethesda, MD) for clinical uses and to  PhoenixSongs
liver functions [31]. Biologicals (PSB, Branford, CT) for non‐clinical, commercial
In summary, reports from clinical trials using stem cells uses. None of the authors have equity or a position in Vesta, and
­versus mature cells for cell therapies offer the most substantive none are paid consultants to the company. By contrast, LMR is
and incontrovertible evidence for rejection of the claim that one of the founders, is the ­scientific director, and does hold an
plasticity alone mediates regenerative responses in liver and equity position in PSB; to  date she has received no salary or
pancreas. We prefer the assumption that mechanisms for stem/ consulting fees for these efforts. Other than LMR’s connections
progenitors and their maturational lineages, along with minor with PSB, the authors declare no conflicts of interest. This
contributions from epigenetic phenomena, contribute to tissue review is derivative of  and updated from a book chapter in a
turnover and repair. book on stem cells and edited by Stewart Sell [45].

ACKNOWLEDGMENTS REFERENCES

Financial support   1. Turner, R., Lozoya, O., Wang, Y.F. et al. Hepatic stem cells and maturational
liver lineage biology. Hepatology, 2011;53:1035–45.
  2. Miyajima, A., Tanaka, M., and Itoh, T. Stem/progenitor cells in liver development,
UNC School of Medicine (Chapel Hill, NC). Funding derived homeostasis, regeneration, and reprogramming. Cell Stem Cell,
from Vesta Therapeutics (Bethesda, MD), a wholly owned 2014;14(5):561–74.
subsidiary of Toucan Capital Investments, and from the
­  3. Itoh, T. Stem/progenitor cells in liver regeneration. Review. Hepatology.
Fibrolamellar Carcinoma Foundation (Greenwich, CT). 2016;64(2):663–8.
 4. Carpino, G., Renzi, A., Franchitto, A. et  al. Stem/progenitor cell niches
Additional support was provided through discounted rates for
involved in hepatic and biliary regeneration. Review. Stem Cell International,
core services via federal funding of the cores: a microscopy ser- 2017:3658013.
vices laboratory in pathology and laboratory medicine core   5. Wang, Y., Cui, C., Miguez, P. et al. Lineage restriction of hepatic stem cells
facility grant (NIH P30DK34987), core director Victoria to mature fates is made efficient by tissue‐specific biomatrix scaffolds.
Madden, PhD; a histology core funded by the Center for Hepatology, 2011;53(1):293–305.
  6. Badylak, S.F., Taylor, D., and Uygun, K. Whole‐organ tissue engineering:
Gastrointestinal and Biliary Disease Biology (CGIBD) via an
decellularization and recellularization of three‐dimensional matrix scaffolds.
NIDDK Grant (DK34987); the Lineberger Cancer Center grant Review. Ann Rev Biomed Eng, 2011;13:27–53.
(NCI grant #CA016086); the Carolina Center for Genome   7. Baptista, P.M., Siddiqui, M.M., Lozier, G. et  al. The use of whole organ
Sciences (Katherine Hoadley, director), and the UNC Center for decellularization for the generation of a vascularized liver organoid.
Bioinformatics (Hemant Kelkar, director). Hepatology, 2011;53(2):604–17.
  8. Uygun, B., Soto‐Gutierrez, A., Yagi, H. et al. Organ re‐engineering through
Diabetes Research Institute (Miami, FL). Studies were funded development of a transplantatble recellularized liver graft using decellular-
by grants from NIH, the Juvenile Diabetes Research Foundation, ized liver matrix. Nat Med, 2010;16(7):814–20.
  9. Cardinale, V., Wang, Y., Gaudio, E. et  al. The biliary tree: a reservoir of
ADA, and the Diabetes Research Institute Foundation. multipotent stem cells. Nat Rev Gastroenterol Hepatol, 2012;9:231–40.
10. Zhang, D., Jiang, W., Liu, M. et al. Highly efficient differentiation of human
Sapienza University Medical Center (Rome, Italy). Professor
ES cells and iPS cells into mature pancreatic insulin‐producing cells. Cell
Gaudio was supported by research project grant from the Res, 2009;19(4):429–38.
University “Sapienza” of Rome and FIRB grant 11. Polo, J.M., Anderssen, E., Walsh, R.M. et al. A molecular roadmap of repro-
#RBAP10Z7FS_001 and by  PRIN grant #2009X84L84_001. gramming somatic cells into iPS cells. Cell, 2012;151(7):1617–32.
536 THE LIVER:  REFERENCES

12. Wang, Y., Qin, J., Wang, S. et al. Conversion of human gastric epithelial cells 38. Kubota, H. and Reid, L.M. Clonogenic hepatoblasts, common precursors
to multipotent endodermal progenitors using defined small molecules. Cell for hepatocytic and biliary lineages, are lacking classical major histocom-
Stem Cell, 2016;19(4):449–61. patibility complex class I antigens. Proc Natl Acad Sci (USA), 2000;97(22):
13. Rezania, A., Bruin, J.E., Riedel, M.J. et al. Maturation of human embryonic 12132–7.
stem cell‐derived pancreatic progenitors into functional islets capable of 39. Schmelzer, E., Wauthier, E., and Reid, L.M. Phenotypes of pluripotent
treating pre‐existing diabetes in mice. Diabetes, 2012;61:2016–29. human hepatic progenitors. Stem Cell, 2006;24(8):1852–8.
14. Zorn, A.M. and Wells, J.M. Molecular basis of vertebrate endoderm develop- 40. Sicklick, J.K., Li, Y.X., Melhem, A. et  al. Hedgehog signaling maintains
ment. Int Rev Cytol, 2007;259:49–111. resident hepatic progenitors throughout life. Am J Physiol Gastrointest Liver
15. McLin V.A., Zorn, A.M. Organogenesis: making pancreas from liver. Curr Physiol, 2006;290(5):G859–70.
Biol, 2003;13(3):R96–8. 41. Schmelzer, E., Zhang, L., Bruce, A. et al. Human hepatic stem cells from
16. Sinner, D., Rankin, S., Lee, M., and Zorn, A.M. SOX 17 and beta‐catenin fetal and postnatal donors. J Exp Med, 2007;204(8):1973–87.
cooperate to regulate the transcription of endodermal genes. Development, 42. Bonner‐Weir, S. and Sharma, A. Pancreatic stem cells. J Pathol, 2002;
2004;131(13):3069–80. 197(4):519–26.
17. Zaret, K. Developmental competence of the gut endoderm: genetic potentia- 43. Lysy, P.A., Weir, G.C., and Bonner‐Weir, S. Concise review: pancreas regen-
tion by GATA and HNF3/fork head proteins. Dev Biol, 1999;209(1):1–10. eration: recent advances and perspectives. Stem Cells Translat Med, 2012;
18. Wandzioch, E. and Zaret, K.S. Dynamic signaling network for the specifica- 1(2):150–9.
tion of embryonic pancreas and liver progenitors. Science, 2009;324(5935): 44. Jiang, W., Sui, X., Zhang, D. et al. CD24: a novel surface marker for PDX1‐
1707–10. positive pancreatic progenitors derived from human embryonic stem cells.
19. Tremblay, K.D. and Zaret, K.S. Distinct populations of endoderm cells con- Stem Cells, 2011;29(4):609–17.
verge to generate the embryonic liver bud and ventral foregut tissues. Dev 45. Furth, M.E., Wang, Y., Cardinale, V. et al. Stem cell populations giving rise
Biol, 2005;280(1):87–99. to liver, biliary tree and pancreas, in, The Stem Cells Handbook, 2nd edn, (ed.
20. Barker, N., van de Wetering, M., and Clevers, H. The intestinal stem cell. S. Sell), Springer Science Publishers, New York, 2013, pp. 75–126.
Genes Dev, 2008;22:1856–64. 46. Macdonald, J.M., Xu, A., Kubota, H. et  al. Liver cell culture and lineage
21. Barker, N., Huch, M., Kujala, P. et al. Lgr5(+ve) stem cells drive self‐renewal biology, in Methods of Tissue Engineering, (eds., A. Atala and R.P. Lanza),
in the stomach and build long‐lived gastric units in vitro. Cell Stem Cell, Academic Press, London, 2002, pp. 151–202.
2010;6:25–36. 47. Wang, Y., Yao, H., Barbier, C. et al. Paracrine signals from mesenchymal cell
22. Lange, A.W., Keiser, A.R., Wells, J.M., Zorn, A.M., and Whitsett, J.A. populations govern the expansion and differentiation of human hepatic stem
SOX17 promotes cell cycle progression and inhibits TGF‐beta/Smad3 sign- cells to adult liver fates. Hepatology, 2010;52(4):1443–54.
aling to initiate progenitor cell behavior in the respiratory epithelium. PLoS 48. Nusse, R. Wnt signaling and stem cell control. Cell Res, 2008;18:523–7.
One, 2009;4(5):e5711. 49. Roskams, T. and Desmet, V. Embryology of extra‐ and intrahepatic bile
23. Snyder, J.C., Teisanu, R.M., and Stripp, B.R. Endogenous lung stem cells ducts, the ductal plate. Anat Rec, 2008;291(6):628–35.
and contribution to disease. J Pathol, 2009;217:254–64. 50. Carpentier, R., Español‐Suñer, R., van Hul, N. et al. Embryonic ductal plate
24. Dor, Y., Brown, J., Martinez, O.I., and Melton, D.A. Adult pancreatic beta‐ cells give rise to cholangiocytes, periportal hepatocytes, and adult liver pro-
cells are formed by self‐duplication rather than stem‐cell differentiation. genitor cells. Gastroenterology, 2011;141(4):1432–8.
Nature, 2004;429(6987):41–6. 51. Zhang, L., Theise, N., Chua, M., and Reid, L.M. The stem cell niche of
25. Houbracken, I. and Bouwens, L. The quest for tissue stem cells in the pan- human livers: symmetry between development and regeneration. Hepatology,
creas and other organs, and their application in beta‐cell replacement. Rev 2008;48(5):1598–607.
Diabet Stud, 2010;7:112–23. 52. Saxena, R. and Theise, N. Canals of Hering: recent insights and current
26. Xu, X., D’Hoker, J., Stange, G. et  al. Beta cells can be generated from knowledge. Semin Liver Dis,2004;24(1):43–8.
endogenous progenitors in injured adult mouse pancreas. Cell, 2008; 53. Theise, N.D., Saxena, R., Portmann, B.C. et al. The canals of Hering and
132(2):197–207. hepatic stem cells in humans. Hepatology, 1999;30(6):1425–33.
27. Wang, Y., Lanzoni, G., Carpino, G. Biliary tree stem cells, precursors to pan- 54. Strobel, O., Rosow, D.E., Rahaklin, E.Y. J. et al. Pancreatic duct glands are
creatic committed progenitors: evidence for life‐long pancreatic organogen- distinct ductal compartments that react to chronic injury and mediate Shh‐
esis. Stem Cells, 2013;31(9):1966–79. induced metaplasia. Gastroenterology, 2010;138:1166–77.
28. Carpino, G., Renzi, A., Cardinale, V. et  al. Progenitor cell niches in the 55. Kushner, J.A., Weir, G.C., and Bonner‐Weir, S. Ductal origin hypothesis of
human pancreatic duct system and associated pancreatic duct glands: an ana- pancreatic regeneration under attack. Cell Metab, 2010;11(1):2–3.
tomical and immunophenotyping study. J Anat, 2016;228(3):474–86. 56. Bonner‐Weir, S., Tosch, E., Inada, A. et al. The pancreatic ductal epithelium
29. Carpino, G., Cardinale, V., Gentile, R. et al. Evidence for multipotent endo- serves as a potential pool of progenitor cells. Pediatric Diabetes, 2004;
dermal stem/progenitor cell populations in human gallbladder. J Hepatology, 5:16–22.
2014;60(6):1194–2020. 57. McClelland, R., Wauthier, E., Uronis, J., and Reid, L.M. Gradient in extra-
30. Cardinale, V., Wang, Y., Carpino, G. Multipotent stem cells in the extrahe- cellular matrix chemistry from periportal to pericentral zones: regulation of
patic biliary tree give rise to hepatocytes, bile ducts and pancreatic islets. hepatic progenitors. Tissue Eng,2008;14(1):59–70.
Hepatology, 2011;54(6):2159–72. 58. Kubota, H., Yao, H., and Reid, L.M. Identification and characterization of
31. Lanzoni, G., Cui, C., Oikawa, T. et al. Clinical programs of stem cell thera- vitamin A‐storing cells in fetal liver. Stem Cell, 2007;25:2339–49.
pies for liver and pancreas. Stem Cells, 2013;31(10):2047–60. 59. Lesley, J., Hascall, V.C., Tammi, M., and Hyman, R. Hyaluronan binding by
32. Lanzoni, G., Cardinale, V., Carpino, G. The hepatic, biliary and pancreatic cell surface CD44. J Biol Chem, 2000;275(35):26967–75.
network of stem/progenitor cells niches in humans: a new reference frame 60. Hayes, A., Tudor, D., Nowell, M., Caterson, B., and Hughes, C. Chondroitin
for disease and regeneration. Hepatology, 2016;64(1):277–86. sulfate sulfation motifs as putative biomarkers for isolation of articular carti-
33. Alvaro, D. and Gaudio, E. Liver capsule. Biliary tree stem cell subpopula- lage progenitor cells. J Histochem Cytochem, 2007;56:125–38.
tions. Hepatology, 2016;64(2)644. 61. McClelland R., Wauthier E., Uronis J, et al. Gradients in the liver’s extracel-
34. Carpino, G., Cardinale, V., Onori, P. et al. Biliary tree stem/progenitor cells lular matrix chemistry from periportal to pericentral zones: influence on
in glands of extrahepatic and intraheptic bile ducts: an anatomical in situ human hepatic progenitors. Tissue Eng Part A, 2008;14(1):59–70.
study yielding evidence of maturational lineages. J Anat, 2012;220(2): 62. Zern, M. and Reid, Z. Extracellular Matrix Chemistry and Biology,
186–99. Academic Press, New York, 1993.
35. Leeson, T.S. and Leeson, C.R. The fine structure of Brunner’s glands in man. 63. Aruffo, A., Stamenkovic, I., Melnick, M., Underhill, C.B., and Seed, B.
J Anat, 1968;103(2):263–76. CD44 is the principal cell surface receptor for hyaluronate. Cell, 1990;
36. Krause, W.J. Brunner’s glands: a structural, histochemical and pathological 61:1303–13.
profile. Prog Histochem Cytochem, 2000;35(4):259–367. 64. Nakanuma, Y., Hoso, M., Sanzen, T., and Sasaki, I.M. Microstructure and
37. Semeraro, R., Carpino, G., Cardinale, V. et al. Multipotent stem/progenitor development of the normal and pathologic biliary tract in humans, including
cells in the human foetal biliary tree. J Hepatol, 2012;220(2):186–99. blood supply. A review. Microsc Res Tech, 1997;15(38):552–70.
42:  Stem Cell‐Fueled Maturational Lineages in Hepatic and Pancreatic Organogenesis 537

65. Nakanuma, Y., Katayanagi, K., Terada, T., and Saito, K. Intrahepatic peribil-  91. McClelland, R., Wauthier, E., Zhang, L. et  al. Ex vivo conditions for
iary glands of humans. I. Anatomy, development and presumed functions. A self‐replication of human hepatic stem cells. Tissue Eng Part C Methods,
review. J Gastroenterol Hepatol, 1994;9(1):75–9. 2008;14(4):314–51.
66. Nakanuma, Y., Sasaki, M., Terada, T., and Harada, K. Intrahepatic peribiliary   92. Schmelzer, E. and Reid, L.M. Telomerase activity in human hepatic stem
glands of humans. II. Pathological spectrum. Review. J Gastroenterol cells, hepatoblasts and hepatocytes from neonatal, pediatric, adult and geri-
Hepatol, 1994;9(1):80–6. atric donors. Eur J Hepatol Gastroenterol, 2009;21(10):1191–8.
67. Portulano, C., Paroder‐Belenitsky, M., and Carrasco, N. The Na+/I‐ sym-   93. Oikawa, T., Kamiya, A., Zeniya, M. et  al. Sal‐like protein 4 (SALL4), a
porter (NIS): mechanism and medical impact. Endocrine Rev,2014;35(1): stem cell biomarker in liver cancers. Hepatology, 2012;57(4):1469–83.
106–49.   94. Turner, R., Gerber, D., and Reid, L.M. Transplantation of cells from solid
68. Lozoya, O.A., Wauthier, E., Turner, R. et  al. Regulation of hepatic stem/ organs requires grafting protocols. Transplantation, 2010;90:807–10.
progenitor phenotype by microenvironment stiffness in hydrogel models of   95. Hodgman, C.D. Handbook of Chemistry and Physics, 37th ed, Chemical
the human liver stem cell niche. Biomaterials, 2011;32(30):7389–402. Rubber Publishing Co., 1955–1956, pp. 3156.
69. Wauthier, E., McClelland, R., Turner, W. et al. Hepatic stem cells and hepa-   96. Turner, W.S., Schmelzer, E., McClelland, R., Wauthier, E., Chen, W., and
toblasts: identification, isolation and ex vivo maintenance. Methods Cell Reid, L.M. Human hepatoblast phenotype maintained by hyaluronan
Biol, 2008;86:137–225. hydrogels. J Biomed Mater Res, 2007;82(1):156–68.
70. Cardinale, V., Carpino, G., Gentile, R. et al. Transplantation of human fetal   97. Turner, W.S., Seagle, C., Galanko, J. et  al. Metabolomic footprinting of
biliary tree stem/progenitor cells into two patients with advanced liver cir- human hepatic stem cells and hepatoblasts cultured in engineered hyaluro-
rhosis. BMC Gastroenterology, 2014;14:204. nan‐matrix hydrogel scaffolds. Stem Cell, 2008;26:1547–55.
71. Michalopoulos, G.K. Liver regeneration: alternative epithelial pathways. Int   98. Turner, R., Mendel, G., Wauthier, E., Barbier, C., and Reid, L.M. Hyaluronan‐
J Biochem Cell Biol, 2011;43:173–9. supplemented buffers preserve adhesion mechanisms facilitating cryopreser-
72. Cohen, D.E. and Melton, D. Turning straw into gold: directing cell fate for vation of human hepatic stem/progenitor cells. Cell Transplant, 2012;21(10):
regenerative medicine. Nat Rev Genet, 2011;12:243–52. 2257–66.
73. Puppi, J., Strom, S.J., Hughes, R.D. et  al. Improving the techniques for   99. Li, W. and Ding, S. Small molecules that modulate embryonic stem cell fate
human hepatocyte transplantation: report from a consensus meeting in and somatic cell reprogramming. Trends Pharmacol Sci,2010;31(1):36–45.
London. Cell Transplant, 2012;21(1)1–10. 100. Turner, R., Wauthier, E., Lozoya, O. et  al. Successful transplantation of
74. Washburn, M.L., Bility, M.T., Zhang, L. et al. A humanized mouse model to human hepatic stem cells with restricted localization to liver using hyaluro-
study hepatitis C virus infection, immune response, and liver disease. nan grafts. Hepatology, 2013;57:775–84.
Gastroenterology, 2011;140(4):1334–44. 101. Weber, A., Mahieu‐Caputo, D., Michelle Hadchoue, M., and Franco, D.
75. Fukumitsu, K., Yagi, H., and Soto‐Gutierrez, A. Bioengineering in organ Hepatocyte transplantation: studies in preclinical models. J Inherit Metab
transplantation: targeting the liver. Transplant Proc,2011;43:2137–8. Dis, 2006;29(2–3):436–41.
76. Gerlach, J.C. Bioreactors for extracorporeal liver support. Cell Transplant, 102. Weber, A., Groyer‐Picard, M.T., Franco, D., and Dagher, I. Hepatocyte
2006;15(1):S91–103. transplantation in animal models. Liver Transplant, 2009;15(1):7–14.
77. Russo, F.P. and Parola, M. Stem and progenitor cells in liver regeneration and 103. Kinoshita, T. and Miyajima, A. Cytokine regulation of liver development.
repair. Cytotherapy, 2011;13:135–44. Biochim Biophys Acta, 2002;1592(3):303–12.
78. Parveen, N., Aleem, A.K., Habeeb, M.A., and Habibullah, C.M. An update 104. Kamiya, A., Kinoshita, T., Ito, Y. et al. Fetal liver development requires a
on hepatic stem cells: bench to bedside. Curr Pharm Bioltechnol, paracrine action of oncostatin M through the gp130 signal transducer.
2011;12(2):226–30. EMBO J, 1999;18(8):2127–36.
79. Duncan, A., Hickey, R.D., Paulk, N.K. et  al. Ploidy reductions in murine 105. Rojkind, M., Gatmaitan, Z., Mackensen, S., Giambrone, M.A., Ponce, P., and
fusion‐derived hepatocytes. PloS Genet, 2009;5(2):e1000385. Reid, L.M. Connective tissue biomatrix: its isolation and utilization for long‐
80. Tanaka, M., Itoh, T., Tanimizu, N., and Miyajima, A. Liver stem/progenitor term cultures of normal rat hepatocytes. J Cell Biol, 1980;87(1):255–63.
cells: their characteristics and regulatory mechanisms. J Biochem, 2011; 106. Reid, L.M., Gaitmaitan, Z., Arias, I., Ponce, P., and Rojkind, M. Long‐term
149(3):231–9. cultures of normal rat hepatocytes on liver biomatrix. Ann N T Acad Sci,
81. Formin, M.E., Tai, L.K., Bárcena, A., and Muench, M.O. Coexpression of 1980;349:70–6.
CD14 and CD326 discriminate hepatic precursors in the human fetal liver. 107. Purushothaman, A., Hurst, D.R., Pisano, C., Mizumoto, S., Sugahara, K.,and
Stem Cells, 2011;20(7):1247–57. Sanderson, R.D. Heparanase‐mediated loss of nuclear syndecan‐1 enhances
82. Thorgeirsson, S., Factor, V., and Grisham, J. Early activation and expansion histone acetyltransferase (HAT) activity to promote expression of genes that
of hepatic stem cells, in Handbook of Stem Cells, 2nd edn, (eds. R. Lanza, H. drive an aggressive tumor phenotype. J Biol Chem, 2011;286(35):30377–83.
Blau, D.A. Melton et al.) Elsevier, New York, 2004, pp. 497–512. 108. Capila, I. and Linhardt, R.J. Heparin ± protein interactions. Angewandte
83. Navarro‐Alvarez, N., Soto‐Gutierrez, A., and Kobayashi, N. Hepatic stem Chemie Int Ed, 2002;41:390–412.
cells and liver development. Method Mol Biol, 2010;640:181–236. 109. Itoh, T. and Miyajima, A. Liver regeneration by stem/progenitor cells.
84. Vessey, C.J. and de la Hall, P.M. Hepatic stem cells: a review. Pathology, Hepatology, 2014;58(4):1617–26.
2001;33(2):130–41. 110. Michalopoulos, G.K. and DeFrances, M.C. Liver regeneration. Review.
85. Reid, L.M., Fiorino, A.S., Sigal, S.H., Brill, S., and Holst, P.A. Extracellular Science, 1997;276(5309):60–6.
matrix gradients in the space of Disse: relevance to liver biology. Hepatology, 111. Chiang, J.Y. Bile acid regulation of gene expression: roles of nuclear hor-
1992;15(6):1198–203. mone receptors. Endocrine Rev, 2002;23(4):443–63.
86. Kopp, J.L., Grompe, M., and Sanders, M. Stem cells versus plasticity in liver 112. Perelman, A. and Brandan, E. Different membrane‐bound forms of acetyl-
and pancreas regeneration. Nat Cell Biol, 2016;18(3):238–45. cholinesterase are present at the cell surface of hepatocytes. Eur J Biochem,
87. Khan, A.A., Parveen, N., Mahaboob, V.S. et al. Management of hyperbili- 1989;182(1):203–7.
rubenemia in biliary atresia by hepatic progenitor cell transplantation 113. Alvaro, D., Alpini, G., Jezequel, A.M. et al. Role and mechanisms of action
through hepatic artery: a case report. Transplant Proc, 2008;40(4):1153–5. of acetylcholine in the regulation of rat cholangiocyte secretory functions.
88. Khan, A.A., Parveen, N., Mahaboob, V.S. et al. Treatment of Crigler‐Najjar J Clin Invest, 1997;100(6):1349–62.
syndrome type 1 by hepatic progenitor cell therapy: a simple procedure for 114. LeSage, E.G., Alvaro, D., Benedetti, A. et  al. Cholinergic system modu-
hyperbilirubinemia. Transplant Proc, 2008;40(4):1148–50. lates growth, apoptosis, and secretion of cholangiocytes from bile duct‐
89. Khan, A.A., Shaik, M.V., Parveen, N. et al. Human fetal liver‐derived stem ligated rats. Gastroenterology, 1999;117(1):191–9.
cell transplantation as supportive modality in the management of end‐stage 115. Vongchan, P., Warda, M., Toyoda, H., Toida, T., Marks, R.M., and Linhardt,
decompensated liver cirrhosis. Cell Transplant, 2010;19(4):409–18. R.J. Structural characterization of human liver heparan sulfate. Biochim
90. Stachelscheid, H., Urbaniak, T., Ring, A., Spengler, B., Gerlach, J.C., and Biophys Acta, 2005;1721(1–3):1–8.
Zeilinger, K. Isolation and characterization of adult human liver progenitors 116. Fujita, M., Spray, D.C., Choi, H. et al. Glycosaminoglycans and proteogly-
from ischemic liver tissue derived from therapeutic hepatectomies. Tissue cans induce gap junction expression and restore transcription of tissue‐­
Eng Part C Methods, 2009;15(7):1633–43. specific mRNAs in primary liver cultures. Hepatology, 1987;7(1):1S–9S.
538 THE LIVER:  REFERENCES

117. Spray, D.C., Fujita, M., Saez, J.C. et al. Proteoglycans and glycosamino- 125. Takahashi, K. and Yamanaka, S. A developmental framework for induced
glycans induce gap junction synthesis and function in primary liver cul- pluripotency. Review. Development, 2015;142(19):3274–85.
tures. J Cell Biol, 1987;105(1):541–51. 126. Dorrell, C., Tarlow, B., Wang, Y. et  al. The organoid‐initiating cells in
118. Masyuk, A.I., Masyuk, T.V., and LaRusso, N.F. Cholangiocyte primary mouse pancreas and liver are phenotypically and functionally similar. Stem
cilia in liver health and disease. Dev Dynam, 2008;237:2007–12. Cell Res, 2014;13(2):275–83.
119. Huang, B.Q., Masyuk, T.V., Muff, M.A., Tietz, P.S., Masyuk, A.I., and 127. Lu, W.Y., Bird, T.G., Boulter, L. et al. Hepatic progenitor cells of biliary
Larusso, N.F. Isolation and characterization of cholangiocyte primary cilia. origin with liver repopulation capacity. Nat Cell Biol, 2015;17(8):971–83.
Am J Physiol Gastrointest Liver Physiol, 2006;291(3):G500–9. 128. Tarlow, B.D., Pelz, C., Naugler, W.E. et al. Bipotential adult liver progeni-
120. Cervantes, S., Lau, J., Cano, D.A., Borromeo‐Austin, C., and Hebrok, M. tors are derived from chronically injured mature hepatocytes. Cell Stem
Primary cilia regulate Gli/Hedgehog activation in pancreas. Proc Nat Acad Cell, 2014;15(5):605–18.
Sci USA, 2010;107(22):10109–14. 129. Fan, B., Malato, Y., Calvisi, D.F. et al. Cholangiocarcinomas can originate
121. Michalopoulos, G.K. and Appasamy, R. Metabolism of HGF‐SF and its from hepatocytes in mice. J Clin Invest, 2012;122(8):2911–5.
role in liver regeneration. Review. EXS, 1993;65:275–83. 130. Smukler, S.R., Arntfield, M.E., Razavi, R. et  al. The adult mouse and
122. Liu, H., Di Cunto, F., Imarisio, S., and Reid, L.M. Citron kinase is a cell human pancreas contain rare multipotent stem cells that express insulin.
cycle‐dependent, nuclear protein required for G2/M transition of hepato- Cell Stem Cell, 2011;8:281–93.
cytes. J Biol Chem, 2003;278(4):2541–8. 131. Grompe, M. and Strom, S. Mice with human livers. Gastroenterology,
123. Sigal, S.H., Rajvanshi, P., Gorla, G.R. et al. Partial hepatectomy‐induced 2013;145(6):1209–14.
polyploidy attenuates hepatocyte replication and activates cell aging 132. Rhim, J.A., Sandgren, E.P., Palmiter, R.D., and Brinster, R.L. Complete
events. Am J Physiol Gastrointest Liver Physiol, 1999;276(5):G1260–72. reconstitution of mouse liver with xenogeneic hepatocytes. Proc Nat Acad
124. Huch, M. and Dollé, L. The plastic cellular states of liver cells: are EpCAM Sci USA, 1995;92(11):4942–6.
and Lgr5 fit for purpose? Review. Hepatology, 2016:64(2);652–62.
Developmental Morphogens
43 and Adult Liver Repair
Mariana Verdelho Machado1 and Anna Mae Diehl2
1
Gastroenterology and Hepatology Department, Hospital de Santa Maria, CHLN, Lisbon, and Faculty of
Medicine, Lisbon University, Lisbon, Portugal
2
School of Medicine, Duke University, Durham, NC, USA

HEDGEHOG SIGNALING interactions regulate Smo activity by controlling cholesterol


modification of Smo. Smo is directly activated by binding cho-
lesterol. Patch suppresses this cholesterol modification of Smo
Pathway overview in the absence of Hh ligands, and this inhibition is relieved when
The canonical Hh pathway is a conserved, highly complex Hh binds to Patch [9]. The Hh–Patch complex is subsequently
signaling cascade with four fundamental components: (i) the internalized and degraded [10]. Three Hh co‐receptors, CAM‐
ligand Hedgehog, (ii) the receptor Patched (Patch), (iii) the related downregulated by oncogene (Cdo), brother of Cdo
­signal transducer Smoothened (Smo), and (iv) the effector (Boc), and growth‐arrest‐specific (GAS)‐2, potentiate Hh
transcription factor, Gli (Figure 43.1). Canonical Hh signaling ­signaling by enhancing Hh‐Ptch interaction [11]. Conversely,
occurs along the primary cilium (PC) with components of the Hh Hhip, a soluble Hh receptor, inhibits Hh signaling by preventing
pathway concentrating in PC [1] and a complex PC trafficking Hh‐Patch binding [12].
system regulating the interaction of Hh pathway components to Smo is a transmembrane G‐protein coupled receptor that
enhance, or block, the Hh‐initiated signal [2]. mediates activation of Gli transcription factors in Hh‐responsive
Hh is a protein produced as a 45 kDa precursor that under- cells. Gli proteins promote transcription of several genes
goes proteolytic processing in the endoplasmic reticulum (ER) important in the regenerative/repair process, including vascular
[3] and subsequent lipid modification to acquire cholesterol and endothelial growth factors, angiopoietin‐1 and ‐2, snail, twist‐2,
palmitoyl groups [4, 5]. Hh is secreted into the extracellular α‐smooth muscle actin, vimentin, nanog, sox2 and sox9 [8].
space, diffusing away from the ligand‐producing cell to bind to In the absence of Hh, Smo activity is repressed by Patch, and
other cells whereby it determines their fate according to the con- Gli binds to fused kinase (Fu), suppressor of fused (Sufu) and
centration and duration of exposure [6]. Extracellular matrix Costal‐2, to form a suppressor protein complex which prevents
proteins, such as proteoglycans, modulate the diffusion of Hh Gli from entering the nucleus [13]. Arrested in the cytoplasm by
through the extracellular space and thus, regulate the concentra- the suppressor protein complex, Gli is sequentially phosphoryl-
tion of Hh to which target cells are exposed [7]. Mammals have ated by protein kinase A (PKA), glycogen synthase kinase‐3
three different Hh proteins: Sonic (Shh), Indian (Ihh), and (GSK3), and calmodulin kinase‐1 (CK1). Phosphorylated Gli
Desert (Dhh) hedgehog. The three ligands similarly activate the then binds to β‐transducin repeat containing protein (βTrCp)
Hh pathway in Hh‐responsive cells, however their expression is and the Gli‐βTrCp complex is targeted to the proteasome where
differently regulated. While Shh and Ihh are widely expressed, Gli can be either degraded entirely or processed to generate a
Dhh is thought to be expressed mainly in the nervous system truncated transcription repressor (Gli‐R) [14]. When Hh binds to
and testis [8]. Patch, Smo is de‐repressed and activated Smo dissociates Gli
Patch, the Hh receptor, is a protein with 12 transmembrane from the suppressor protein complex, preventing Gli phospho-
domains. When Hh ligands are absent, Patch localizes to the PC rylation and subsequent degradation. This enables full‐length Gli
and constitutively inhibits the Hh pathway by blocking Smo, the to move to the nucleus where it acts as a transcription factor.
signal transducer protein, from being activated and entering Mammals have three known Gli proteins: Gli1, ‐2 and ‐3.
the  PC. When Hh ligand binds to Patch, these inhibitory Gli1 does not undergo proteosomal degradation and hence,
actions of Patch are relieved and Smo becomes active. Hh‐Patch remains untruncated and always promotes transcription. Gli1 is

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
540 THE LIVER:  HEDGEHOG SIGNALING

PC PC

Hh
Smo
Patch
Plasma Plasma
membrane membrane

CKS
PKA
GKA
Phosphorylation CKS
PKA
Ubiquitination Gli
GKA
Smo Gli
SCF-β-
TrCP
Partial Processing
Hh-Patch
Proteasome complex

Nucleus Nucleus
STOP
Gli-R PATHWAY Gli-A PATHWAY
OFF ON

Figure 43.1  The Hedgehog signaling pathway. (a) Pathway off: Patched (Patch) blocks the entry of Smoothened (Smo) into the primary cilium
(PC), which represses Smo activity and allows sequential phosphorylation of Gli by several kinases: protein kinase A (PKA), glycogen synthase‐3β
(GSK3β) and casein kinase‐1 (CK1). Phosphorylated Gli undergoes ubiquitination by Skp‐Cullin‐F‐box (SCF) protein/β‐transducing repeat con-
taining protein (TrCP), which primes Gli to limited degradation in the proteasome. Truncated Gli (Gli‐R) is a repressor of gene transcription. (b)
Pathway on: Hedgehog binds to Patch and removes it from the PC, which allows Smo entering into the PC and Smo activation. Active Smo blocks
phosphorylation and subsequent degradation of Gli. Full length Gli translocates to the nucleus and promotes transcription of several target genes.

an important target gene for Gli2 [15]. Full‐length Gli2 accumu- Liver development
lates when Smo is activated because activated Smo protects Gli2
from proteosomal degradation. When Smo is inactive, both Gli2 The exact contribution of the Hh pathway to liver embryogene-
and Gli3 are targeted to the proteasome; Gli2 is usually fully sis remains unknown although it is clear that the pathway is cru-
degraded but Gli3 is frequently partially processed to a trun- cial for differentiation of the liver bud from the ventral foregut
cated form that represses transcription [16]. Hence, Gli3 can act endoderm [27]. Later in development, activation and repression
as either a transcriptional repressor (when Smo is inactive) or as of Hh signaling control hepatoblast fate, with activation stimu-
an activator of transcription (when activated Smo protects it lating a proliferative undifferentiated phenotype and repression
from proteosomal degradation). In contrast, Gli1 and Gli2 act permitting hepatoblasts to differentiate into mature hepatocytes
predominantly as transcription promoters. Emerging evidence [28]. Hedgehog is also crucial for the proper development of the
suggests that there may be a “Gli code” whereby the different muscle layer and functioning of the gallbladder and disruption
Gli factors interact to control their expression. According to this of the Hh pathway at that point in development abrogates the
model, Gli3 and Gli1 promote each other’s expression when protective effect of the gallbladder against the toxic effects of
Gli2 is low but become mutually antagonistic when Gli2 accu- bile acids on the biliary system, which may contribute to the
mulates [17]. development of biliary atresia [29].
Besides the canonical Hh pathway, there are also two known
types of non‐canonical Hh signaling. Type 1 non‐canonical Hh Liver neoplasia
signaling depends on Patch but is Smo‐independent. In the
Primary liver cancers
absence of Hh, Patch has direct pro‐apoptotic and antiprolifera-
tive effects, by activating caspase‐3 [18] and preventing nuclear Hedgehog pathway activation has been demonstrated in various
localization of cyclin D [19], respectively. Both effects of Patch types of primary liver cancer, including hepatocellular carci-
are lost when Hh binds to Patch. Type 2 noncanonical Hh sign- noma [30, 31], cholangiocarcinoma [32], and fibrolamellar
aling depends on Smo but it does not require PC [20]. This non- hepatocellular carcinoma [33]. In all of these cancers, the level
canonical signaling depends on the Gαi activity of Smo that of pathway activity correlates with worse clinical outcomes and
directly regulates metabolism (e.g. it promotes a Warburg‐like reduced cancer‐free, and overall, survival [31]. Relatively
effect promoting glycolysis in muscle, adipose tissue, and uncommon genetic mechanisms, and more prevalent epigenetic
myofibroblasts [21, 22]), proliferation, calcium flux, and migra- mechanisms, contribute to hedgehog dysregulation in liver can-
tion (in myofibroblasts and endothelial cells [23, 24]). cers. For example, liver cancer cell growth is increased both by
Additionally, Gli signaling can occur in the absence of Hh via a certain mutations of Smo that activate it constitutively [30], and
process that also appears to be Patch and Smo‐independent, as by hypermethylation of Hh signaling inhibitors which epige-
demonstrated by evidence that Gli induction is a direct down- netically suppresses expression of factors that typically con-
stream consequence of transforming growth factor (TGF) beta strain pathway activity [34]. Dysregulated Hh signaling has
and RAS signaling [25, 26]. been demonstrated in diverse cell types of liver cancers,
43:  Developmental Morphogens and Adult Liver Repair 541

including the malignant epithelial cells themselves, cancer‐ Hepatocytes produce Hedgehog ligands
associated fibroblasts (CAFs), and tumor‐associated mac-
One mechanism that may explain how liver injury is tightly
rophages (TAMs). Hedgehog signaling in CAFs promotes
coupled to Hh pathway activation was suggested by the discov-
fibrogenesis and generation of factors that maintain a supportive
ery that adult hepatocytes produce and release large amounts of
niche for cancer stem‐like cells, responses that likely foster can-
Shh and Ihh ligands when subjected to challenges that promote
cer growth [35]. Similarly, Hh signaling in TAMs promotes
ER stress and apoptosis [49, 50]. In such circumstances,
immune tolerance that permits cancer cells to escape immune
­biologically‐active Hh ligands appear to localize in hepatocyte‐
surveillance mechanisms [36]. On the other hand, Hh signaling
derived exosomes and microparticles [51]. Once released into
also promotes vasculogenesis and this might enhance chemo-
the hepatic microenvironment, bile, and blood, these mem-
therapy bioavailability [37, 38]. These opposing actions of Hh
brane‐bound Hh ligands are capable of activating Hh signaling
make it difficult to predict if treatments that inhibit Hh signaling
in local and distant Hedgehog‐responsive cells and hence,
might be beneficial in primary liver cancers. Enthusiasm has
function as hepatocyte‐derived hormones [51]. In injured liv-
also been dampened by inconsistent efficacy and relatively
ers, increased exposure to Hh ligands triggers changes in pro-
prevalent toxicities observed when Hedgehog was inhibited in
liferation, viability, and differentiation of various types of local
other cancers [39]. Clinicaltrials.gov lists ongoing trials of
target cells to orchestrate reconstruction of adult liver tissue via
hedgehog inhibitors in hepatocellular carcinoma.
regenerative mechanisms that resemble those involved in fetal
liver development [47]. The impact of liver‐derived Hh ligands
Hepatic adenomas on Hh signaling in extrahepatic tissues is less studied but might
A subtype of hepatic adenomas was recently identified in which be significant because Hh signaling suppresses adipogenesis
a chromosomal deletion results in constitutive activation of Gli1 in  white adipose depots [52], promotes vasculogenesis [53],
in hepatocytes. Because Gli1 is both a downstream target and a and modulates immune function [54, 55], and advanced liver
proximal effector of the Hh signaling pathway, neoplasia is disease is characterized by cachexia, vascular remodeling, and
attributed to dysregulated Hedgehog signaling. Hedgehog sign- immune dysfunction.
aling drives vasculogenesis during development and these ade-
nomas are particularly prone to clinically‐significant hemorrhage
Hepatic stellate cells both produce and respond
[40]. Interestingly, the Hedgehog‐high adenoma subclass is also
to Hedgehog ligands
marked by high expression of arginosuccinate synthase 1, a urea
cycle enzyme [41]. This finding supports emerging evidence The major fibrogenic cell type in liver, the hepatic stellate cell
that Hh signaling activity in hepatocytes may regulate metabolic (HSC), produces and is highly responsive to, Hedgehog ligands
zonation in healthy adult liver [42]. According to this model, Hh [56]. Coupled with the fact that HSCs live in the space of Disse
activity is higher in periportal hepatocytes (which use the urea immediately adjacent to hepatocytes and sinusoidal endothelial
cycle to detoxify ammonia) than in perivenous hepatocytes cells (other liver‐resident cell types that produce and/or respond
(which exhibit negligible urea cycle activity) [41]. Conversely, to Hh ligands), HSC are positioned to be a key node in Hh‐
Wnt signaling is known to be higher in zone three hepatocytes regulated regenerative responses. HSCs are also capable of
(which synthesize glutamine to scavenge ammonia) than in ­noncanonical Hh signaling because they express various G
zone one hepatocytes (which do not) [43]. Differential Hh and ­protein‐coupled receptors that regulate Smo independently of
Wnt activity also occurs along the intestinal crypt–villus axis. In Hedgehog–Patched interaction [22]. In addition, they can acti-
that tissue, Hh and Wnt appear to be mutually antagonistic, with vate Gli2 via morphogen‐driven mechanisms that do not require
Hh restricting propagation of Wnt signals and thereby restrict- Smo [57]. Such signaling flexibility supports the concept that
ing the stem cell compartment to the crypt [44]. net Hh pathway activity must be tightly regulated in HSC
in  order to assure optimal liver repair. Indeed, Hh pathway
activation is necessary for quiescent HSC to become and
Liver repair remain myofibroblasts and in mice, liver fibrosis is inhibited by
The adult liver is designed to resist injury. Hence, hepatocyte various approaches that inhibit Smo and/or suppress Gli1/2
attrition is mainly due to natural senescence and liver mass is activity [56, 58, 59]. Conversely, simply overexpressing Shh
easily maintained by a small subpopulation of hepatocytes that ligand in hepatocytes is sufficient to induce progressive liver
are more proliferative than the vast majority of adult hepato- fibrosis in mice [60].
cytes, which are devoted to performing various liver‐specific
functions [45]. In contrast, liver injuries that kill hepatocytes
Liver sinusoidal cells (LSECs) respond to Hh ligands
dramatically increase the demand for hepatocyte replacement
and this triggers multifaceted wound healing responses that aim Liver sinusoidal cells (LSECs) respond to Hh ligands with a
to restore functional hepatocyte mass [46]. Upregulation of Hh repertoire of behaviors that results in blood vessel formation
signaling has been demonstrated in injured livers, regardless of and growth during liver injury. In LSECs, increased Hh
injury etiology [47]. Further, the level of pathway activity typi- ­signaling induces loss of fenestrae (capillarization) [61] and
cally parallels the severity of liver damage and fibrosis [48]. may promote portal hypertension/portal–systemic shunting
These observations suggest that Hh critically regulates how [62]. Liver endothelial cells can also produce Hh ligands
adult livers respond to injury. The mechanisms involved are which modulate liver repair via autocrine and paracrine
summarized below. mechanisms [61].
542 THE LIVER:  NOTCH SIGNALING

The innate immune system is also highly engaging their receptor, Patched [77]. Interestingly, Smo is
Hedgehog‐responsive directly inhibited by certain other lipoprotein‐associated lipids
[78] and thus, the healthy liver may modulate Hh signaling by
T cells and natural killer T (NKT) cells require Hh signaling to
varying lipoprotein particle composition. Hedgehog ligand pro-
remain viable [54]. Hedgehog signaling also promotes the syn-
ducing cells that traffic into and out of the liver (e.g. immune
thesis of certain chemotactic factors for NKT cells (CXCL16),
cells, bone marrow‐derived mononuclear cells) might be another
and macrophages (CCL2) [63, 64]. Further, Hedgehog‐sensitive
source of Hh ligands for healthy hepatocytes. Hedgehog ligands
mechanisms modulate immune polarization of T cells, NKT
might also be provided by subpopulations of liver‐resident cells
cells, and macrophages: Hh signaling typically promotes Th2/
that are capable of generating Hh ligands (e.g. HSCs, LSECs,
M2 polarization that enhances immune tolerance [55]. Tolerance
ductal cells). Signaling downstream of Hedgehog–Patched
is mediated, in part, via increased production of cytokines that
could also be modulated by non‐canonical Hh signaling that
also induce fibrogenesis (e.g. IL‐4, IL‐13, TGFβ) by re‐enforc-
operates independently of Patched or Smo, affording hepato-
ing Hh signaling in hepatic stellate cells [54].
cytes multiple methods to titrate pathway activity. Indeed, Hh
signaling is likely to be tightly regulated in most cells because
Ductal cells produce and respond to Hh ligands
pathway activity critically modulates DNA methylation and
Paracrine Hh signaling between ductal cells and neighboring chromatin remodeling by regulating one carbon metabolism and
myofibroblasts induces a migratory, less mature, more prolifer- redox state to exert epigenetic control of cell fate decisions [79].
ative phenotype in ductal cells and re‐enforces the fibrogenic
phenotype of the myofibroblasts [65]. Together with the afore-
mentioned effects on Hh responsive immune cells, these actions Liver diseases with hedgehog dysregulation
promote the ductular reaction, a fibroproliferative inflammatory Hedgehog pathway activity increases with the severity of liver
response to severe acute or chronic liver injury [66]. damage and fibrosis in multiple human liver diseases, including
non‐alcoholic fatty liver disease (NAFLD) [80], alcoholic liver
Liver metabolism disease [81], viral hepatitis [82], schistosomiasis [55], primary
biliary cholangitis (PBC) [83], primary sclerosing cholangitis
The Hh pathway is known to regulate metabolism in adipocytes (PSC) [84], and biliary atresia [85]. Further, liver damage and
[67], fibroblasts [22, 68], stem cells [69], immune cells [70], fibrosis were improved by inhibiting Hh signaling in animal
and many cancer cells [35]. Pathway activity generally stimu- models of some of these conditions (e.g. NAFLD [86, 87],
lates glycolysis [22] and inhibits lipogenesis [71] but the mech- schistosomiasis [88], PSC [56], biliary atresia [89]). To date,
anisms involved are multi‐faceted and not fully understood. For clinical trials of Hh inhibitors have not been done in any of these
example, Smo can directly activate AMP kinase, a master regu- diseases. The major impediment seems to be the risk of adverse
lator of cellular energy balance [21]. Hedgehog signaling may events caused by inhibiting Hh activity in extrahepatic tissues.
also control mitochondrial mass [72] and has been shown to Sustained inhibition of Hh might also have negative impacts on
interact with other pathways that modulate metabolism, such as the liver itself because genetic defects that globally impair Smo
Notch [73] and Wnt [44]. Hence, contextual factors are likely to activation promote the metabolic syndrome [90] and hepatic
influence exactly how Hh impacts metabolism in any given cell. steatosis [91] in humans and inhibiting Smo blocks liver regen-
Until recently, Hh signaling was not thought to be directly eration after partial hepatectomy in mice [92]. The aggregate
involved in hepatocyte metabolism. However, emerging evi- data suggest that future therapeutic approaches to target exces-
dence suggests that Hh may be a critical regulator of hepatic sive Hh signaling must bring pathway activity back to physio-
lipid metabolism [17]. This activity appears to be exquisitely logical levels, rather than silencing signaling completely.
controlled by circadian forces [17] and might have systemic, as
well as local, implications for tissue health. This realization is
triggering research to delineate the mechanisms that control Hh
signaling in hepatocytes. NOTCH SIGNALING
Healthy hepatocytes may produce Hh ligands, although this
has been difficult to demonstrate by immunostaining intact liver
tissue. Failure to visualize Hh ligands in such cells could reflect
Pathway overview
the fact that healthy hepatocytes generally produce low levels of The Notch signaling pathway is fundamentally different from
Hh ligands and/or release them efficiently. The latter possibility the Hh pathway in that it requires cell‐to‐cell contact and the
is supported by two lines of recent evidence. First, Hh ligands interaction of a transmembrane receptor in the Notch target cell
can traffic along cytonemes, providing a mechanism to shuttle with a transmembrane ligand in the ligand‐producing cell [93].
ligands from ligand‐producing hepatocytes to immediately Both pathways ultimately activate transcription factors. The
adjacent ligand‐receiving cells without releasing ligands into canonical Notch pathway has three fundamental components:
the extracellular space [74]. Second, VLDL and other lipopro- (i) the ligands (delta‐like ligands and jagged), (ii) the receptor
tein particles harbor biologically‐active Hh ligands in healthy (Notch), (iii) the effector (transcription factor CBF‐1/Su(H)/
adults [75]. The association of Hh ligands with cholesterol‐ LAG1 [CSL] and coactivator master‐mind‐like [MAML]) [94]
containing lipoproteins is particularly intriguing because cho- (Figure 43.2).
lesterol can bind to the extracellular domain of Smo and activate Both ligands and receptors undergo complex post‐translational
it directly [76], while Hh ligands activate Smo only after modifications before reaching the cell membrane. Ligands are
43:  Developmental Morphogens and Adult Liver Repair 543

Notch ligand
expressing cell

Jag
Dll
Notch

Notch receptor
expressing cell γ-secretase

NICD

NICD

NICD
Nucleus
Hes
RBP-Jk Hey

Figure 43.2  The Notch signaling pathway. Transmembrane ligands (Jagged or Dll) in one cell, bind to transmembrane receptor (Notch) in another
cell, and exposes Notch to cleavage by γ‐secretases, resulting in the release of Notch intracellular domain (NICD). NICD enters the nucleus and
binds to the transcription factor RBP‐Jk, promoting the expression of several target genes such as Hes and Hey.

expressed by the afferent cell as transmembrane proteins from mediated by GSK3β enhances stability of Notch1, whereas it
the Delta/Serrate/LAG‐2 (DSL) family. Mammals have three diminishes activity of Notch2 [100].
Delta‐like ligands (Dll1, Dll3, and Dll4) and two Jagged ligands The NRR region in Notch prevents cleavage by proteases
(Jag1 and Jag2) [94]. Before reaching maturation, ligands [101]. The interaction of ligand with Notch receptor changes the
undergo a cycle of endocytosis and recycling to the cell surface structure of the receptor, exposing cleavage sites to the sequen-
via a process that involves ubiquitylation by the E3 ubiquitin tial action of ADAM metalloproteases and the γ‐secretase com-
ligases, Neuralized and Mindbomb [93]. Interactions between plex [102]. This results in the release of Notch intracellular
ligand and receptor can occur between two cells (in trans) lead- domain (NICD), which translocates into the nucleus due to its
ing to receptor activation, or in the same cell (in cis), resulting nuclear localization sequence. NICD can translocate directly to
in signaling inhibition [95]. There are four receptor paralogs the nucleus or traffic there indirectly via the endosome compart-
(Notch1–4), with different signal strength and thus, ligands ment. The latter process is highly regulated; for example, Numb
may also have different effects depending on which receptor negatively regulates Notch through modulation of NICD endo-
they engage [96]. The extracellular domain of Notch is com- somal trafficking [103]. In contrast, atypical protein kinase C
posed of epidermal growth factor (EGF)‐like repeats and a amplifies Notch signaling by enhancing NICD relocalization
negative regulatory region (NRR). The Notch receptor is modi- from late endosomes to the nucleus [104]. Besides translocating
fied by O‐glycans: O‐fucose adducts generated by Pofut into the nucleus, endosomal Notch can also recycle to the cell
enzyme are required for Notch function; O‐glucose modifica- membrane or undergo lysosomal degradation [105].
tion by Rumi enzymes enhances Notch cleavage; Fringe pro- In the nucleus NICD interacts with the DNA‐binding protein
teins elongate O‐fucose by addition of GlcNac, and O‐xylose CSL (also known as DNA‐binding recombination signal‐bind-
adducts. All of these glycations modulate ligand‐receptor ing protein Jκ [RBP‐Jκ]), and together recruit MAML [106].
interaction [93]. Other post‐translational modifications also
­ Unbound CSL acts as a transcription repressor by binding sev-
modulate cell signaling. For example, methylation by methyl- eral corepressors. In contrast, the complex NICD/CSL/MAML
transferase CARM1 enhances Notch activation [97]; Notch act as a transcription factor, binding to activating cofactors [93].
hydroxylation by protein factor inhibiting HIF‐1 modulates the Classical target genes from the Notch pathway are hairy
hypoxia response [98]. Acetylation of Notch by PCAF and enhancer of split (Hes) and hairy related (Hey) families, but
p300 increases Notch stability, and deacetylation by SIRT1 Notch regulates the expression of components of many impor-
reduces Notch stability [99]; ubiquitylation targets Notch to tant pathways in liver disease, such as hepatocyte nuclear ­factors
rapid proteasome degradation; phosphorylation of Notch1 (HNF), sox9, Hh, Wnt, PDGF, TGF, and VEGF [93].
544 THE LIVER:  NOTCH SIGNALING

Notch signaling can also occur via three noncanonical routes: (e.g. MAML1), and target genes (e.g. Sox9, Spp1, and Hey1) [124].
type 1, ligand independent; type 2, CSL‐independent, and type Although the Notch genetic signature does not correlate with
3, Notch independent. Regarding the type 1 noncanonical path- overall prognosis, it might help select patients who would
way, other ligands have been shown to activate Notch (for respond to Notch‐targeted therapies, since cell lines that share
example MAGP1,2 and DLK1) but the functional consequences Notch activation signature respond to Notch inhibition/inacti-
are not fully understood [107, 108]. On the other hand, types 2 vation with decreased proliferation [124]. A subset of hepato-
and 3 noncanonical signaling seem likely to have functional cellular carcinomas also showed increased components of
effects: CSL‐independent Notch‐dependent signaling upregu- Notch pathway expression at the protein level [126]. Those
lates IL‐6 [109] and some viral proteins directly activate CSL tumors are enriched with CK19+ and Sox9+ cells, markers of
activation independently of Notch [110]. biliary epithelial cells and progenitor cells with bi‐phenotypic
potential and/or stem characteristics. Furthermore, tumor
expression of Notch components correlates with reduced patient
Liver development survival [127, 128] and Notch polymorphisms in tumor cells
Notch is crucial during liver development because it coordinates correlate with increased susceptibility to tumor recurrence post‐
biliary differentiation and morphogenesis of intra‐ and extrahe- surgery [128]. Preclinical studies also suggest that Notch can be
patic bile ducts, as well as the gallbladder [111]. Mutations in a target for cancer therapy, since inhibition of different Notch
Jag1 [112], or less frequently Notch2 [113], cause Alagille syn- receptors or Jag1 in the mice after hepatocellular tumorigenesis
drome, an autosomal disease characterized by intrahepatic bile had occurred, inhibited proliferation, blocked epithelial‐to‐
duct paucity and cholestasis, as well as heart, ocular, and verte- mesenchymal transition, and resulted in reduced tumor burden
bral defects. Mice heterozygous for mutations in Jag1 or Notch2 and metastization [123].
have an Alagille‐like phenotype [114, 115]. Severe cases of The Notch pathway also seems to play a key role in the patho-
extrahepatic biliary atresia have also been associated with mis- genesis of intrahepatic cholangiocarcinoma (ICC) given that the
sense mutations in Jag1 [116]. Notch pathway is a top overexpressed pathway [129], and Jag1
Large extrahepatic bile ducts derive from branching of the overexpression has been demonstrated in virtually all ICC [130].
primitive gut‐derived diverticulum, whereas small intrahepatic ICC can derive from transformation of cholangiocytes [131] or
bile ducts derive from tubular structures within the ductal plate. from the transdifferentiation of fully mature hepatocytes [132],
The ductal plate is a layer of hepatoblasts surrounding the portal both processes being dependent on Notch activation [133].
vein branches in a ring‐like array with ductal commitment, as Interestingly, Jag1‐expressing Kupffer cells stimulate hepatocyte
opposed to hepatoblasts distant to the portal vein, which differ- to cholangiocyte transition and hence, cholangiocarcinogenesis
entiate into hepatocytes [117]. Interestingly, some studies sug- [134]. In human ICC, the overexpression of different Notch
gest that hepatoblasts express the Notch2 receptor and receptors has been shown to correlate with tumor burden, tumor
cholangiocyte differentiation is driven by interaction with Jag1‐ aggressiveness, and reduced overall survival [135]. There is also
expressing periductal and periportal myofibroblasts [118]. increasing evidence linking Notch deregulation with extrahe-
Notch induces transcription factors critical for biliary specifica- patic cholangiocarcinoma and gallbladder cancer [136, 137].
tion (including Hes1, HNF1β and Sox9) and downregulates the
expression of HNF1α and HNF4 [119]. The role of Notch in
biliary differentiation is highly regulated and dependent on an
Liver repair
intricate network of intercommunicating signaling pathways The Notch pathway is crucial to liver regeneration/repair, regu-
such as TGFβ [120], Wnt [121], and Hippo [122]. lating biliary repair, progenitor cell‐mediated liver repair,
­vascular remodeling, and fibrogenesis. All Notch receptors are
expressed in the liver [138]. Notch1 and ‐2 predominate in chol-
Liver neoplasia angiocytes and hepatic progenitor cells (HPC) and increase after
The effect of the Notch pathway in liver carcinogenesis is biliary damage, whereas Notch3 and ‐4 expression increase
diverse and receptor‐dependent. Mouse models of liver cancer after endothelial cell damage [139]. Interestingly, quiescent
suggest that Notch1 and Notch2 promote hepatocellular carci- HSC express Notch1 but activated myofibroblasts downregulate
noma [123, 124]. However, while Notch2 also seems to promote Notch1 and the Notch‐inhibitor Numb, while upregulating
the development of cholangiocarcinoma, Notch1 seems to Notch2 and ‐3 [73, 140]. The main Notch ligands expressed in
inhibit it [123]. The Notch oncogenic effect was linked with the liver are Jag1 on HPC, biliary cells and HSC, and Dll4 on
upregulation of insulin growth factor 2 and Sox9 [124]. endothelial cells [139].
Interestingly, cancer associated fibroblasts may promote car- In the partial hepatectomy animal model of liver regenera-
cinogenesis by expressing Notch ligands that act on neighbor tion, the Notch pathway is upregulated and promotes biliary and
tumor cells [125]. However, the role of Notch in hepatocellular vascular regeneration. After partial hepatectomy, Notch‐RBP/Jk
tumorigenesis is complex. For example, other authors demon- activation on HPC stimulates these cells to become cholangio-
strated a p53‐dependent proapoptotic TRAIL‐sensitizing and cytes and blocks their hepatocyte differentiation [141]. This
chemotherapy‐sensitizing effect of Notch activation in tumor regulation of cell‐fate resembles embryonic morphogenesis,
cells in vitro [126]. being similarly mediated by RBP/Jk downregulation of YAP (a
Regarding human hepatocellular carcinoma, one third of factor that is known to promote hepatocyte regeneration [142]),
patients present a genetic signature of Notch activation, with suppression of HNF1α and HNF4 (factors that promote hepato-
upregulation of Jag1, α‐secretase ADAM17, Notch effectors cytic differentiation), and upregulation of factors that promote
43:  Developmental Morphogens and Adult Liver Repair 545

biliary differentiation (e.g. HNF1β and Sox9) [143]. The Notch SREBP‐1 regulated lipogenic genes (e.g. acetyl‐CoA carboxy-
pathway also promotes hepatocyte regeneration, although not lase‐1 and fatty acids synthase). Additionally, Notch inhibits
directly through activation of Notch signaling in hepatocytes fatty acid oxidation. The aggregate outcomes of increased Notch
[144]. Rather, activation of Notch on endothelial cells induces activity in hepatocytes promote insulin resistance and hepatic
their de‐differentiation, proliferation, vascular remodeling, and steatosis [157]. The Notch pathway can also change metabolism
release of hepatocyte trophic factors such as Wnt2 and hepatic to be cancer cell‐like, modulating mitochondrial function and
growth factor (HGF) [145], and profibrogenic factors leading to deregulating glutamine catabolism such that cell growth becomes
HSC activation [146]. Furthermore, Notch signaling promotes independent of exogenous glutamine [154].
the homing of bone‐marrow derived endothelial progenitor
cells, which are also major sources of trophic factors such as
HGF that are essential for hepatocyte proliferation and restora- Liver diseases with notch dysregulation
tion of liver weight [146, 147]. The role of Notch dysregulation on human disease was first
Regarding wound‐healing responses, HSC respond to biliary described in diseases associated with ductal plate malforma-
damage with upregulation of Jag1, which promotes Notch sign- tions such as Alagille syndrome and biliary atresia. However,
aling on HPC and biliary specification to cholangiocytes [148]. recent evidence has linked Notch to other chronic liver diseases.
On the other hand, after hepatocyte injury, macrophages engulf Adult primary cholangiopathies differ with regards to Notch
hepatocyte debris and upregulate Wnt3, which induces Numb to signaling. For example, PBC has more striking Notch upregula-
block Notch signaling on HPC and promote specification to tion and less Wnt activation than PSC. The ductular reaction
hepatocytes [148]. Evidence also suggests that during biliary also differs in these two diseases: it tends to be exuberant earlier
injury, hepatocytes can transdifferentiate into primitive ductular in the course of PBC and associates with increased expression
cells and regenerate the biliary epithelium via a Notch‐depend- of ductal markers (Sox9 and K‐19) but reduced expression of
ent process [149]. hepatocyte markers [158]. NAFLD animal models show that
The Notch pathway promotes fibrogenesis in liver repair, act- Notch signaling upregulation contributes to obesity‐induced
ing synergically with the Hh pathway [73]. Notch promotes Shh hepatic steatosis [157]. In human steatohepatitis, Notch path-
signaling through the regulation of transport in and out the PC way activation not only correlates with severity of steatosis, but
[150], and Shh promotes Notch signaling through direct upregu- also with severity of liver disease (necroinflammatory activity
lation of Hes1 and Jag2 [151]. Both in vitro and in vivo studies and aminotransferases levels) [159]. Further, Notch‐3 expres-
showed that the Notch pathway drives an epithelial to mesen- sion associates with ductular reaction and Notch‐4 with sinusoi-
chymal‐like transition that transdifferentiates HSC into myofi- dal neovessel proliferation and Kupffer cell activation [160].
broblasts, promoting fibrogenesis [73, 152]. In patients with
chronic liver disease, activation of Notch correlates with the
severity of fibrosis, regardless of etiology of liver disease [153].
Lastly, Notch also enhances inflammatory responses and M1 SUMMARY
polarization on macrophages [153].
Hedgehog and Notch are among the morphogenic signaling
pathways that are typically inactivated once liver development
Liver metabolism
finishes. Both pathways reactivate in adulthood in response to
There is growing evidence that Notch signaling reconfigures cel- liver injury so that healthy liver tissue can be regenerated.
lular metabolic pathways [154], with consequences for the meta- Because these pathways have pleiotropic actions that control cell
bolic syndrome, as well as cancer biology. The effect of Notch fate, they are tightly regulated. Dysfunction of these regulatory
on adipocyte metabolism/function seems to be vary according to mechanisms can lead to either insufficient, or excessive, pathway
the receptor. Activation of Notch pathway generally blocks activity – both result in progressive liver damage. For example,
expansion of the adipose tissue, rendering it less able to cope hypoactivation of Notch signaling can impair biliary regenera-
with energy surplus. Notch also impairs adipocyte function lead- tion and thereby promote progressive ductopenia and/or ductal
ing to insulin resistance, decreased fatty acid oxidation, spill out sclerosis, as occurs in certain primary cholangiopathies.
of fatty acids from adipocytes, and ectopic fat accumulation in Conversely, hyperactivation of the Hh pathway promotes pro-
hepatocytes [155, 156]. Paradoxically, Notch‐1 activation pro- gressive liver fibrosis in many chronic liver diseases, as well as
motes adipogenesis through upregulation of peroxisome prolif- aggressive cancer biology in different types of primary liver can-
erator‐activated receptor (PPAR)‐δ and ‐γ. Interestingly, Notch‐1 cer. Hence, both Hedgehog and Notch are emerging as targets for
also associates with an adipogenic PPARγ dependent quiescent biomarker and therapeutic development in adult liver diseases.
phenotype in HSC [73]. Both obesity/energy surplus, two condi-
tions associated with increased free serum fatty acids, induce
Notch pathway activity in hepatocytes [157]. This results in
increased hepatic glucose production due to NCID enhancement REFERENCES
of forkhead box protein O1 (FOXO1), with consequent tran-
scriptional upregulation of key enzymes of gluconeogenesis (e.g.   1. Roy, S. Cilia and Hedgehog: when and how was their marriage solemnized?
Differentiation, 2012;83(2):S43–8.
glucose 6‐phosphatase and phosphoenolpyruvate carboxykin-   2. Liu, A., Wang, B., and Niswander, L.A. Mouse intraflagellar transport pro-
ase‐1). Simultaneously, Notch activation promotes mTORC1/ teins regulate both the activator and repressor functions of Gli transcription
raptor pathway of lipogenesis, upregulating expression of factors. Development, 2005;132(13):3103–11.
546 THE LIVER:  REFERENCES

3. Chen, Y., Sasai, N., Ma, G. et al. Sonic Hedgehog dependent phospho- 27. Zhao, R. and Duncan, S.A. Embryonic development of the liver. Hepatology,
rylation by CK1alpha and GRK2 is required for ciliary accumulation and 2005;41(5):956–67.
activation of smoothened. PLoS Biol, 2011;9(6):e1001083. 28. Hirose, Y., Itoh, T., and Miyajima, A. Hedgehog signal activation coordinates
4. Porter, J.A., Young, K.E., and Beachy, P.A. Cholesterol modification of hedge- proliferation and differentiation of fetal liver progenitor cells. Exp Cell Res,
hog signaling proteins in animal development. Science, 1996;274(5285):255–9. 2009;315(15):2648–57.
5. Pepinsky, R.B., Zeng, C., Wen, D. et  al. Identification of a palmitic acid‐ 29. Higashiyama, H., Ozawa, A., Sumitomo, H. et al. Embryonic cholecystitis
modified form of human Sonic hedgehog. J Biol Chem, 1998;273(22): and defective gallbladder contraction in the Sox17‐haploinsufficient mouse
14037–45. model of biliary atresia. Development, 2017;144(10):1906–17.
6. Briscoe, J. and Therond, P.P. The mechanisms of Hedgehog signalling 30. Sicklick, J.K., Li, Y.X., Jayaraman, A. et al. Dysregulation of the Hedgehog
and  its roles in development and disease. Nat Rev Mol Cell Biol, pathway in human hepatocarcinogenesis. Carcinogenesis, 2006;27(4):748–57.
2013;14(7):416–29. 31. Della Corte, C.M., Viscardi, G., Papaccio, F. et  al. Implication of the
7. Ayers, K.L., Gallet, A., Staccini‐Lavenant, L., and Therond, P.P. The long‐ Hedgehog pathway in hepatocellular carcinoma. World J Gastroenterol,
range activity of Hedgehog is regulated in the apical extracellular space by 2017;23(24):4330–40.
the glypican Dally and the hydrolase Notum. Dev Cell, 2010;18(4):605–20. 32. Jinawath, A., Akiyama, Y., Sripa, B., and Yuasa, Y. Dual blockade of the
8. Merchant, J.L. and Saqui‐Salces, M. Inhibition of Hedgehog signaling in the Hedgehog and ERK1/2 pathways coordinately decreases proliferation and
gastrointestinal tract: targeting the cancer microenvironment. Cancer Treat survival of cholangiocarcinoma cells. J Cancer Res Clin Oncol,2007;
Rev, 2014;40(1):12–21. 133(4):271–8.
9. Xiao, X., Tang, J.J., Peng, C. et al. Cholesterol Modification of Smoothened 33. Li, Y.C., Deng, Y.H., Guo, Z.H. et al. Prognostic value of hedgehog signal
is required for Hedgehog signaling. Mol Cell,Cell, 2017;66(1):154–62. component expressions in hepatoblastoma patients. Eur J Med Res,
10. Denef, N., Neubuser, D., Perez, L., and Cohen, S.M. Hedgehog induces 2010;15(11):468–74.
opposite changes in turnover and subcellular localization of patched and 34. Tada, M., Kanai, F., Tanaka, Y. et al. Down‐regulation of hedgehog‐interact-
smoothened. Cell, 2000;102(4):521–31. ing protein through genetic and epigenetic alterations in human hepatocel-
11. Izzi, L., Levesque, M., Morin, S. et  al. Boc and Gas1 each form distinct lular carcinoma. Clin Cancer Res, 2008;14(12):3768–76.
Shh receptor complexes with Ptch1 and are required for Shh‐mediated cell 35. Chan, I.S., Guy, C.D., Chen, Y. et  al. Paracrine Hedgehog signaling
proliferation. Dev Cell, 2011;20(6):788–801. drives  metabolic changes in hepatocellular carcinoma. Cancer Res,
12. Chuang, P.T. and McMahon, A.P. Vertebrate Hedgehog signalling modulated 2012;72(24):6344–50.
by induction of a Hedgehog‐binding protein. Nature, 1999;397(6720):617–21. 36. Wan, S., Zhao, E., Kryczek, I. et al. Tumor‐associated macrophages produce
13. Teperino, R., Aberger, F., Esterbauer, H., Riobo, N., and Pospisilik, J.A. interleukin 6 and signal via STAT3 to promote expansion of human hepato-
Canonical and non‐canonical Hedgehog signalling and the control of metab- cellular carcinoma stem cells. Gastroenterology, 2014;147(6):1393–404.
olism. Semin Cell Dev Biol, 2014.;33:81–92. 37. Li, W., Miao, S., Miao, M. et al. Hedgehog signaling activation in hepatic
14. Wang, B. and Li, Y. Evidence for the direct involvement of {beta}TrCP in stellate cells promotes angiogenesis and vascular mimicry in hepatocellular
Gli3 protein processing. Proc Nat Acad Sci USA, 2006;103(1):33–8. carcinoma. Cancer Invest. 2016;34(9):424–30.
15. Ikram, M.S., Neill, G.W., Regl, G., Eichberger, T., Frischauf, A.M., 38. Olive, K.P., Jacobetz, M.A., Davidson, C.J. et  al. Inhibition of Hedgehog
Aberger, F. et al. GLI2 is expressed in normal human epidermis and BCC signaling enhances delivery of chemotherapy in a mouse model of pancreatic
and induces GLI1 expression by binding to its promoter. J Invest Dermatol, cancer. Science, 2009;324(5933):1457–61.
2004;122(6):1503–9. 39. Gan, G.N. and Jimeno, A. Emerging from their burrow: Hedgehog pathway
16. Pan, Y. and Wang, B. A novel protein‐processing domain in Gli2 and Gli3 inhibitors for cancer. Expert Opin Investig Drugs, 2016;25(10):1153–66.
differentially blocks complete protein degradation by the proteasome. J Biol 40. Nault, J.C., Couchy, G., Balabaud, C. et al. Molecular classification of hepa-
Chem, 2007;282(15):10846–52. tocellular adenoma associates with risk factors, bleeding, and malignant
17. Matz‐Soja, M., Rennert, C., Schonefeld, K. et al. Hedgehog signaling is a transformation. Gastroenterology, 2017;152(4):880–94.
potent regulator of liver lipid metabolism and reveals a GLI‐code associated 41. Nault, J.C., Couchy, G., Caruso, S. et al. Argininosuccinate synthase 1 and
with steatosis. Elife, 2016;5. periportal gene expression in sonic hedgehog hepatocellular adenomas.
18. Chinchilla, P., Xiao, L., Kazanietz, M.G., and Riobo, N.A. Hedgehog pro- Hepatology, 2018;68(3):964–76.
teins activate pro‐angiogenic responses in endothelial cells through non‐ 42. Schmidt‐Heck, W., Matz‐Soja, M., Aleithe, S., Marbach, E., Guthke, R., and
canonical signaling pathways. Cell Cycle, 2010;9(3):570–79. Gebhardt, R. Fuzzy modeling reveals a dynamic self‐sustaining network of
19. Barnes, E.A., Kong, M., Ollendorff, V., and Donoghue, D.J. Patched1 the GLI transcription factors controlling important metabolic regulators in
interacts with cyclin B1 to regulate cell cycle progression. EMBO J,
­ adult mouse hepatocytes. Mol Biosyst, 2015;11(8):2190–7.
2001;20(9):2214–23. 43. Gebhardt, R. and Coffer, P.J. Hepatic autophagy is differentially regulated in
20. Yuan, X., Cao, J., He, X. et  al. Ciliary IFT80 balances canonical versus periportal and pericentral zones ‐ a general mechanism relevant for other
non‐canonical hedgehog signalling for osteoblast differentiation. Nat tissues? Cell Commun Signal, 2013;11(1):21.
Commun, 2016;7:11024. 44. van den Brink, G.R., Bleuming, S.A., Hardwick, J.C. et al. Indian Hedgehog
21. Teperino, R., Amann, S., Bayer, M. et  al. Hedgehog partial agonism is an antagonist of Wnt signaling in colonic epithelial cell differentiation. Nat
drives  Warburg‐like metabolism in muscle and brown fat. Cell, Genet, 2004;36(3):277–82.
2012;151(2):414–26. 45. Lin, S., Nascimento, E.M., Gajera, C.R. et  al. Distributed hepatocytes
22. Chen, Y., Choi, S.S., Michelotti, G.A. et al. Hedgehog controls hepatic stel- expressing telomerase repopulate the liver in homeostasis and injury. Nature,
late cell fate by regulating metabolism. Gastroenterology, 2012;143(5):1319– 2018;556(7700):244–8.
29 e1–11. 46. Machado, M.V. and Diehl, A.M. Liver renewal: detecting misrepair and opti-
23. Polizio, A.H., Chinchilla, P., Chen, X., Manning, D.R., and Riobo, N.A. mizing regeneration. Mayo Clin Proc, 2014;89(1):120–30.
Sonic Hedgehog activates the GTPases Rac1 and RhoA in a Gli‐independent 47. Machado, M.V. and Diehl, A.M. Hedgehog signalling in liver pathophysiol-
manner through coupling of smoothened to Gi proteins. Sci Signal, ogy. J Hepatol, 2018;68(3):550–62.
2011;4(200):pt7. 48. Fleig, S.V., Choi, S.S., Yang, L. et al. Hepatic accumulation of Hedgehog‐
24. Bijlsma, M.F., Borensztajn, K.S., Roelink, H., Peppelenbosch, M.P., and reactive progenitors increases with severity of fatty liver damage in mice.
Spek, C.A. Sonic hedgehog induces transcription‐independent cytoskeletal Lab Invest, 2007;87(12):1227–39.
rearrangement and migration regulated by arachidonate metabolites. Cell 49. Rangwala, F., Guy, C.D., Lu, J. et al. Increased production of sonic hedge-
Signal, 2007;19(12):2596–604. hog by ballooned hepatocytes. J Pathol, 2011;224(3):401–10.
25. Dennler, S., Andre, J., Verrecchia, F., and Mauviel, A. Cloning of the human 50. Machado, M.V., Michelotti, G.A., Pereira, T.A. et al. Reduced lipoapoptosis,
GLI2 Promoter: transcriptional activation by transforming growth factor‐beta hedgehog pathway activation and fibrosis in caspase‐2 deficient mice with
via SMAD3/beta‐catenin cooperation. J Biol Chem, 2009;284(46):31523–31. non‐alcoholic steatohepatitis. Gut, 2015;64(7):1148–57.
26. Nolan‐Stevaux, O., Lau, J., Truitt, M.L. et  al. GLI1 is regulated through 51. Witek, R.P., Yang, L., Liu, R. et al. Liver cell‐derived microparticles activate
Smoothened‐independent mechanisms in neoplastic pancreatic ducts and medi- hedgehog signaling and alter gene expression in hepatic endothelial cells.
ates PDAC cell survival and transformation. Genes Dev, 2009;23(1):24–36. Gastroenterology, 2009;136(1):320–30 e2.
43:  Developmental Morphogens and Adult Liver Repair 547

52. Kopinke, D., Roberson, E.C., and Reiter, J.F. Ciliary Hedgehog signaling 77. Murone, M., Rosenthal, A., and de Sauvage, F.J. Sonic hedgehog signaling
restricts injury‐induced adipogenesis. Cell, 2017;170(2):340–51. by the patched‐smoothened receptor complex. Curr Biol, 1999;9(2):76–84.
53. Byrd, N. and Grabel, L. Hedgehog signaling in murine vasculogenesis and 78. Jiang, X.L., Chen, T., and Zhang, X. Activation of sonic hedgehog signaling
angiogenesis. Trends Cardiovasc Med, 2004;14(8):308–13. attenuates oxidized low‐density lipoprotein‐stimulated brain microvascular
54. Syn, W.K., Witek, R.P., Curbishley, S.M. et al. Role for hedgehog pathway endothelial cells dysfunction in vitro. Int J Clin Exp Pathol, 2015;8(10):
in regulating growth and function of invariant NKT cells. Eur J Immunol, 12820–8.
2009;39(7):1879–92. 79. Jeon, S. and Seong, R.H. Anteroposterior limb skeletal patterning requires
55. Pereira, T.A., Xie, G., Choi, S.S. et  al. Macrophage‐derived Hedgehog the bifunctional action of SWI/SNF chromatin remodeling complex in
ligands promotes fibrogenic and angiogenic responses in human schistoso- Hedgehog pathway. PLoS Genet, 2016;12(3):e1005915.
miasis mansoni. Liver Int, 2013;33(1):149–61. 80. Guy, C.D., Suzuki, A., Zdanowicz, M. et al. Hedgehog pathway activation
56. Michelotti, G.A., Xie, G., Swiderska, M. et al. Smoothened is a master regu- parallels histologic severity of injury and fibrosis in human nonalcoholic
lator of adult liver repair. J Clin Invest, 2013;123(6):2380–94. fatty liver disease. Hepatology, 2012;55(6):1711–21.
57. Fingas, C.D., Bronk, S.F., Werneburg, N.W. et  al. Myofibroblast‐derived 81. Jung, Y., Brown, K.D., Witek, R.P. et  al. Accumulation of hedgehog‐
PDGF‐BB promotes Hedgehog survival signaling in cholangiocarcinoma responsive progenitors parallels alcoholic liver disease severity in mice and
cells. Hepatology, 2011;54(6):2076–88. humans. Gastroenterology, 2008;134(5):1532–43.
58. Kumar, V., Mundra, V., and Mahato, R.I. Nanomedicines of Hedgehog inhib- 82. Pereira Tde, A., Witek, R.P., Syn, W.K. et al. Viral factors induce Hedgehog
itor and PPAR‐gamma agonist for treating liver fibrosis. Pharm Res, pathway activation in humans with viral hepatitis, cirrhosis, and hepatocel-
2014;31(5):1158–69. lular carcinoma. Lab Invest, 2010;90(12):1690–703.
59. El‐Agroudy, N.N., El‐Naga, R.N., El‐Razeq, R.A., and El‐Demerdash, E. 83. Jung, Y., McCall, S.J., Li, Y.X., and Diehl, A.M. Bile ductules and stromal
Forskolin, a hedgehog signalling inhibitor, attenuates carbon tetrachloride‐ cells express hedgehog ligands and/or hedgehog target genes in primary
induced liver fibrosis in rats. Br J Pharmacol, 2016;173(22):3248–60. biliary cirrhosis. Hepatology, 2007;45(5):1091–6.
60. Chung, S.I., Moon, H., Ju, H.L. et al. Hepatic expression of Sonic Hedgehog 84. Carpino, G., Cardinale, V., Renzi, A. et al. Activation of biliary tree stem
induces liver fibrosis and promotes hepatocarcinogenesis in a transgenic cells within peribiliary glands in primary sclerosing cholangitis. J Hepatol,
mouse model. J Hepatol, 2016;64(3):618–27. 2015;63(5):1220–8.
61. Xie, G., Choi, S.S., Syn, W.K. et  al. Hedgehog signalling regulates liver 85. Jung, H.Y., Jing, J., Lee, K.B., and Jang, J.J. Sonic hedgehog (SHH) and
sinusoidal endothelial cell capillarisation. Gut, 2013;62(2):299–309. glioblastoma‐2 (Gli‐2) expressions are associated with poor jaundice‐free
62. Uschner, F.E., Ranabhat, G., Choi, S.S. et al. Statins activate the canonical survival in biliary atresia. J Pediatr Surg, 2015;50(3):371–6.
hedgehog‐signaling and aggravate non‐cirrhotic portal hypertension, but 86. Syn, W.K., Jung, Y., Omenetti, A. et al. Hedgehog‐mediated epithelial‐to‐
inhibit the non‐canonical hedgehog signaling and cirrhotic portal hyperten- mesenchymal transition and fibrogenic repair in nonalcoholic fatty liver
sion. Sci Rep, 2015;5:14573. disease. Gastroenterology, 2009;137(4):1478–88 e8.
63. Omenetti, A., Syn, W.K., Jung, Y. et al. Repair‐related activation of hedge- 87. Hirsova, P., Ibrahim, S.H., Bronk, S.F., Yagita, H., and Gores, G.J.
hog signaling promotes cholangiocyte chemokine production. Hepatology, Vismodegib suppresses TRAIL‐mediated liver injury in a mouse model of
2009;50(2):518–27. nonalcoholic steatohepatitis. PloS One, 2013;8(7):e70599.
64. Xie, J. The hedgehog’s trick for escaping immunosurveillance: The molecu- 88. de Almeida Pereira, T., Borthwick, L., Xie, G. et  al. Crosstalk between
lar mechanisms driving myeloid‐derived suppressor cell recruitment in IL13 and Hedgehog pathways contributes to Schistosomiasis mansoni
hedgehog signaling‐dependent tumors. Oncoimmunology, 2014;3:e29180. fibrosis. J Hepatol, 2016;62(2):S204–S.
65. Omenetti, A., Porrello, A., Jung, Y. et al. Hedgehog signaling regulates epi- 89. Tang, V., Cofer, Z.C., Cui, S., Sapp, V., Loomes, K.M., and Matthews, R.P.
thelial‐mesenchymal transition during biliary fibrosis in rodents and humans. Loss of a candidate biliary atresia susceptibility gene, add3a, causes biliary
J Clin Invest, 2008;118(10):3331–42. developmental defects in zebrafish. J Pediatr Gastroenterol Nutr, 2016;63(5):
66. Omenetti, A., Popov, Y., Jung, Y. et  al. The hedgehog pathway regulates 524–30.
remodelling responses to biliary obstruction in rats. Gut, 2008;57(9): 90. Ali, O., Cerjak, D., Kent, J.W., Jr. et al. An epigenetic map of age‐associated
1275–82. autosomal loci in northern European families at high risk for the metabolic
67. Fleury, A., Hoch, L., Martinez, M.C. et al. Hedgehog associated to micropar- syndrome. Clin Epigenetics, 2015;7:12.
ticles inhibits adipocyte differentiation via a non‐canonical pathway. Sci 91. Sacoto, M.J.G., Martinez, A.F., Abe, Y. et al. Human germline Hedgehog
Rep, 2016;6:23479. pathway mutations predispose to fatty liver. J Hepatol, 2017;67(4):809–17.
68. Du, K., Hyun, J., Premont, R.T. et  al. Hedgehog‐YAP signaling pathway 92. Swiderska‐Syn, M., Syn, W.K., Xie, G. et al. Myofibroblastic cells function
regulates glutaminolysis to control activation of hepatic stellate cells. as progenitors to regenerate murine livers after partial hepatectomy. Gut,
Gastroenterology, 2018;154(5):1465–79. 2014;63(8):1333–44.
69. Fu, X., Zhu, M.J., Dodson, M.V., and Du, M. AMP‐activated protein kinase 93. Siebel, C. and Lendahl, U. Notch Signaling in development, tissue homeo-
stimulates Warburg‐like glycolysis and activation of satellite cells during stasis, and disease. Physiol Rev, 2017;97(4):1235–94.
muscle regeneration. J Biol Chem, 2015;290(44):26445–56. 94. Meurette, O. and Mehlen, P. Notch signaling in the tumor microenviron-
70. Song, M., Han, L., Chen, F.F. et al. Adipocyte‐derived exosomes carrying ment. Cancer Cell, 2018;34(4):536–48.
Sonic Hedgehog mediate M1 macrophage polarization‐induced insulin 95. Jacobsen, T.L., Brennan, K., Arias, A.M., and Muskavitch, M.A. Cis‐inter-
resistance via Ptch and PI3K pathways. Cell Physiol Biochem, 2018;48(4): actions between Delta and Notch modulate neurogenic signalling in
1416–32. Drosophila. Development, 1998;125(22):4531–40.
71. Bhatia, B., Hsieh, M., Kenney, A.M., and Nahle, Z. Mitogenic Sonic hedge- 96. Van de Walle, I., Waegemans, E., De Medts, J. et  al. Specific Notch
hog signaling drives E2F1‐dependent lipogenesis in progenitor cells and receptor‐ligand interactions control human TCR‐alphabeta/gammadelta
­
medulloblastoma. Oncogene, 2011;30(4):410–22. development by inducing differential Notch signal strength. J Exp Med,
72. Yao, P.J., Manor, U., Petralia, R.S. et al. Sonic hedgehog pathway activation 2013;210(4):683–97.
increases mitochondrial abundance and activity in hippocampal neurons. 97. Hein, K., Mittler, G., Cizelsky, W. et al. Site‐specific methylation of Notch1
Mol Biol Cell, 2017;28(3):387–95. controls the amplitude and duration of the Notch1 response. Sci Signal,
73. Xie, G., Karaca, G., Swiderska‐Syn, M. et  al. Cross‐talk between Notch 2015;8(369):ra30.
and  Hedgehog regulates hepatic stellate cell fate in mice. Hepatology, 98. Coleman, M.L., McDonough, M.A., Hewitson, K.S. et  al. Asparaginyl
2013;58(5):1801–13. hydroxylation of the Notch ankyrin repeat domain by factor inhibiting
74. Gonzalez‐Mendez, L., Seijo‐Barandiaran, I., and Guerrero, I. Cytoneme‐ hypoxia‐inducible factor. J Biol Chem, 2007;282(33):24027–38.
mediated cell‐cell contacts for Hedgehog reception. Elife, 2017;6. 99. Guarani, V., Deflorian, G., Franco, C.A. et  al. Acetylation‐dependent
75. Panakova, D., Sprong, H., Marois, E., Thiele, C., and Eaton, S. Lipoprotein regulation of endothelial Notch signalling by the SIRT1 deacetylase.
­
particles are required for Hedgehog and Wingless signalling. Nature, Nature, 2011;473(7346):234–8.
2005;435(7038):58–65. 100. Foltz, D.R., Santiago, M.C., Berechid, B.E., and Nye, J.S. Glycogen syn-
76. Huang, P., Nedelcu, D., Watanabe, M. et al. Cellular cholesterol directly acti- thase kinase‐3beta modulates notch signaling and stability. Curr Biol,
vates Smoothened in Hedgehog signaling. Cell, 2016;166(5):1176–87 e14. 2002;12(12):1006–11.
548 THE LIVER:  REFERENCES

101. Kovall, R.A., Gebelein, B., Sprinzak, D., and Kopan, R. The canonical 124. Villanueva, A., Alsinet, C., Yanger, K. et al. Notch signaling is activated in
Notch signaling pathway: structural and biochemical insights into shape, human hepatocellular carcinoma and induces tumor formation in mice.
sugar, and force. Dev Cell, 2017;41(3):228–41. Gastroenterology, 2012;143(6):1660–9.
102. Gordon, W.R., Zimmerman, B., He, L. et al. Mechanical allostery: evidence 125. Xiong, S., Wang, R., Chen, Q. et al. Cancer‐associated fibroblasts promote
for a force requirement in the proteolytic activation of Notch. Dev Cell, stem cell‐like properties of hepatocellular carcinoma cells through IL‐6/
2015;33(6):729–36. STAT3/Notch signaling. Am J Cancer Res, 2018;8(2):302–16.
103. Couturier, L., Mazouni, K., and Schweisguth, F. Numb localizes at 126. Giovannini, C., Gramantieri, L., Chieco, P. et  al. Selective ablation of
endosomes and controls the endosomal sorting of notch after asymmetric Notch3 in HCC enhances doxorubicin’s death promoting effect by a p53
division in Drosophila. Curr Biol, 2013;23(7):588–93. dependent mechanism. J Hepatol, 2009;50(5):969–79.
104. Sjoqvist, M., Antfolk, D., Ferraris, S. et al. PKCzeta regulates Notch recep- 127. Wu, T., Jiao, M., Jing, L. et al. Prognostic value of Notch‐1 expression in hepa-
tor routing and activity in a Notch signaling‐dependent manner. Cell Res, tocellular carcinoma: a meta‐analysis. Onco Targets Ther, 2015;8:3105–14.
2014;24(4):433–50. 128. Yu, T., Han, C., Zhu, G. et al. Prognostic value of Notch receptors in post-
105. Leitch, C.C., Lodh, S., Prieto‐Echague, V., Badano, J.L., and Zaghloul, surgical patients with hepatitis B virus‐related hepatocellular carcinoma.
N.A. Basal body proteins regulate Notch signaling through endosomal traf- Cancer Med, 2017;6(7):1587–600.
ficking. J Cell Sci, 2014;127(11):2407–19. 129. Xue, T.C., Zhang, B.H., Ye, S.L., and Ren, Z.G. Differentially expressed
106. Nam, Y., Weng, A.P., Aster, J.C., and Blacklow, S.C. Structural require- gene profiles of intrahepatic cholangiocarcinoma, hepatocellular carci-
ments for assembly of the CS.L.intracellular Notch1.Mastermind‐like 1 noma, and combined hepatocellular‐cholangiocarcinoma by integrated
transcriptional activation complex. J Biol Chem, 2003;278(23):21232–9. microarray analysis. Tumour Biol, 2015;36(8):5891–9.
107. Miyamoto, A., Lau, R., Hein, P.W., Shipley, J.M., and Weinmaster, G. 130. Che, L., Fan, B., Pilo, M.G. et al. Jagged 1 is a major Notch ligand along
Microfibrillar proteins MAGP‐1 and MAGP‐2 induce Notch1 extracellular cholangiocarcinoma development in mice and humans. Oncogenesis,
domain dissociation and receptor activation. J Biol Chem, 2006;281(15): 2016;5(12):e274.
10089–97. 131. Guest, R.V., Boulter, L., Kendall, T.J. et al. Cell lineage tracing reveals a
108. Traustadottir, G.A., Jensen, C.H., Thomassen, M. et al. Evidence of non‐ biliary origin of intrahepatic cholangiocarcinoma. Cancer Res, 2014;
canonical NOTCH signaling: delta‐like 1 homolog (DLK1) directly inter- 74(4):1005–10.
acts with the NOTCH1 receptor in mammals. Cell Signal, 2016;28(4): 132. Fan, B., Malato, Y., Calvisi, D.F. et al. Cholangiocarcinomas can originate
246–54. from hepatocytes in mice. J Clin Invest, 2012;122(8):2911–5.
109. Jin, S., Mutvei, A.P., Chivukula, I.V. et al. Non‐canonical Notch signaling 133. Wang, J., Dong, M., Xu, Z. et al. Notch2 controls hepatocyte‐derived chol-
activates IL‐6/JAK/STAT signaling in breast tumor cells and is controlled angiocarcinoma formation in mice. Oncogene, 2018;37(24):3229–42.
by p53 and IKKalpha/IKKbeta. Oncogene. 2013;32(41):4892–902. 134. Terada, M., Horisawa, K., Miura, S. et al. Kupffer cells induce Notch‐medi-
110. Zimber‐Strobl, U. and Strobl, L.J. EBNA2 and Notch signalling in Epstein‐ ated hepatocyte conversion in a common mouse model of intrahepatic chol-
Barr virus mediated immortalization of B lymphocytes. Semin Cancer Biol, angiocarcinoma. Sci Rep, 2016;6:34691.
2001;11(6):423–34. 135. Wu, W.R., Shi, X.D., Zhang, R. et al. Clinicopathological significance of
111. Zong, Y., Panikkar, A., Xu, J. et al. Notch signaling controls liver develop- aberrant Notch receptors in intrahepatic cholangiocarcinoma. Int J Clin
ment by regulating biliary differentiation. Development, 2009;136(10): Exp Pathol, 2014;7(6):3272–9.
1727–39. 136. Aoki, S., Mizuma, M., Takahashi, Y. et  al. Aberrant activation of Notch
112. Oda, T., Elkahloun, A.G., Pike, B.L. et al. Mutations in the human Jagged1 signaling in extrahepatic cholangiocarcinoma: clinicopathological features
gene are responsible for Alagille syndrome. Nat Genet, 1997;16(3):235–42. and therapeutic potential for cancer stem cell‐like properties. BMC Cancer,
113. McDaniell, R., Warthen, D.M., Sanchez‐Lara, P.A. et al. NOTCH2 muta- 2016;16(1):854.
tions cause Alagille syndrome, a heterogeneous disorder of the notch sign- 137. Yoon, H.A., Noh, M.H., Kim, B.G. et al. Clinicopathological significance
aling pathway. Am J Hum Genet, 2006;79(1):169–73. of altered Notch signaling in extrahepatic cholangiocarcinoma and gall-
114. McCright, B., Lozier, J., and Gridley, T. A mouse model of Alagille syn- bladder carcinoma. World J Gastroenterol, 2011;17(35):4023–30.
drome: Notch2 as a genetic modifier of Jag1 haploinsufficiency. 138. Nijjar, S.S., Crosby, H.A., Wallace, L., Hubscher, S.G., and Strain, A.J. Notch
Development, 2002;129(4):1075–82. receptor expression in adult human liver: a possible role in bile duct forma-
115. Andersson, E.R., Chivukula, I.V., Hankeova, S. et  al. Mouse model of tion and hepatic neovascularization. Hepatology, 2001;34(6):1184–92.
Alagille syndrome and mechanisms of Jagged1 missense mutations. 139. Morell, C.M., Fiorotto, R., Fabris, L., and Strazzabosco, M. Notch signal-
Gastroenterology, 2018;154(4):1080–95. ling beyond liver development: emerging concepts in liver repair and onco-
116. Kohsaka, T., Yuan, Z.R., Guo, S.X. et al. The significance of human jagged genesis. Clin Res Hepatol Gastroenterol, 2013;37(5):447–54.
1 mutations detected in severe cases of extrahepatic biliary atresia. 140. Sawitza, I., Kordes, C., Reister, S., and Haussinger, D. The niche of stellate
Hepatology, 2002;36(4 Pt 1):904–12. cells within rat liver. Hepatology, 2009;50(5):1617–24.
117. Lemaigre, F. and Zaret, K.S. Liver development update: new embryo mod- 141. Lu, J., Zhou, Y., Hu, T. et al. Notch signaling coordinates progenitor cell‐
els, cell lineage control, and morphogenesis. Curr Opin Genet Dev, mediated biliary regeneration following partial hepatectomy. Sci Rep,
2004;14(5):582–90. 2016;6:22754.
118. Furubo, S., Sato, Y., Harada, K., and Nakanuma, Y. Roles of myofibroblasts 142. Yimlamai, D., Christodoulou, C., Galli, G.G. et al. Hippo pathway activity
and notch and hedgehog signaling pathways in the formation of intrahe- influences liver cell fate. Cell, 2014;157(6):1324–38.
patic bile duct lesions in polycystic kidney rats. Pediatr Dev Pathol, 143. Fabris, L., Cadamuro, M., Guido, M. et al. Analysis of liver repair mecha-
2013;16(3):177–90. nisms in Alagille syndrome and biliary atresia reveals a role for notch sign-
119. Strazzabosco, M. and Fabris, L. Notch signaling in hepatocellular carci- aling. Am J Pathol, 2007;171(2):641–53.
noma: guilty in association! Gastroenterology, 2012;143(6):1430–4. 144. Dill, M.T., Rothweiler, S., Djonov, V. et al. Disruption of Notch1 induces
120. Wang, W., Feng, Y., Aimaiti, Y., Jin, X., Mao, X., and Li, D. TGFbeta vascular remodeling, intussusceptive angiogenesis, and angiosarcomas in
signaling controls intrahepatic bile duct development may through
­ livers of mice. Gastroenterology, 2012;142(4):967–77 e2.
regulating the Jagged1‐Notch‐Sox9 signaling axis. J Cell Physiol,
­ 145. Ding, B.S., Nolan, D.J., Butler, J.M. et  al. Inductive angiocrine signals
2018;233(8):5780–91. from sinusoidal endothelium are required for liver regeneration. Nature,
121. So, J., Khaliq, M., Evason, K. et  al. Wnt/beta‐catenin signaling controls 2010;468(7321):310–5.
intrahepatic biliary network formation in zebrafish by regulating notch 146. DeLeve, L.D. Liver sinusoidal endothelial cells and liver regeneration. J
activity. Hepatology, 2018;67(6):2352–66. Clin Invest, 2013;123(5):1861–6.
122. Wu, N., Nguyen, Q., Wan, Y. et al. The Hippo signaling functions through 147. Wang, L., Wang, Y.C., Hu, X.B. et al. Notch‐RBP‐J signaling regulates the
the Notch signaling to regulate intrahepatic bile duct development in mam- mobilization and function of endothelial progenitor cells by dynamic mod-
mals. Lab Invest, 2017;97(7):843–53. ulation of CXCR4 expression in mice. PloS One, 2009;4(10):e7572.
123. Huntzicker, E.G., Hotzel, K., Choy, L. et al. Differential effects of targeting 148. Boulter, L., Govaere, O., Bird, T.G. et al. Macrophage‐derived Wnt opposes
Notch receptors in a mouse model of liver cancer. Hepatology, Notch signaling to specify hepatic progenitor cell fate in chronic liver dis-
2015;61(3):942–52. ease. Nat Med, 2012;18(4):572–9.
43:  Developmental Morphogens and Adult Liver Repair 549

149. Jeliazkova, P., Jors, S., Lee, M. et  al. Canonical Notch2 signaling deter- 155. Chartoumpekis, D.V., Palliyaguru, D.L., Wakabayashi, N. et  al. Notch
mines biliary cell fates of embryonic hepatoblasts and adult hepatocytes intracellular domain overexpression in adipocytes confers lipodystrophy in
independent of Hes1. Hepatology, 2013;57(6):2469–79. mice. Mol Metab, 2015;4(7):543–50.
150. Ingram, W.J., McCue, K.I., Tran, T.H., Hallahan, A.R., and Wainwright, 156. Song, N.J., Yun, U.J., Yang, S. et al. Notch1 deficiency decreases hepatic
B.J. Sonic Hedgehog regulates Hes1 through a novel mechanism that is lipid accumulation by induction of fatty acid oxidation. Sci Rep,
independent of canonical Notch pathway signalling. Oncogene, 2016;6:19377.
2008;27(10):1489–500. 157. Pajvani, U.B., Qiang, L., Kangsamaksin, T., Kitajewski, J., Ginsberg, H.N.,
151. Stasiulewicz, M., Gray, S.D., Mastromina, I. et  al. A conserved role and Accili, D. Inhibition of Notch uncouples Akt activation from
for  Notch signaling in priming the cellular response to Shh through hepatic  lipid accumulation by decreasing mTorc1 stability. Nat Med,
ciliary  localisation of the key Shh transducer Smo. Development, 2013;19(8):1054–60.
2015;142(13):2291–303. 158. Carpino, G., Cardinale, V., Folseraas, T. et al. Hepatic stem/progenitor cell
152. Chen, Y., Zheng, S., Qi, D. et al. Inhibition of Notch signaling by a gamma‐ activation differs between primary sclerosing and primary biliary cholangi-
secretase inhibitor attenuates hepatic fibrosis in rats. PloS One, tis. Am J Pathol, 2018;188(3):627–39.
2012;7(10):e46512. 159. Valenti, L., Mendoza, R.M., Rametta, R. et al. Hepatic notch signaling cor-
153. He, F., Guo, F.C., Li, Z. et al. Myeloid‐specific disruption of recombina- relates with insulin resistance and nonalcoholic fatty liver disease. Diabetes,
tion signal binding protein Jkappa ameliorates hepatic fibrosis by attenu- 2013;62(12):4052–62.
ating inflammation through cylindromatosis in mice. Hepatology, 160. Liew, P.L., Wang, W., Lee, Y.C., Huang, M.T., and Lee, W.J. Roles
2015;61(1):303–14. of  hepatic progenitor cells activation, ductular reaction proliferation
154. Basak, N.P., Roy, A., and Banerjee, S. Alteration of mitochondrial pro- and  Notch signaling in morbid obesity. Hepatogastroenterology, 2012;
teome due to activation of Notch1 signaling pathway. J Biol Chem, 59(118):1921–7.
2014;289(11):7320–34.
Liver Repopulation by Cell
44 Transplantation and the
Role of Stem Cells in Liver
Biology
David A. Shafritz1 and Markus Grompe2
1
Marion Bessin Liver Research Center, Albert Einstein College of Medicine, Bronx, NY, USA
2
Papé Pediatric Research Institute, Oregon Health Sciences University, Portland, OR, USA

INTRODUCTION HEPATOCYTE TRANSPLANTATION:


RATIONALE AND EARLY STUDIES
The major impetus for trying to reconstitute the liver by cell
transplantation is based on the very high regenerative capacity Currently, orthotopic liver transplantation (OLT) is the only
of this organ. The very first experiments were done in the early effective method available to cure acquired and inherited hepatic
1900s, when liver fragments were transplanted into the anterior disorders, especially when these diseases reach their end stages
chamber of the eye, but the transplanted liver tissue degenerated [3, 4]. However, the number of patients who are fortunate
rapidly and disappeared within a few days [1]. The first success- enough to receive a liver transplant is limited by donor organ
ful report of liver cell transplantation was in the 1970s, when availability. Liver transplantation is also expensive, carries sig-
isolated hepatocytes were transplanted to the liver, leading to nificant morbidity and mortality, and requires long‐term
transient reduction in serum bilirubin in the Gunn rat model for immunosuppression.
Crigler–Najjar syndrome, type 1 [2]. More recently, substantial Hepatocyte transplantation has been shown to be therapeutic
progress has been made defining the cell types that can repopu- in animal models of acute liver failure [5] and limited human
late the liver and restore function, and the host liver conditions clinical studies have been promising [6]. Inherited liver disorders
under which effective repopulation can be achieved. Both adult are also candidates for cell therapy because they are caused by
hepatocytes and hepatic stem‐like progenitor cells (stem/pro- loss of expression or dysfunction of a single hepatocyte‐specific
genitor cells), as well as stem and progenitor cells of non‐hepatic gene (i.e. monogenic). Replacement of these “diseased” hepato-
origin, have been explored for transplantation into the liver. cytes by normal hepatocytes should be therapeutic. Specific dis-
Studies concerning hepatocyte regeneration during both homeo- ease examples include Crigler–Najjar syndrome, type 1,
stasis and injury, including cell plasticity and interconversion ornithine transcarbamylase deficiency and other urea cycle dis-
between hepatocytes and bile duct epithelial cells and the cur- orders, familial hypercholesterolemia, factors VII and IX defi-
rent status of hepatic cell transplantation in humans will be cov- ciency, and phenylketonuria (diseases in which there is no
ered, as well as expectations for the future. underlying liver injury or damage) and other diseases such as

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
44:  Liver Repopulation by Cell Transplantation and the Role of Stem Cells in Liver Biology 551

Wilson’s disease, α1‐antitrypsin deficiency, and hemochromato- subsequently received a liver transplant, so that the long‐term
sis (diseases in which there is extensive liver injury). In inherited effects of hepatocyte transplantation could not be assessed.
monogenic liver disorders, autologous cell transplantation may However, in the remaining patients, the function of transplanted
be possible, in which the patient’s cells are genetically manipu- hepatocytes was not sustained. Whether this resulted from rejec-
lated ex vivo and then transplanted back into his/her liver without tion of transplanted cells, their lack of proliferation, or eventual
the need for immunosuppression (and indeed this has been done) apoptosis has not been determined. In no patient was expansion
[7]. However, only a small percentage of the total hepatocellular of transplanted hepatocytes or progressive improvement of met-
mass (around 1–2% maximum) can be replaced by simple hepat- abolic function over time documented.
ocyte transplantation without causing portal hypertension and/or Hepatocyte transplantation therapy for fulminant hepatic fail-
hepatic infarction. Therefore, in most instances, it will be neces- ure as a bridge to orthotopic liver transplantation has had some
sary to selectively expand the therapeutic cells in vivo after their encouraging results [5, 6]. The numbers of patients treated by
engraftment. The therapeutic threshold will be different for each this modality are low, the diseases are of varying etiologies and
disease, but generally cell replacement levels of at least 5–10% therefore it is not possible to draw definitive conclusions regard-
will be needed for most disorders. ing the efficacy of this intervention. Data from several pilot
Attempts to increase the proportion of transplanted hepato- studies have been reported, many with reductions in both blood
cytes in the liver simply by stimulating liver regeneration (e.g. ammonia and encephalopathy and successful organ transplant
through partial hepatectomy (PH) or carbon tetrachloride (CCl4)‐ [5]. In several cases, the patients recovered without the need for
induced hepatic necrosis) have generally failed to show signifi- a transplant, but others died.
cant benefit. This is not surprising, considering that hepatocytes At this stage it can be said that allogeneic hepatocyte trans-
need to undergo only one or two rounds of cell division to replace plantation in humans is safe and that the available data suggest
the mass of liver removed by two‐thirds PH [8, 9]. Performing partial efficacy for both genetic and acquired hepatic disorders.
repeated PH or CCl4 administration does not significantly
increase repopulation by donor hepatocytes [10], since both
transplanted and host hepatocytes respond similarly to this
regenerative stimulus. Repeated cell transplantations have also BASIC REQUIREMENTS FOR LIVER
been performed, [11, 12] but this also did not significantly REGENERATION
increase the efficacy of liver replacement by transplanted cells.
In the normal adult liver, hepatocytes are quiescent, turning over
very slowly (only two to three times per year). However, follow-
ing surgical reduction of liver mass or extensive acute toxic liver
CLINICAL TRIALS OF HEPATOCYTE injury, hepatocytes rapidly enter the cell cycle and proliferate to
TRANSPLANTATION restore liver mass. During liver regeneration, 70–90% of the
residual mature hepatocytes engage in DNA synthesis and
To date, allogeneic hepatocyte transplantation in humans has undergo cell division [9]. Liver regeneration is a highly organ-
had limited clinical success. In 1998, a child with Crigler–Najjar ized, complex process involving growth factors and cytokines,
syndrome type I, suffering from severe hyperbilirubinemia, was transcription factors, cell signaling pathways, and expression of
infused with 7.5 × 109 allogeneic donor hepatocytes via a portal cell cycle regulatory genes [17]; (see Chapter 45). From many
vein catheter [13]. This resulted in a significant reduction of studies, it has been concluded that the proliferative activity of
serum bilirubin. However, after two and a half years serum bili- adult hepatocytes is sufficient to regenerate the liver following
rubin progressively increased to pretreatment levels, although two‐thirds PH without participation of stem cells [18].
bilirubin glucuronides were still detectable in the bile (J. Roy
Chowdhury, personal communication). Studies have also been
conducted in patients with chronic liver disease and cirrhosis,
again with only marginal, if any, success [14]. LIVER SIZE CONTROL AND “HIPPO
Hepatocyte transplantation has also been used in conjunction SIGNALING PATHWAY”
with ex vivo retroviral gene therapy in five patients with a defect
in the low‐density lipoprotein (LDL) receptor [7]. In this study, For many years, it has been known that the liver size (mass) is
the proportion of liver cell mass replaced was estimated to be proportional to total body weight, ranging from 3 to 5% in differ-
around 1%. The procedure resulted in a very modest decrease in ent mammalian species (see Chapter  45). After two‐thirds PH,
plasma cholesterol in several of the patients. cellular proliferation begins within 12–18 hours and the liver size
The worldwide experience with human hepatocyte transplan- returns to normal within 1–2 weeks in rodents and after 4–12
tation has been updated recently [15, 16]. Usually, suspensions months in humans. When an undersized liver is transplanted, it
of adult hepatocytes (autographs or allographs) were utilized in grows to the expected full size for the host, and when an oversized
a single injection, although some patients received multiple liver is transplanted, its size is reduced to the expected mass com-
infusions. Many patients had inherited metabolic disorders, pared with total body weight (see Chapter  45). However, until
including familial hypercholesterolemia, Crigler–Najjar very recently, little was known concerning how this process is
­syndrome type 1, factor VII deficiency, urea cycle defects, gly- controlled. In 2007, Pan and colleagues [19] showed that mam-
cogen storage diseases, and others [14, 16]. Initial improvement malian genes comparable to those in the Drosophila Hippo kinase
of metabolic function was frequently observed. Some patients signaling cascade, that regulates wing mass during development,
552 THE LIVER:  ANIMAL MODELS FOR THERAPEUTIC LIVER REPOPULATION

control hepatocyte proliferation. When YAP, the mammalian after  which they observed extensive liver repopulation
counterpart to Yorki, the last gene in the Drosophila Hippo kinase (Figure  44.1b). They estimated that each donor hepatocyte
cascade, is overexpressed in a transgenic mouse model, hepato- engrafted into the albumin uPA host liver, underwent 12–14
cyte proliferation becomes unchecked and there is massive liver cell divisions [23].
hyperplasia, leading to hepatic carcinogenesis. When YAP hyper- Another mouse model for liver repopulation was generated
expression is turned off or blocked within four to six weeks after by targeted disruption of the last gene in tyrosine catabolism,
birth, liver size returns to normal [19]. fumarylacetoacetate hydrolase (Fah) [24]. Deletion of Fah leads
YAP is synthesized in the cytoplasm and translocated to the to accumulation of upstream intermediates in tyrosine catabo-
nucleus, where it functions as a transcriptional coactivator of lism, some of which (namely, fumarylacetoacetate) are toxic
TEADs, p73, RUNX2, and other genes. Control of YAP func- and cause extensive and continuous liver injury. The Fah null
tion is achieved through phosphorylation by upstream kinases mouse represents an animal model for the human metabolic dis-
in the Hippo signaling pathway, Mst1/2 that phosphorylates order hereditary tyrosinemia, type 1 (HT1), which causes exten-
Lats1/2, and pLats1/2 that phosphorylates YAP at amino acid sive liver injury, hepatocellular carcinoma, and death at an early
S127. pYAP phosphorylated at S127 is retained in the cyto- age in affected individuals. Administration of 2‐(2‐nitro‐4‐trif-
plasm, where it is subsequently degraded [19, 21]. YAP is a luoromethylbenzoyl) cyclohexane‐1,3‐dione (NTBC), a phar-
proliferative gene with around 500 known downstream targets, macological inhibitor of tyrosine catabolism upstream of
including many that effect cell growth and the cell cycle. YAP homogentisic acid, prevents accumulation of fumarylacetoace-
also induces two anti‐apoptotic genes, Birc2 and Birc5, so that tate and is successful in treating patients with HT1 [25]. NTBC
it also increases survival of cells in which YAP nuclear func- also allows Fah null mice to survive and liver failure occurs in
tion is enhanced (i.e. by reduced phosphorylation of YAP at these mice only when NTBC is discontinued.
S127). Most studies reporting on the tumorigenic properties of After transplanting syngeneic wild‐type (wt) hepatocytes
YAP in hepatocytes have been performed in transgenic mice or into Fah null mice maintained on NTBC, only scattered small
in hepatocytes expressing a non‐phosphorylatable mutant of clusters of transplanted hepatocytes are detected (Figure 44.1c
YAP (YAP S127A), in which YAP expression/function is con- left panel). However, if NTBC treatment is discontinued
stitutive and under control of a strong promoter (i.e. the tet shortly after cell transplantation, liver injury resumes and
operon) [20].Therefore, it unclear whether the YAP is inher- transplanted cells proliferate extensively, forming large clus-
ently oncogenic or whether this reported phenotype is the ters within three weeks (middle panel) and replacing most of
result of aberrant or uncontrolled YAP hyperexpression/ the liver mass within six weeks (right panel). Fah null mice
function. with repopulated livers remain healthy, have normal liver
function tests and show a relatively normal liver structure for
many months after wt hepatocyte transplantation [24]. These
studies provide proof of principle that liver repopulation can
ANIMAL MODELS FOR THERAPEUTIC effectively cure a monogenic liver disease, namely, the mouse
LIVER REPOPULATION equivalent to HT1.
Transplanted repopulating hepatocytes integrate into the
For many years, it was thought that mature hepatocytes could host hepatic architecture and express all functions necessary
undergo only two or three divisions, after which they become for normal health of the animals. In the Fah null mouse, not
terminally differentiated and incapable of further prolifera- only do transplanted wt hepatocytes replace Fah null hepato-
tion. More recently, however, it has been shown in several cytes, but the transplanted cells can also be serially transplanted
rodent model systems that extensive modifications of the liver through up to twelve consecutive Fah null mice, while retain-
microenvironment permit hepatocytes to retain high prolifera- ing full ability to proliferate and replace host hepatocytes [26,
tive capability and to effectively repopulate the liver. Under 27]. In these studies, it was calculated that each serially trans-
these conditions, the level of cell replacement can be 90% and planted hepatocyte underwent an average of at least 69 cell
higher, providing proof of principle for therapeutic liver divisions. Thus, murine hepatocytes exhibit essentially infinite
repopulation by transplanted cells. Originally, Sandgren et al. capacity to proliferate and restore liver function under circum-
developed a transgenic mouse model in which a protease, stances in which there is both massive and continuous liver
urokinase plasminogen activator (uPA), is expressed exclu- injury and the transplanted hepatocytes have a significant
sively in hepatocytes under control of the albumin (Alb) pro- selective advantage for survival compared with host hepato-
moter [22]. In this model, tissue protease activity caused cytes. Therefore, under the specialized circumstances existing
continuous and extensive liver injury and sub‐fulminant liver in the liver of Fah null mice, hepatocytes exhibit many proper-
failure, leading to death of the mice at four to six weeks of ties of stem cells, except for the ability to differentiate into
age. However, some mice survived and, in these there were more than one lineage, in this case into hepatocytes but not into
scattered nodules of normal liver tissue distributed throughout bile duct epithelial (BDE) cells.
the hepatic parenchyma (Figure  44.1a). This occurred by More recently, several additional murine liver repopulation
deletion or inactivation of the uPA transgene from individual models have been developed (reviewed in [28]). In AFC8 [29]
hepatocytes, which then clonally expanded into large clusters and TK‐NOG [30] transgenic mice selection can be induced by
that replaced damaged tissue. These findings prompted administration of a small molecule drug. Wild‐type hepatocytes
the  investigators to transplant normal hepatocytes (contain- can also repopulate the livers of mice expressing a mutant form
ing a β‐galactosidase marker gene) into uPA transgenic mice, of human alpha1‐antitrypsin [31]. Similarly, transgenic mice
44:  Liver Repopulation by Cell Transplantation and the Role of Stem Cells in Liver Biology 553

Figure 44.1  Liver regeneration and hepatocyte transplantation in the uPA transgenic mouse. (a) Wt mouse liver (left), regenerating uPA trans-
genic mouse liver by deletion of uPA transgene (middle), uPA transgenic liver (right). (b) LacZ transgenic mouse liver (left), uPA transgenic mouse
liver (middle), uPA transgenic mouse liver repopulated with LacZ hepatocytes (right). (c) Repopulation of the Fah null mouse liver by wt hepato-
cytes under NTBC administration. Immunohistochemistry with FAH staining (dark), 400× magnification: (left panel) 2 days, (middle) 3 weeks, and
(right) 6 weeks after cell transplantation.

overexpressing the p53 regulator mdm2 can be repopulated by inhibition of hepatocyte proliferation. However, the basic meta-
transplanted donor cells [32]. bolic functions of DNA damaged hepatocytes are maintained
In all of these genetic‐based liver repopulation models, suc- and the animals survive. Two to four weeks after retrorsine
cessful liver repopulation by adult hepatocytes required two administration, the animals are subjected to two‐thirds PH or
experimental conditions: (i) the liver was under massive and CCl4 administration in conjunction with transplantation of
continuous liver injury and (ii) the transplanted hepatocytes hepatocytes from normal animals. This leads to a brisk regen-
had a strong cell‐autonomous selective advantage for survival erative response specifically by transplanted hepatocytes and
in the host liver. In the Fah model, it has been shown that this there is near total liver repopulation in three to six months [34].
selective advantage is based on p21‐dependent cell cycle arrest Another method to achieve effective liver repopulation by
in host hepatocytes, resulting from p21 induction caused by transplanted hepatocytes is to induce DNA damage with
DNA damage [33]. Thus, an alternative strategy to obtain a ­selective irradiation of parts of the liver in conjunction with
high level of liver repopulation by transplanted hepatocytes is hepatocyte transplantation and either two‐thirds PH, CCl4
to block proliferation of endogenous hepatocytes using exoge- administration, or ischemic liver injury [35, 36]. In X‐irradi-
nous DNA‐damaging agents and then transplant normal hepat- ated rodents, administration of HGF has been used to replace
ocytes in conjunction with a liver proliferative stimulus. The PH as a liver regenerative stimulus [35]. As with retrorsine or
first method described to achieve this effect was to treat rats monocrotaline treatment of the host liver, transplanted hepato-
with retrorsine, a plant alkaloid that is taken up and metabo- cytes have a proliferative advantage over host hepatocytes in
lized selectively by hepatocytes to produce a DNA alkylating irradiated liver lobes. Recently, this approach is beginning to be
agent that cross‐links cellular DNA and disrupts hepatocyte explored in humans [37]. However, definitive data demonstrat-
division [34]. When retrorsine or a related compound, mono- ing selective liver repopulation in humans that underwent such
crotaline, is administered to rats or mice, there is a long‐lived liver conditioning have not yet been published.
554 THE LIVER:  STEM CELLS: THEIR ORIGIN, PROPERTIES, AND TRANSPLANTATION

XENOREPOPULATION MODELS improved surgical techniques and perioperative management


have vastly increased the percentage of harvested livers that
Extensive repopulation of the liver by transplanted hepato- can be used directly for transplantation. Therefore, considera-
cytes can also be achieved with human hepatocytes. Since ble attention has been focused on the potential use of a renew-
human cells are rejected by other animals, this requires the able and expandable cell source, such as stem cells, to
use of immune‐deficient recipients or pharmacological reconstitute liver mass. Several kinds of stem cells have been
immune suppression. Small rodents capable of harboring explored as potential hepatocyte precursors for transplanta-
human hepatocytes are not only of interest as a test bed for tion, including fetal liver stem cells, hepatocytes derived from
cell therapy, but also for a multitude of preclinical pharma- pluripotent stem cells, and adult liver stem/progenitor cells.
ceutical research applications, including models of drug While the existence of a bipotential stem cell capable of gen-
metabolism, infectious diseases (including malaria, hepatitis erating both cholangiocytes and hepatocytes in fetal life – the
B and C, gene therapy, and toxicology) [28]. Three murine hepatoblast – is well accepted, the notion of a true adult liver
models capable of supporting extensive repopulation of the stem cell is currently controversial. Much of the existing lit-
mouse liver with human hepatocytes have been reported erature on liver stem/progenitor cells can be explained by cell
(reviewed in [28]). All three models have been developed plasticity [41]. The current state of this field will be presented
commercially and are being used by the pharmaceutical in the following sections. First, the literature on hepatic stem
industry as well as academia in preclinical experimentation. cells will be reviewed. This will then be followed by a discus-
In 2001, Dandri et al. first showed that immune‐deficient uPA sion of cell plasticity as a possible alternative mechanism for
transgenic mice could be engrafted with human hepatocytes the published data, and recent studies using genetically modi-
and used as a model for hepatitis B [38]. Subsequently, this fied hepatocytes to increase their proliferative/repopulation
model was further developed to permit extensive repopula- potential [21].
tion, reaching levels as high as 90% human cells [39]. Fah During embryogenesis, the earliest stem cells originate from
knockout mice can also be repopulated extensively with the inner cell mass of the blastocyst and are pluripotent (capable
human cells from a variety of sources when they are crossed of differentiating into all the cell types in mammalian species);
onto severely immune deficient backgrounds (Figure  44.2) these cells are generally referred to as embryonic stem (ES)
[40]. This model has the advantage that the hepatic injury is cells [42]. These cells give rise to somatic stem cells that subse-
not constitutive but can be titrated by NTBC. Most recently, a quently differentiate into multipotent tissue‐specific stem cells
third xenorepopulation platform has emerged, the TK‐NOG [42–44]. The latter give rise to lineage‐committed progenitor
mouse [30]. All of these models are of potential use to assess cells that proliferate and differentiate into mature phenotypes
the therapeutic potential of transplanted cells, be they primary that ultimately become the somatic, tissue‐specific, working
human hepatocytes or stem cell derivatives. cells in different organs (Figure 44.3). Through studies of cell
transplantation to reconstitute the hematopoietic system and
studies of cell turnover in other tissues undergoing rapid and
continuous regeneration, such as skin and intestinal epithelium,
STEM CELLS: THEIR ORIGIN, stem cells have been shown to exhibit four essential properties:
PROPERTIES, AND TRANSPLANTATION (i) they have the capacity to maintain themselves (self‐renew),
while at the same time they generate progeny that differentiate
Given the promising results seen in experimental cell trans- into mature cellular phenotypes; this is referred to as asymmet-
plantation models, the source of cells that could be used clini- ric cell division [45, 46]; (ii) they are multipotent, that is, capa-
cally has been of considerable interest. Historically, ble of producing differentiated cells in at least two lineages; (iii)
hepatocytes isolated from adult donors were used for trans- their progeny are stable, reconstitute organ mass, and remain
plantation. However, in the clinical setting the availability of functional in the tissue for a long time; and (iv) by virtue of their
such cells from cadaveric donors is even more constrained capability of self‐renewal, stem cells can be transplanted seri-
than whole organs for orthotopic transplantation, because ally through successive hosts.

Figure 44.2  Repopulation of the Fah−/−/Rag2−/−/IL2rg−/− mouse liver by primary human hepatocytes at six weeks after cell transplantation, shown
at (a) low and (b) high magnification.
44:  Liver Repopulation by Cell Transplantation and the Role of Stem Cells in Liver Biology 555

Figure 44.3  Stem cells and tissue differentiation.

LIVER REPOPULATION BY FETAL LIVER In ED14 rat fetal liver, there are three distinct populations of
STEM/PROGENITOR CELLS epithelial cells, those positive for AFP and Alb but negative for
CK‐19, those positive for AFP, Alb, and CK‐19, and those posi-
During days 11–15 of embryonic development a bipotent liver tive for CK‐19 but negative for AFP and Alb [48]. The number
epithelial cell, the hepatoblast, can be found in rats and mice (see of AFP+/Alb+/CK‐19+ cells decreased dramatically at ED16,
Chapter  2). These hepatoblasts co‐express markers of adult after which liver repopulation potential of rat fetal liver cells
hepatocytes (albumin) and cholangiocytes (CK19) as well as also decreased significantly [47]. The level of liver repopulation
AFP. The ultimate test for a putative stem cell is to demonstrate by ED14 fetal liver cells under non‐selective conditions (i.e. in
its ability to self‐renew in vivo and to repopulate functionally a a normal liver) can also be increased to 20–25% by simply
tissue or organ, long‐term. Sandhu et  al. [47] reported 5–10% increasing the number of ED14 fetal liver cells transplanted
repopulation of DPPIV− mutant F344 rat liver by transplanting (Figure 44.4) [49]. Repopulation continues to increase for up to
wt ED14 fetal liver epithelial cells in conjunction with two‐thirds 1 year, reaching an average of around 30% for the total liver, and
PH. Liver repopulation by transplanted cells increased progres- remains stable for the life of the animal [50]. This represents a
sively over six months, and the bulk of repopulating clusters con- several thousand‐fold amplification of transplanted fetal liver
tained both hepatocytes and mature bile duct cells. The epithelial cells in the host organ. Both hepatic parenchymal
transplanted cells were integrated into the host parenchyma and cords and mature bile ducts are formed by transplanted fetal
formed hybrid bile canaliculi with host hepatocytes. Thus, trans- liver cells, and the progeny of the transplanted cells express nor-
planted rat ED14 fetal liver epithelial cells exhibited three major mal levels of hepatocytic and cholangiocytic genes in the
properties of liver stem cells: (i) extensive proliferation, (ii) bipo- respective cell types [49, 51]. Since serial transplantation has
tency, and (iii) long‐term repopulation in vivo [47]. Liver repopu- not yet been demonstrated with fetal liver epithelial cells, they
lation by transplanted rat fetal liver cells was achieved in a are referred to as fetal liver stem/progenitor cells (FLSPCs),
non‐selective host liver environment but required PH to initiate rather than stem cells.
the process. This is consistent with studies in hematopoietic stem The mechanism of liver repopulation by rat FLSPCs has
cells, in which hematopoietic reconstitution does not occur been shown to be cell competition between the transplanted
unless there is near total ablation of the host bone marrow. cells and host hepatocytes [49], a process that was originally
556 THE LIVER:  THE LIVER STEM CELL “NICHE”

Figure 44.4  Repopulation of the normal adult rat liver by fetal liver stem/progenitor cells. (a) Two examples of whole rat liver sections repopu-
lated by ED14 fetal liver cells. (b) Selected areas of repopulated liver at higher magnification showing both hepatocytes and mature bile duct
structures.

described in Drosophila during wing development [52, 53]. THE LIVER STEM CELL “NICHE”
These cells have been cryopreserved with full ability to repopu-
late the normal adult liver after thawing [54] and rat FLSPCs If bipotential liver stem cells do exist, the question remains as to
have been enriched to 95% purity by selection with immuno- where they reside [61]? The original idea of a stem cell “niche”
magnetic beads [51]. evolved from the concept that stem cells reside in tissues within
an “inductive microenvironment” that directs their differentia-
tion [63, 64]. More recently, the stem cell niche has been
described further as “a specific location in a tissue where stem
STEM CELLS IN THE ADULT LIVER cells can reside for an indefinite period and produce progeny
cells, while self‐renewing” [65]. Local stromal cells and other
Although many studies claim to have identified, isolated, and extracellular environmental factors attract stem cells to these
purified hepatic epithelial stem cells from the adult liver, much niches and affect their behavior (i.e. gene expression program,
of the evidence is derived from in vitro culture of clonogenic proliferation, and/or differentiation), and such niches have been
epithelial cells [55–60]. Whether these cells represent bona fide identified in the bone marrow, brain, skin, and intestinal mucosa
in vivo bipotential stem cells remains uncertain. In the mouse, [66–70].
all of the reported liver “stem cell” markers are also found in At present, the most likely candidate for a liver stem cell
cholangiocytes and therefore the existence of true stem cells niche is the canal of Hering (Figure 44.5). The canals of Hering
distinct from cholangiocytes in the adult mouse liver still were originally identified more than 100 years ago as luminal
remains to be established. Currently, the published data regard- channels linking the hepatocyte canalicular system to the biliary
ing the emergence of bipotential progenitors can all be system [71]. These channels contain small undifferentiated epi-
explained by cell plasticity, that is, fate conversion of hepato- thelial cells [72, 73] that are in direct physical continuity with
cytes to cholangiocytes and vice versa [41]. In the rat, however, hepatocytes at one membrane boundary and bile duct cells at
classic experiments are suggestive of a specialized adult liver another boundary to form duct‐like structures, enclosing a
stem cell. Unlike in the mouse, AFP becomes expressed in cells lumen (i.e. the canal; Figure  44.5b). In most models of liver
of ductal origin upon liver injury [61]. In conjunction with [3H] structure, the canals of Hering are depicted as very limited
thymidine pulse labeling and chase, AFP expression was shown structures confined to the portal space. However, studies con-
to decrease as [3H] labeled cells differentiated into basophilic ducted in humans have demonstrated that the canals of Hering
hepatocytes, which were strongly positive for albumin [62]. are much more elaborate than previously thought and extend far
Thus, AFP can be considered a specific marker for stem/pro- into the hepatic parenchyma [73, 74]. These structures contain
genitor cells in rats. epithelial cells with dual expression of bile ductular and fetal
44:  Liver Repopulation by Cell Transplantation and the Role of Stem Cells in Liver Biology 557

Figure 44.5  (a) Canal of Hering as the proposed liver stem cell “niche”. “Oval cells” within the canal are the precursors of both hepatocytes and
bile ducts. (b) Electron micrograph of the canal of Hering. The canal of Hering (*) is bounded by two mature hepatocytes (Hc) and one undifferenti-
ated epithelial cell (Ec).

hepatocytic markers (AFP, HepPar1, CK‐19) and were thus con- Oval cell induction methods were also reported for mice,
sidered to represent “facultative hepatic stem cells” [73, 75]. including a choline‐deficient (CD)/ethionine‐substituted diet [86,
87], treatment with dipin [88], or 3,5‐diethoxycarbonyl‐1,
4‐dihydrocollidine (DDC) [89]. However, their utility has been
controversial, because multiple groups have shown that “oval
“OVAL CELLS” AS HEPATOCYTE cells” emerging in these models generally lack bipotentiality [90].
PROGENITORS A role for “oval cells” in normal liver physiology and cell
turnover has not been established. However, as indicated above,
The term “oval cells” was coined by Farber [76] to describe non‐ “oval cells” are induced to proliferate when liver injury is super-
parenchymal cells in the periportal region that were ­present after imposed on circumstances in which hepatocyte proliferation is
treatment of rats with carcinogens: ethionine, α‐­acetylaminofluorene impaired. In the rat, these cells exhibit many features of pro-
(2‐AAF), and 3‐methyl‐4‐diethylaminobenzene. Other methods genitor cells, dividing rapidly and appearing to differentiate into
to induce proliferation of “oval cells” are to treat rats with both hepatocytes and BDE cells.
d‐galactosamine [77,78], or allyl alcohol [79]. In each of these Attempts to establish specific markers for “oval cells” to dis-
models, cells are induced in the adult liver that have a small, tinguish them from mature hepatocytes and BDE cells, and to
oval shaped, pale‐stained blue nucleus and very scant, lightly determine their lineage origin (mesoderm or endoderm), have
basophilic cytoplasm. Farber did not believe that “oval cells” led to conflicting findings. All investigators agree that “oval
were hepatocyte progenitors, but Thorgeirsson and co‐workers cells” express common liver epithelial progenitor cell markers,
[61] demonstrated that “oval cells” induced to ­proliferate in the such as AFP and Alb for hepatocyte progenitors and CK‐19 (and
periportal region after treatment of rats with 2‐AAF followed by OV6 in the rat) for bile duct progenitor cells. They were initially
two‐thirds PH subsequently differentiated into distinct clusters thought to express hematopoietic stem cell markers, c‐kit,
of basophilic hepatocytes. This was demonstrated by pulse labe- CD34, and Thy 1 [80, 81, 84, 92], but subsequent studies
ling of the liver with [3H]thymidine and following the progres- reported that both fetal liver progenitor cells and “oval cells” are
sion of label from periportal “oval cells” to hepatocytic clusters negative for these markers [75, 93, 94–96].
in the mid‐parenchyma in conjunction with the kinetic pattern If “oval cells” are indeed stem cells or hepatic progenitor cells,
of expression of bile ductular (CK‐7 and CK‐19) and hepato- they should be able to restore liver mass after their transplantation.
cytic (α‐fetoprotein [AFP] and Alb) markers over time [61, 62]. About 30 years ago, Faris and Hixson [97] reported that “oval
Other indirect evidence suggesting that “oval cells” are hepatic cells” isolated from the liver of rats fed a CD diet, treated with 2‐
progenitors is their expression of c‐kit [80], CD34 [81], flt3 AAF, and transplanted into the liver of secondary hosts produced
receptor [82], and LIF [83], all known to be expressed in hemat- “colonies” or clusters of cells with an hepatocytic phenotype in
opoietic stem cells or their immediate derivatives. recipients that had also been subjected to the CD diet but not in
Studies in the rat 2‐AAF/PH model have demonstrated that recipients that had received a normal diet. However, the level of
proliferating “oval cells” are indeed located in the canals of liver repopulation by transplanted CD/2‐AAF “oval cells” was not
Hering [84]. Thorgeirsson and colleagues [85] performed determined. “Oval cells” isolated from the liver of rats treated with
extensive immunohistochemical and ultrastructural studies d‐galactose (d‐Gal) also proliferate and differentiate into hepato-
in which they demonstrated that “oval cells,” induced to pro- cytes after their transplantation into rats undergoing two‐thirds PH
liferate by 2‐AAF/PH, are derived from undifferentiated cells [98]. Duct‐like epithelial cells isolated from the atrophic pancreas
in the canals of Hering, after which they pass through discon- of rats undergoing treatment with a copper chelating agent (trien)
tinuities in the laminar basement membrane of the ductal also proliferate modestly and differentiate into hepatocytes after
limiting plate and join together with stellate cells as they transplantation into normal rat liver [99]. Isolated pancreatic cells
enter the hepatic parenchyma, proliferate, and differentiate from normal mice also repopulate the liver of Fah null mice [100].
into hepatocytes. “Oval cells” isolated from the liver of DDC‐fed mice also
558 THE LIVER: PLASTICITY, TRANSDIFFERENTIATION, AND CLONOGENIC SUBSETS OF MATURE EPITHELIAL CELLS

repopulate the liver of Fah null mice, albeit with less efficiency humans [56, 57]. The cell of origin has a ductal phenotype and
than with mature hepatocytes [101]. Similarly, “oval cells” from liver organoids are extensively cholangiocytic in phenotype.
GFP transgenic mice maintained on a DDC diet repopulate the Nonetheless, they can be induced to express hepatocyte markers
liver of wt mice treated with monocrotaline in conjunction with in culture [56, 57]. Upon transplantation, they can produce bona
PH [102]. Other studies also showed effective repopulation of the fide hepatocytes, but the frequency is very low. The future poten-
liver by purified “oval cells” in both rats treated with retrorsine tial of clonally derived Lgr5+ cells is discussed in Chapter 77.
[75] and in Fah null mice [60], but not in animals with a normal
liver. Numerous studies have reported the isolation of “oval cell”
lines from mice and rats, and also from humans, that are clonal,
PLASTICITY, TRANSDIFFERENTIATION,
bipotent, and exhibit other stem and progenitor cell characteristics
in vitro and in vivo (for a review see [18]). This has provided valu- AND CLONOGENIC SUBSETS
able information concerning the basic biological properties of OF MATURE EPITHELIAL CELLS
these cell lines, but, in general, in vivo repopulation by trans-
planted “oval cell” lines has been very low. Much of the data on liver injury responses in the historic litera-
ture has been interpreted as either proliferation of existing
mature epithelial cells (hepatocytes or cholangiocytes) or acti-
vation of facultative stem cells. More recently, however, the
HUMAN “OVAL CELLS” AND phenomenon of plasticity and transdifferentiation between
STEM CELLS ductal epithelial cells and hepatocytes has become more appre-
ciated and experimentally proven. Several studies showed con-
A human counterpart to “oval cell” activation has been described clusively that proliferating ductal cells, previously considered to
in liver tissue obtained from patients with extensive chronic be derived from facultative stem cells, were in fact not the
liver injury or submassive hepatic necrosis, that is, the so‐called source of parenchymal regeneration in several classical models
“ductular reaction” (for a detailed description, see [103]). of oval cell injury in the mouse [90, 114–116]. To the contrary,
“Ductular reactions” are comprised of collections of cells in it has been shown clearly that mature hepatocytes can efficiently
ductular arrays with the morphological appearance and immu- convert into cholangiocyte‐like cells [115] and even form a
nohistochemical markers comparable to those found in rodent functional biliary tree [117]. Hepatocyte‐derived ducts can
“oval cells.” These cells are present primarily in the portal tracts extensively proliferate and retain the ability to reconvert to their
with extension into the parenchyma and express both hepato- cell of origin, the hepatocytes, once the injury subsides [115].
cytic and bile ductular markers, and also certain neuroendocrine The ability of hepatocytes to differentiate into BDE cells has
genes [103–106]. Using double and triple label immunohisto- also been well demonstrated in several rat models under injury
chemistry, Zhou et  al. [107] have shown that “ductular reac- conditions [118, 119]. These studies have taken advantage of
tions” are bipolar structures with cells at one pole exhibiting recently developed lineage tracing techniques to identify the
hepatocytic morphology and gene expression (HepPar1 or origin of cells in liver tissue under different physiologic and
HepPar1/NCAM) and cells at the other pole exhibiting biliary pathophysiologic states (see Chapter 85).
morphology and gene expression (CK‐19 or CK‐19/NCAM), Taken together these findings raise the question, whether the
with undifferentiated epithelial cells in the center expressing bipotential stem/progenitor cells and oval cells previously
only NCAM. Cells with similar morphological and immunohis- reported may actually have also been transdifferentiated hepato-
tochemical properties have also been identified in the human cytes, not stem cells of ductal origin. Recent papers have taken
fetal liver beginning at four weeks gestation [108]. advantage of this plasticity of hepatocytes. While mature hepat-
A number of investigators have isolated, cultured, and/or pas- ocytes cannot be efficiently expanded in culture, transdifferenti-
saged human fetal liver epithelial cells with bipotent properties, ated hepatocytes can be grown extensively [120]. Excitingly,
and several of these studies have demonstrated their differentia- these cells retain the ability to redifferentiate into mature hepat-
tion into hepatocytes after transplantation into SCID or nude ocytes and function in transplantation.
mice [109–111]. Schmelzer et al. [55] have identified two popu- While non‐hepatocytes are clearly not the source of new
lations of clonogenic hepatic epithelial cells from human fetal, hepatocytes in most of the classic mouse oval cell literature,
neonatal, and pediatric liver that exhibit stem cell properties. newer injury models which involve activation of p21 and senes-
While these cells are progenitor‐like in vitro and can extensively cence in mature hepatocytes have shown that cholangiocytes
expand, their ability to efficiently give rise to functional hepato- can become hepatocytes [32, 121, 122]. Thus, transdifferentia-
cytes either in vitro or upon transplantation remains unproven. tion has now been shown to work in both directions. If cholan-
In the intestine, a self‐renewing classic stem cells resides at giocytes can produce hepatocytes and vice versa, the question
the bottom of intestinal crypts and can be defined by expression arises whether typical stem/progenitor cells exist in the adult
of the wnt‐target gene Lgr5 [112]. Clevers and co‐workers devel- liver or whether all regeneration can be explained by plasticity
oped tissue culture conditions which permit the clonal expansion of mature epithelial cells.
of these stem cells and were able to obtain all mature intestinal An important question arising from the concept of plastic-
cell types in “organoids” derived from single Lgr5+ cells [113]. ity is whether all mature epithelial cells are equally prolifera-
They applied similar tissue culture conditions to other endoder- tive and plastic. Recently, several labs have demonstrated
mal tissues, including liver, stomach, and pancreas and were able heterogeneity within hepatocytes in the adult mouse [123,
to grow immortal Lgr5+ liver organoids from both mice and 124], once again giving rise to claims of having discovered
44:  Liver Repopulation by Cell Transplantation and the Role of Stem Cells in Liver Biology 559

“the” liver stem cell. In addition, significant differences in the retrorsine treatment. These studies are promising, but the ability
proliferative ability of cholangiocyte subsets have also been of mesenchymal stem cells to repopulate the adult liver under
found [125]. more normal, clinically viable circumstances has not been
established.

LIMITED LIVER REPOPULATION BY


EXTRAHEPATIC AND EMBRYONIC PLURIPOTENT STEM CELLS
STEM CELLS (ES and iPSC)
Various studies have reported that cells released from the bone Because of their extensive proliferative capacity, pluripotent
marrow (BM) into the circulation, migrate to the liver and dif- stem cells are an attractive potential source of transplantable
ferentiate into hepatocytes. However, the extent to which this hepatocytes. Not only can these cells divide indefinitely, but
occurs and the mechanism(s) involved remain highly controver- they also retain the ability to differentiate into multiple different
sial (for reviews, see [126–128]). Estimates of liver repopula- mature cell types [42]. Until 2007, ES cells were the sole source
tion by hematopoietic cells vary widely, ranging from less than of pluripotent human cells and are ethically controversial. Now,
0. 01 to 40%. Originally, Petersen et al. reported that BM stem however, pluripotent cells can be derived from direct genetic
cells from DPPIV+ F344 rats transplanted into sublethally irra- reprogramming of somatic cell types, such as dermal fibroblasts
diated DPPIV− F344 rats repopulate the BM after which they [157, 158]. Pluripotent stem cells (both ES and iPSC) cells in
migrate to the liver and “transdifferentiate” into hepatocytes culture can be induced along the endodermal and hepatocytic
through the liver “oval cell” progenitor pathway [129]. This lineages by addition of specific cytokines and growth factors
mechanism was generally accepted until studies by Wang et al. [159–166]. The first step typically involves the induction of
[101] using lacZ marking showed that BM cells did not enter the definitive endoderm using activin A. Numerous slightly differ-
“oval cell” pool in wt mice treated with DDC or contribute to ent step‐wise differentiation protocols have been developed by
liver repopulation by “oval cells” in secondary Fah−/− mouse different labs with the aim of generating mature hepatocytes as
recipients. Menthena et  al. [130] also showed in rats that the final product [159, 166]. Historically, cells produced in this
DPPIV+ BM cells transplanted into DPPIV− rats contributed less fashion expressed typical markers such as Alb, but usually also
than 1% to “oval cells” derived from three different model sys- express AFP. More recent protocols yield cells that have silenced
tems: (i) 2‐AAF/PH, (ii) retrorsine/PH, and (iii) d‐Gal induced AFP and express multiple mature hepatocyte markers. Despite
liver injury. this, genome‐wide expression analysis demonstrates that these
In Fah−/− mice as well as other model systems, it has been hepatocyte‐like cells are not fully mature and are significantly
demonstrated that cell fusion and reprogramming, rather than different from primary human hepatocytes. Such cells can be
transdifferentiation, is the mechanism by which hematopoi- transplanted into the liver of immune deficient recipients with
etic cells acquire an hepatocytic phenotype. Initial studies in differentiation into cells expressing mature hepatocyte [165,
cell culture showed that BM and neuronal cells can fuse with 166] and cholangiocyte markers [166]. Nonetheless, the level of
ES cells [131, 132]. Wang et al. [133] and Vassilopoulos et al. liver repopulation obtained with hepatocyte‐differentiated,
[134] subsequently showed that hematopoietic stem cells fuse pluripotent stem cells is generally low and functional tests, such
with hepatocytes in Fah null mice to produce cells expressing as blood albumin measurements, fail to demonstrate equiva-
the deficient enzyme, which then expand massively to restore lence to primary hepatocytes [167]. The repopulation level is
liver mass and function [133, 134]. Fusion also occurs somewhat higher when the cells are transplanted into MUP‐
between hematopoietic cells and neurons or muscle cells uPA/SCID mice [166]. To date, all pluripotent stem cell differ-
[135, 136] and it has been shown that myelomonocytic cells entiation protocols generate only “hepatocyte‐like” cells, but
can fuse with hepatocytes [137, 138] or muscle cells [139] to not fully functional, mature, and transplantable equivalents of
produce somatic hybrids expressing genes from both parental hepatocytes isolated from adult liver. More mature “hepatocyte‐
cell types. like” cells have been selected using surface markers, such as the
Other studies reported that fusion does not appear to be asialoglycoprotein receptor for purification [167], but even this
required for BM‐derived cells to differentiate into hepatocytes approach does not yield fully mature donor cells. In the future,
[140–142]. Unfractionated or CD34+ enriched cells from human it is hoped that conditions will be developed in which lineage‐
cord blood [143–146], multipotent adult progenitor cells specified pluripotent derivatives will be therapeutically equiva-
(MAPCs) [147, 148], or mesenchymal stem cells [149–153] lent to primary cells.
have been transplanted into the liver of immunodeficient mice.
These transplanted cells express a differentiated hepatocytic
phenotype [143, 154], but liver repopulation was once again
very low. Several studies have reported that mesenchymal stem DIRECT REPROGRAMMING
cells, isolated from adipose tissue and differentiated in culture
along the hepatocytic lineage, can also engraft in the liver paren- Since it is possible to generate hepatocyte‐like cells from der-
chyma and contribute to liver regeneration [155, 156]. One mal fibroblasts via a pluripotent intermediate stage, several
study [156] reported large repopulation clusters with hepato- groups have reported direct reprogramming of somatic cells into
cyte‐differentiated mesenchymal stem cells, but this required hepatocyte‐like cells, terming them iHeps, induced hepatocytes
560 THE LIVER: FUTURE HORIZONS

[168–172]. The concept of direct reprogramming using a cock- to a modified estrogen receptor gene (ERT2), whose function
tail of cell‐type specific transcription factors was first developed could be controlled be tamoxifen. Thus, hYapERT2 expression/
for neurons [173]. For the liver, a combination Hnf4α plus function and liver repopulation would be regulated by tamox-
Foxa1, Foxa2, or Foxa3 was successfully used to directly con- ifen administration [21]. The hYapERT2 sequence was incor-
vert fibroblasts into transplantable hepatocyte‐like cells [171]. porated into a lentivirus vector which produced stable Yap
Soon thereafter, the generation of iHeps from human cells using expression after transduction of normal adult hepatocytes.
a similar combination of factors (FOXA3, HNF1α, and HNF4α) Liver repopulation in the DPPIV‐ Fischer (F) 344 rat transplan-
was also reported [170]. While transplantation success has been tation system produced 8–10% liver repopulation after six
reasonably good with human iHeps in rodents and other ani- months on continuous tamoxifen feeding (Figure  44.6).
mals, the cells in the repopulated liver are not fully differenti- Repopulation was tightly controlled by tamoxifen administra-
ated [174]. Nonetheless, these cells may be functional enough to tion, transplanted hepatocytes expanded into large clusters of
be of use in an extracorporeal bioartificial liver device (BAL) morphologically normal hepatocytes and integrated into host
(Lijian Hui, personal communication) and further developments hepatic parenchymal plates with no evidence for dysplasia,
may eventually render them useful for in vivo human transplan- dedifferentiation or tumorigenesis.
tation in the future. A major concern in using Yap to stimulate proliferation of
cells is potential tumor risk. A very recent study by Peterson
et  al. [180] showed increased repopulation by transplanted
hYapERT2 transduced hepatocytes on continuous tamoxifen
OTHER ROLES FOR BM CELLS IN feeding for one year in the DPPIV model system with 10–20%
LIVER REGENERATION repopulation, no evidence for tumorigenesis, and a cell com-
petition mechanism for repopulation, as previously found with
Several studies have reported that injections of BM‐derived ED14 fetal liver cells [49]. Using hYapERT2 transduced
stem cells can restore liver function during chronic liver injury Wistar RHA rat hepatocytes transplanted into the hyperbiliru-
by enhancing the degradation of liver fibrosis in mice [175, binemic Gunn rat liver, there was comparable liver repopula-
176]. This has been associated with induction of metalloprotein- tion with a progressive 70–80% reduction in serum bilirubin
ases, especially MMPs 2, 9, and 13 [177]. It has been reported on continuous tamoxifen feeding for six months. If tamoxifen
that BM‐derived endothelial progenitor cells (EPCs), injected feeding was initiated six months after hepatocyte transplanta-
into the spleen during liver injury, engraft in the liver, form new tion, serum bilirubin levels dropped progressively over the
blood vessels, and secrete growth factors, such as HGF, TGFα, next six months with the same slope as observed when tamox-
EGF, and VEGF, stimulating liver regeneration and improving ifen administration was initiated at the time of cell transplanta-
survival of animals with massive liver injury [178]. Thus, the tion. If DPPIV− rats transplanted with hYapERT2 transduced
role of BM stem cells in liver regeneration may be supportive in wt hepatocytes were taken off tamoxifen after six months of
generating new parenchymal mass and, under some circum- liver repopulation, the liver repopulation level did not decrease
stances, in ameliorating hepatic fibrosis. after an additional six months on a normal diet; thus, trans-
planted hepatocytes remained in the liver long‐term. These
studies provide hope that transplantation of genetically modi-
fied hepatocytes will provide a therapeutic avenue to treat
USE OF EX VIVO GENETICALLY selected liver diseases.
MODIFIED HEPATOCYTES
TO REPOPULATE THE LIVER
IN A NORMAL HEPATIC FUTURE HORIZONS
MICROENVIRONMENT
Many questions regarding the basic biology of stem cells in the
As noted previously, bipotent fetal hepatoblasts can repopulate liver remain unresolved, including the all‐important question of
the liver under normal physiologic circumstances. Because of whether hepatic stem cells exist at all in the adult liver. Most
ethical concerns, use of adult hepatocytes for this purpose would data suggest that it is possible to generate new hepatocytes
be preferred. Thus far, however, significant liver repopulation from cells other than hepatocytes themselves and that there is
by adult hepatocytes has required extensive and ongoing genetic, an intrahepatic source of these precursors. Their precise nature
physical, or chemical injury to the host liver. A substantial num- and location, however, remains unclear, as do the molecular
ber of monogenic liver disorders in which there is no underlying mechanisms that lead to their activation. Much of this uncer-
or ongoing liver injury could be treated by transplantation of tainly is related to the lack of markers that would permit the
normal adult hepatocytes, if the cells to be transplanted were dissection of the complex cellular heterogeneity of the liver.
first genetically modified to provide a proliferative and/or sur- However, cell surface markers, and also genetic lineage tracing
vival advantage compared to host hepatocytes. tools, are now available to address these issues. Thus, biologi-
Shafritz and coworkers determined that Yap1 is hyperex- cal studies of antigenically defined and genetically marked
pressed in ED14 rat fetal liver cells, as well as the anti‐apop- cells are now feasible and are likely to solve some of these
totic BIRC5 (survivin) gene [21]. Based on these findings, they dilemmas. In addition, pluripotent stem cells now afford the
transduced normal adult hepatocytes with an hYap gene linked opportunity for “developmental biology in a dish,” making
44:  Liver Repopulation by Cell Transplantation and the Role of Stem Cells in Liver Biology 561

Figure 44.6  (a) Hepatocyte transplantation protocol. (b) Detection of lentivirus TTR‐hYapERT2 transduced hepatocytes in DPPIV− host liver by
DPPIV enzyme histochemistry. Clusters of DPPIV+ cells are clearly visible at three months after cell transplantation in tamoxifen fed rats (b1). In
the absence of tamoxifen feeding, clusters of transplanted cells are not apparent (b2). At six months after transplantation of lenti TTR‐hYapERT2
transduced hepatocytes, clusters of DPPIV+ cells are much larger in tamoxifen fed rats, and in some instances comprise whole liver lobules (b3). In
the absence of tamoxifen feeding, transplanted DPPIV+ cells are still not visible (b4). (c) Quantification of liver repopulation by scanning digital
images was performed in DPPIV− rats transplanted with lenti TTR‐hYapERT2 transduced hepatocytes, lenti TTR‐GFP transduced hepatocytes, or
non‐transduced hepatocytes, all maintained on tamoxifen feeding (+Tam) or on rats transplanted with lenti TTR‐hYapERT2 transduced hepatocytes
maintained on normal chow (−Tam).

attractive models to study molecular mechanisms of hepatic functional hepatocytes, and maintain differentiated hepatic
lineage development. function, long‐term [50]. This requires a liver proliferative
Although substantial progress has been made during the past stimulus or liver injury at the time the cells are transplanted,
10–15 years concerning the possibility of liver repopulation by but no selective advantage other than their ability to replace
transplanted cells, much still needs to be learned. Factors gov- host hepatocytes by cell competition. Use of stem cells from
erning engraftment of transplanted cells into the liver and their pediatric or adult cadaveric liver or from other sources, such as
homing to the correct niche, factors regulating proliferation and bone marrow, cord blood, ES cells, or iPS cells, is a possibil-
differentiation of transplanted cells into specific phenotypes ity, and also cultured human fetal cells or adult hepatocytes
required for organ function, and specific host conditions under modified to favor engraftment and proliferation in the host, but
which effective liver replacement can be achieved, all need to be this will require substantial additional research. Cell lines have
determined. been established, including ES cells, iPS cells, fetal liver cells,
The best starting point for therapeutic liver repopulation will and “oval (progenitor) cells” that also exhibit stem cell proper-
probably be a genetic disorder with ongoing liver injury, which ties and differentiate into hepatocytes and/or bile ducts in vitro
will hopefully induce or augment proliferation of transplanted and in vivo. However, these cell lines have shown only limited
cells in the host liver. A good example of such a condition is repopulation of the normal liver at the current state of the art.
Wilson’s disease, in which transplanted cells might also have a In order to advance the field of liver cell therapy further, it will
modest selective advantage [179], since they will not store high be necessary to find conditions under which cells and cell lines
levels of copper. Another example is α1‐antitrypsin deficiency, derived from ES cells, iPS cells, fetal liver, or adult liver can
in which a mutated form of α1‐antitrypsin is not secreted from be expanded in culture and successfully repopulate the liver
the cell and causes liver injury but much less severe than in uPA under conditions that will be clinically acceptable. Use of
transgenic mice. Most encouraging is the recent success in newly emerging three dimensional or “organoid” culture con-
repopulating the liver with normal hepatocytes in a mouse ditions to maintain and expand hepatic derived cells or cells in
model of a1‐antitrypsin deficiency [31]. the hepatic lineage will help to advance this field (see
To date, very few patients have been transplanted with fetal Chapter 77). Although we are not yet there, restoration of liver
liver cells. Studies in rats have shown that these cells have the function by therapeutic cell transplantation holds great prom-
capacity to proliferate in the host, replace hepatic mass with ise for the future.
562 THE LIVER:  REFERENCES

ACKNOWLEDGMENTS 23. Rhim, J.A., Sandgren, E.P., Degen, J.L. et  al. Replacement of diseased
mouse liver by hepatic cell transplantation. Science, 1994;263:1149–52.
24. Overturf, K., Al‐Dhalimy, M., Tanguay, R. et al. Hepatocytes corrected by
The authors would like to thank Anna Caponigro for assistance gene therapy are selected in vivo in a murine model of hereditary tyrosinae-
in preparing the manuscript and Dr Irmin Sternlieb, a former mia type I. Nat Genet, 1996;12:266–73.
colleague at Albert Einstein College of Medicine, who provided 25. Lindstedt, S., Holme, E., Lock, E.A. et al. Treatment of hereditary tyrosinae-
the electron micrograph illustrating a canal of Hering to D.A.S. mia type I by inhibition of 4‐hydroxyphenylpyruvate dioxygenase. Lancet,
1992;340:813–7.
26. Overturf, K., al‐Dhalimy, M., Ou, C.N. et al. Serial transplantation reveals
the stem‐cell‐like regenerative potential of adult mouse hepatocytes. Am J
Pathol, 1997;151:1273–80.
REFERENCES 27. Wang, M.J., Chen, F., Li, J.X. et al. Reversal of hepatocyte senescence after
continuous in vivo cell proliferation. Hepatology, 2014;60:349–61.
  1. Boeck, J. and Popper, H. Ueber Lebertransplantation in die Vorderkammer 28. Grompe, M. and Strom, S. Mice with human livers. Gastroenterology,
des Auges. Virchow Arch Path Anat, 1937;299:219–234. 2013;145:1209–14.
  2. Matas, A.J., Sutherland, D.E., Steffes, M.W. et al. Hepatocellular transplan- 29. Washburn, M.L., Bility, M.T., Zhang, L. et al. A humanized mouse model to
tation for metabolic deficiencies: decrease of plasms bilirubin in Gunn rats. study hepatitis C virus infection, immune response, and liver disease.
Science, 1976;192:892–4. Gastroenterology, 2011;140:1334–44.
  3. Grompe, M. Liver repopulation for the treatment of metabolic diseases. J 30. Hasegawa, M., Kawai, K., Mitsui, T. et  al. The reconstituted “humanized
Inherit Metab Dis, 2001;24:231–44. liver” in TK‐NOG mice is mature and functional. Biochem Biophys Res
  4. Oishi, K., Arnon, R., Wasserstein, M.P. et al. Liver transplantation for pediat- Commun, 2011;405:405–10.
ric inherited metabolic disorders: Considerations for indications, complica- 31. Ding, J., Yannam, G.R., Roy‐Chowdhury, N. et  al. Spontaneous hepatic
tions, and perioperative management. Pediatr Transplant, 2016;20:756–69. repopulation in transgenic mice expressing mutant human alpha1‐antitrypsin
  5. Strom, S.C., Fisher, R.A., Thompson, M.T. et al. Hepatocyte transplantation by wild‐type donor hepatocytes. J Clin Invest, 2011;121:1930–4.
as a bridge to orthotopic liver transplantation in terminal liver failure. 32. Lu, W.Y., Bird, T.G., Boulter, L. et  al. Hepatic progenitor cells of biliary
Transplantation, 1997;63:559–69. origin with liver repopulation capacity. Nat Cell Biol, 2015;17:971–83.
  6. Bilir, B.M., Guinette, D., Karrer, F. et al. Hepatocyte transplantation in acute 33. Willenbring, H., Sharma, A.D., Vogel, A. et al. Loss of p21 permits carcino-
liver failure. Liver Transpl, 2000;6:32–40. genesis from chronically damaged liver and kidney epithelial cells despite
  7. Grossman, M., Rader, D.J., Muller, D.W. et al. A pilot study of ex vivo gene unchecked apoptosis. Cancer Cell, 2008;14:59–67.
therapy for homozygous familial hypercholesterolaemia. Nat Med, 34. Laconi, E., Oren, R., Mukhopadhyay, D.K. et al. Long‐term, near‐total liver
1995;1:1148–54. replacement by transplantion of isolated hepatocytes in rats treated with ret-
  8. Bucher, N.L. and Swaffield, M.N. The rate of incorporation of labeled thy- rorsine. Am J Pathol, 1998;153:319–29.
midine into the deoxyribonucleic acid of regenerating rat liver in relation to 35. Guha, C., Sharma, A., Gupta, S. et  al. Amelioration of radiation‐induced
the amount of liver excised. Cancer Res, 1964;24:1611–25. liver damage in partially hepatectomized rats by hepatocyte transplantation.
  9. Grisham, J.W. A morphologic study of deoxyribonucleic acid synthesis and Cancer Res, 1999;59:5871–4.
cell proliferation in regenerating rat liver; autoradiography with thymidine‐ 36. Malhi, H., Gorla, G.R., Irani, A.N. et  al. Cell transplantation after
H3. Cancer Res, 1962;22:842–9. ­oxidative hepatic preconditioning with radiation and ischemia‐reperfusion
10. Rajvanshi, P., Kerr, A., Bhargava, K.K. et al. Studies of liver repopulation leads  to extensive liver repopulation. Proc Natl Acad Sci USA,
using the dipeptidyl peptidase IV‐deficient rat and other rodent recipients: 2002;99:13114–9.
cell size and structure relationships regulate capacity for increased trans- 37. Soltys, K.A., Setoyama, K., Tafaleng, E.N. et  al. Host conditioning and
planted hepatocyte mass in the liver lobule. Hepatology, 1996;23:482–96. rejection monitoring in hepatocyte transplantation in humans. J Hepatol,
11. Rajvanshi, P., Kerr, A., Bhargava, K.K. et al. Efficacy and safety of repeated 2017;66:987–1000.
hepatocyte transplantation for significant liver repopulation in rodents. 38. Dandri, M., Burda, M.R., Torok, E. et al. Repopulation of mouse liver with
Gastroenterology, 1996;111:1092–1102. human hepatocytes and in vivo infection with hepatitis B virus. Hepatology,
12. Rozga, J., Holzman, M., Moscioni, A.D. et al. Repeated intraportal hepato- 2001;33:981–8.
cyte transplantation in analbuminemic rats. Cell Transplant, 39. Tateno, C., Yoshizane, Y., Saito, N. et al. Near completely humanized liver in
1995;4:237–43. mice shows human‐type metabolic responses to drugs. Am J Pathol,
13. Fox, I.J., Chowdhury, J.R., Kaufman, S.S. et al. Treatment of the Crigler– 2004;165:901–12.
Najjar syndrome type I with hepatocyte transplantation. N Engl J Med, 40. Azuma, H., Paulk, N., Ranade, A. et al. Robust expansion of human hepato-
1998;338:1422–6. cytes in Fah‐/‐/Rag2‐/‐/Il2rg‐/‐ mice. Nat Biotechnol, 2007;25:903–10.
14. Strom, S.C., Chowdhury, J.R., Fox, I.J. Hepatocyte transplantation for the 41. Kopp, J.L., Grompe, M., and Sander, M. Stem cells versus plasticity in liver
treatment of human disease. Semin Liver Dis, 1999;19:39–48. and pancreas regeneration. Nat Cell Biol, 2016;18:238–45.
15. Iansante, V., Chandrashekran, A. and Dhawan, A. Cell‐based liver therapies: 42. Thomson, J.A., Itskovitz‐Eldor, J., Shapiro, S.S. et al. Embryonic stem cell
past, present and future. Philos Trans R Soc Lond B Biol Sci, 2018;373. lines derived from human blastocysts. Science, 1998;282:1145–7.
16. Squires, J.E., Soltys, K.A., McKiernan, P. et al. Clinical hepatocyte trans- 43. Fuchs, E. and Segre, J.A. Stem cells: a new lease on life. Cell,
plantation: what is next? Curr Transplant Rep, 2017;4:280–289. 2000;100:143–55.
17. Michalopoulos, G.K. Advances in liver regeneration. Expert Rev 44. Weissman, I.L. Stem cells: units of development, units of regeneration, and
Gastroenterol Hepatol, 2014;8:897–907. units in evolution. Cell, 2000;100:157–68.
18. Miyajima, A., Tanaka, M., and Itoh, T. Stem/progenitor cells in liver devel- 45. Marshman, E., Booth, C., and Potten, C.S. The intestinal epithelial stem cell.
opment, homeostasis, regeneration, and reprogramming. Cell Stem Cell, Bioessays, 2002;24:91–8.
2014;14:561–74. 46. Lechler, T. and Fuchs, E. Asymmetric cell divisions promote stratification
19. Dong, J., Feldmann, G., Huang, J. et al. Elucidation of a universal size‐con- and differentiation of mammalian skin. Nature, 2005;437:275–80.
trol mechanism in Drosophila and mammals. Cell, 2007;130:1120–33. 47. Sandhu, J.S., Petkov, P.M., Dabeva, M.D. et  al. Stem cell properties and
20. Camargo, F.D., Gokhale, S., Johnnidis, J.B. et al. YAP1 increases organ size repopulation of the rat liver by fetal liver epithelial progenitor cells. Am J
and expands undifferentiated progenitor cells. Curr Biol, 2007;17:2054–60. Pathol, 2001;159:1323–34.
21. Yovchev, M., Jaber, F.L., Lu, Z. et al. experimental model for successful liver 48. Dabeva, M.D., Petkov, P.M., Sandhu, J. et al. Proliferation and differentia-
cell therapy by lenti TTR‐YapERT2 transduced hepatocytes with tamoxifen tion of fetal liver epithelial progenitor cells after transplantation into adult rat
control of Yap subcellular location. Sci Rep, 2016;6:19275. liver. Am J Pathol, 2000;156:2017–31.
22. Sandgren, E.P., Palmiter, R.D., Heckel, J.L. et al. Complete hepatic regen- 49. Oertel, M., Menthena, A., Dabeva, M.D. et al. Cell competition leads to a
eration after somatic deletion of an albumin‐plasminogen activator transgene. high level of normal liver reconstitution by transplanted fetal liver stem/pro-
Cell, 1991;66:245–56. genitor cells. Gastroenterology, 2006;130:507–20.
44:  Liver Repopulation by Cell Transplantation and the Role of Stem Cells in Liver Biology 563

50. Shafritz, D.A. and Oertel, M. Model systems and experimental conditions   79. Yin, L., Lynch, D., and Sell, S. Participation of different cell types in the
that lead to effective repopulation of the liver by transplanted cells. Int J restitutive response of the rat liver to periportal injury induced by allyl alco-
Biochem Cell Biol, 2011;43:198–213. hol. J Hepatol, 1999;31:497–507.
51. Oertel, M., Menthena, A., Chen, Y.Q. et al. Purification of fetal liver stem/   80. Fujio, K., Evarts, R.P., Hu, Z. et al. Expression of stem cell factor and its
progenitor cells containing all the repopulation potential for normal adult rat receptor, c‐kit, during liver regeneration from putative stem cells in adult
liver. Gastroenterology, 2008;134:823–32. rat. Lab Invest, 1994;70:511–6.
52. Moreno, E. and Basler, K. dMyc transforms cells into super‐competitors.   81. Omori, N., Omori, M., Evarts, R.P. et al. Partial cloning of rat CD34 cDNA
Cell, 2004;117:117–29. and expression during stem cell‐dependent liver regeneration in the adult
53. de la Cova, C., Abril, M., Bellosta, P. et al. Drosophila myc regulates organ rat. Hepatology, 1997;26:720–7.
size by inducing cell competition. Cell, 2004;117:107–16.   82. Omori, M., Omori, N., Evarts, R.P. et al. Coexpression of flt‐3 ligand/flt‐3
54. Oertel, M., Menthena, A., Chen, Y.Q. et al. Properties of cryopreserved fetal and SCF/c‐kit signal transduction system in bile‐duct‐ligated SI and W
liver stem/progenitor cells that exhibit long‐term repopulation of the normal mice. Am J Pathol, 1997;150:1179–87.
rat liver. Stem Cells, 2006;24:2244–51.   83. Omori, N., Evarts, R.P., Omori, M. et al. Expression of leukemia inhibitory
55. Schmelzer, E., Wauthier, E., and Reid, L.M. The phenotypes of pluripotent factor and its receptor during liver regeneration in the adult rat. Lab Invest,
human hepatic progenitors. Stem Cells, 2006;24:1852–8. 1996;75:15–24.
56. Huch, M., Dorrell, C., Boj, S.F. et al. In vitro expansion of single Lgr5+ liver   84. Crosby, H.A., Kelly, D.A., and Strain, A.J. Human hepatic stem‐like cells
stem cells induced by Wnt‐driven regeneration. Nature, 2013:1–6. isolated using c‐kit or CD34 can differentiate into biliary epithelium.
57. Huch, M., Gehart, H., van Boxtel, R. et al. Long‐term culture of genome‐sta- Gastroenterology, 2001;120:534–44.
ble bipotent stem cells from adult human liver. Cell, 2015;160:299–312.   85. Paku, S., Schnur, J., Nagy, P. et al. Origin and structural evolution of the
58. Dorrell, C., Erker, L., Schug, J. et al. Prospective isolation of a bipotential clo- early proliferating oval cells in rat liver. Am J Pathol, 2001;158:1313–23.
nogenic liver progenitor cell in adult mice. Genes Dev, 2011;25:1193–203.   86. Sells, M.A., Katyal, S.L., Shinozuka, H. et al. Isolation of oval cells and
59. Shin, S., Walton, G., Aoki, R. et al. Foxl1‐Cre‐marked adult hepatic progeni- transitional cells from the livers of rats fed the carcinogen DL‐ethionine. J
tors have clonogenic and bilineage differentiation potential. Genes Dev, Natl Cancer Inst, 1981;66:355–62.
2011;25:1185–92.   87. Akhurst, B., Croager, E.J., Farley‐Roche, C.A. et al. A modified choline‐
60. Suzuki, A., Sekiya, S., Onishi, M. et al. Flow cytometric isolation and clonal deficient, ethionine‐supplemented diet protocol effectively induces oval
identification of self‐renewing bipotent hepatic progenitor cells in adult cells in mouse liver. Hepatology, 2001;34:519–22.
mouse liver. Hepatology, 2008;48:1964–78.   88. Factor, V.M., Radaeva, S.A., Thorgeirsson, S.S. Origin and fate of oval cells
61. Evarts, R.P., Nagy, P., Marsden, E. et  al. A precursor‐product relationship in dipin‐induced hepatocarcinogenesis in the mouse. Am J Pathol,
exists between oval cells and hepatocytes in rat liver. Carcinogenesis, 1994;145:409–22.
1987;8:1737–40.  89. Preisegger, K.H., Factor, V.M., Fuchsbichler, A. et  al. Atypical ductular
62. Evarts, R.P., Nagy, P., Nakatsukasa, H. et  al. In vivo differentiation of rat proliferation and its inhibition by transforming growth factor beta1 in the
liver oval cells into hepatocytes. Cancer Res, 1989;49:1541–7. 3,5‐diethoxycarbonyl‐1,4‐dihydrocollidine mouse model for chronic alco-
63. Schofield, R. The relationship between the spleen colony‐forming cell and holic liver disease. Lab Invest, 1999;79:103–9.
the haemopoietic stem cell. Blood Cells, 1978;4:7–25.   90. Yanger, K., Knigin, D., Zong, Y. et al. Adult hepatocytes are generated by
64. Trentin, J.J. Determination of bone marrow stem cell differentiation by stro- self‐duplication rather than stem cell differentiation. Cell Stem Cell,
mal hemopoietic inductive microenvironments (HIM). Am J Pathol, 2014;15:340–9.
1971;65:621–8.   91. Wright, N., Samuelson, L., Walkup, M.H. et al. Enrichment of a bipotent
65. Ohlstein, B., Kai, T., Decotto, E. et al. The stem cell niche: theme and varia- hepatic progenitor cell from naive adult liver tissue. Biochem Biophys Res
tions. Curr Opin Cell Biol, 2004;16:693–9. Commun, 2008;366:367–72.
66. Lapidot, T., Dar, A., and Kollet, O. How do stem cells find their way home?   92. Petersen, B.E., Goff, J.P., Greenberger, J.S. et al. Hepatic oval cells express the
Blood, 2005;106:1901–10. hematopoietic stem cell marker Thy‐1 in the rat. Hepatology, 1998;27:433–45.
67. Christiano, A.M. Epithelial stem cells: stepping out of their niche. Cell,   93. Nierhoff, D., Ogawa, A., Oertel, M. et al. Purification and characterization
2004;118:530–2. of mouse fetal liver epithelial cells with high in vivo repopulation capacity.
68. Williams, E.D., Lowes, A.P., Williams, D. et al. A stem cell niche theory of Hepatology, 2005;42:130–9.
intestinal crypt maintenance based on a study of somatic mutation in colonic  94. Suzuki, A., Zheng, Y., Kondo, R. et  al. Flow‐cytometric separation and
mucosa. Am J Pathol, 1992;141:773–6. enrichment of hepatic progenitor cells in the developing mouse liver.
69. Tumbar, T., Guasch, G., Greco, V. et  al. Defining the epithelial stem cell Hepatology, 2000;32:1230–9.
niche in skin. Science, 2004;303:359–63.   95. Dezso, K., Jelnes, P., Laszlo, V. et al. Thy‐1 is expressed in hepatic myofi-
70. Parati, E.A., Pozzi, S., Ottolina, A. et al. Neural stem cells: an overview. J broblasts and not oval cells in stem cell‐mediated liver regeneration. Am J
Endocrinol Invest, 2004;27:64–7. Pathol, 2007;171:1529–37.
71. Hering, E. Ueber den Bau der Wirbeltierleber. Sitzungsberichte der  96. Tanimizu, N., Nishikawa, M., Saito, H. et  al. Isolation of hepatoblasts
Kaiserlichen Akademie der Wissenschaften. Mathematisch‐ based on the expression of Dlk/Pref‐1. J Cell Sci, 2003;116:1775–86.
Naturwissenschaftliche Klasse 1866;Bd. 54:335–341.   97. Faris, R.A. and Hixson, D.C. Selective proliferation of chemically altered
72. Steiner, J.W. and Carruthers, J.S. Studies on the fine structure of the terminal rat liver epithelial cells following hepatic transplantation. Transplantation,
branches of the biliary tree: II. Observations of pathologically altered bile 1989;48:87–92.
canaliculi. Am J Pathol, 1961;39:41–63.   98. Dabeva, M.D., Hwang, S.G., Vasa, S.R. et al. Differentiation of pancreatic
73. Theise, N.D., Saxena, R., Portmann, B.C. et al. The canals of Hering and epithelial progenitor cells into hepatocytes following transplantation into
hepatic stem cells in humans. Hepatology, 1999;30:1425–33. rat liver. Proc Natl Acad Sci USA, 1997;94:7356–61.
74. Saxena, R., Theise, N.D., and Crawford, J.M. Microanatomy of the human  99. Dabeva, M.D., Hurston, E., and Shafritz, D.A. Transcription factor and
liver‐exploring the hidden interfaces. Hepatology, 1999;30:1339–46. liver‐specific mRNA expression in facultative epithelial progenitor cells of
75. Yovchev, M.I., Grozdanov, P.N., Joseph, B. et al. Novel hepatic progenitor liver and pancreas. Am J Pathol, 1995;147:1633–48.
cell surface markers in the adult rat liver. Hepatology, 2007;45:139–49. 100. Wang, X., Al‐Dhalimy, M., Lagasse, E. et al. Liver repopulation and cor-
76. Farber, E. Similarities in the sequence of early histological changes induced rection of metabolic liver disease by transplanted adult mouse pancreatic
in the liver of the rat by ethionine, 2‐acetylamino‐fluorene, and 3’‐methyl‐4‐ cells. Am J Pathol, 2001;158:571–9.
dimethylaminoazobenzene. Cancer Res, 1956;16:142–8. 101. Wang, X., Foster, M., Al‐Dhalimy, M. et al. The origin and liver repopulating
77. Lemire, J.M., Shiojiri, N., and Fausto, N. Oval cell proliferation and the ori- capacity of murine oval cells. Proc Natl Acad Sci USA, 2003;100(1):11881–8.
gin of small hepatocytes in liver injury induced by D‐galactosamine. Am J 102. Song, S., Witek, R.P., Lu, Y. et al. Ex vivo transduced liver progenitor cells
Pathol, 1991;139:535–52. as a platform for gene therapy in mice. Hepatology, 2004;40:918–24.
78. Dabeva, M.D. and Shafritz, D.A. Activation, proliferation, and differentia- 103. Roskams, T., van den Oord, J.J., De Vos, R. et al. Neuroendocrine features
tion of progenitor cells into hepatocytes in the D‐galactosamine model of of reactive bile ductules in cholestatic liver disease. Am J Pathol,
liver regeneration. Am J Pathol, 1993;143:1606–20. 1990;137:1019–25.
564 THE LIVER:  REFERENCES

104. Demetris, A.J., Seaberg, E.C., Wennerberg, A. et al. Ductular reaction after 131. Terada, N., Hamazaki, T., Oka, M. et al. Bone marrow cells adopt the phe-
submassive necrosis in humans. Special emphasis on analysis of ductular notype of other cells by spontaneous cell fusion. Nature, 2002;416:542–5.
hepatocytes. Am J Pathol, 1996;149:439–48. 132. Ying, Q.L., Nichols, J., Evans, E.P. et al. Changing potency by spontaneous
105. Roskams, T., De Vos, R., Van Eyken, P. et al. Hepatic OV‐6 expression in fusion. Nature, 2002;416:545–8.
human liver disease and rat experiments: evidence for hepatic progenitor 133. Wang, X., Willenbring, H., Akkari, Y. et  al. Cell fusion is the principal
cells in man. J Hepatol, 1998;29:455–63. source of bone‐marrow‐derived hepatocytes. Nature, 2003;422:897–901.
106. Roskams, T.A., Theise, N.D., Balabaud, C. et al. Nomenclature of the finer 134. Vassilopoulos, G., Wang, P.R., and Russell, D.W. Transplanted bone mar-
branches of the biliary tree: canals, ductules, and ductular reactions in row regenerates liver by cell fusion. Nature, 2003;422:901–4.
human livers. Hepatology, 2004;39:1739–45. 135. Alvarez‐Dolado, M., Pardal, R., Garcia‐Verdugo, J.M. et  al. Fusion of
107. Zhou, H., Rogler, L.E., Teperman, L. et al. Identification of hepatocytic and bone‐marrow‐derived cells with Purkinje neurons, cardiomyocytes and
bile ductular cell lineages and candidate stem cells in bipolar ductular reac- hepatocytes. Nature, 2003;425:968–73.
tions in cirrhotic human liver. Hepatology, 2007;45:716–24. 136. Weimann, J.M., Johansson, C.B., Trejo, A. et al. Stable reprogrammed het-
108. Haruna, Y., Saito, K., Spaulding, S. et al. Identification of bipotential pro- erokaryons form spontaneously in Purkinje neurons after bone marrow
genitor cells in human liver development. Hepatology, 1996;23:476–81. transplant. Nat Cell Biol, 2003;5:959–66.
109. Malhi, H., Irani, A.N., Gagandeep, S. et al. Isolation of human progenitor 137. Camargo, F.D., Finegold, M., and Goodell, M.A. Hematopoietic myelo-
liver epithelial cells with extensive replication capacity and differentiation monocytic cells are the major source of hepatocyte fusion partners. J Clin
into mature hepatocytes. J Cell Sci, 2002;115:2679–88. Invest, 2004;113:1266–70.
110. Lazaro, C.A., Croager, E.J., Mitchell, C. et al. Establishment, characteriza- 138. Willenbring, H., Bailey, A.S., Foster, M. et al. Myelomonocytic cells are
tion, and long‐term maintenance of cultures of human fetal hepatocytes. sufficient for therapeutic cell fusion in liver. Nat Med, 2004;10:744–8.
Hepatology, 2003;38:1095–106. 139. Camargo, F.D., Green, R., Capetanaki, Y. et al. Single hematopoietic stem
111. Mahieu‐Caputo, D., Allain, J.E., Branger, J. et al. Repopulation of athymic cells generate skeletal muscle through myeloid intermediates. Nat Med,
mouse liver by cryopreserved early human fetal hepatoblasts. Hum Gene 2003;9:1520–7.
Ther, 2004;15:1219–28. 140. Newsome, P.N., Johannessen, I., Boyle, S. et al. Human cord blood‐derived
112. Barker, N., van Es, J.H., Kuipers, J. et  al. Identification of stem cells in cells can differentiate into hepatocytes in the mouse liver with no evidence
small intestine and colon by marker gene Lgr5. Nature, 2007;449:1003–7. of cellular fusion. Gastroenterology, 2003;124:1891–900.
113. Sato, T., Vries, R.G., Snippert, H.J. et al. Single Lgr5 stem cells build crypt‐ 141. Harris, R.G., Herzog, E.L., Bruscia, E.M. et al. Lack of a fusion require-
villus structures in vitro without a mesenchymal niche. Nature, ment for development of bone marrow‐derived epithelia. Science,
2009;459:262–5. 2004;305:90–3.
114. Schaub, J.R., Malato, Y., Gormond, C. et al. Evidence against a stem cell 142. Jang, Y.Y., Collector, M.I., Baylin, S.B. et al. Hematopoietic stem cells con-
origin of new hepatocytes in a common mouse model of chronic liver vert into liver cells within days without fusion. Nat Cell Biol, 2004;
injury. Cell Rep, 2014;8:933–9. 6:532–9.
115. Tarlow, B.D., Pelz, C., Naugler, W.E. et al. Bipotential adult liver progeni- 143. Danet, G.H., Luongo, J.L., Butler, G. et al. C1qRp defines a new human
tors are derived from chronically injured mature hepatocytes. Cell Stem stem cell population with hematopoietic and hepatic potential. Proc Natl
Cell, 2014;15:605–18. Acad Sci USA, 2002;99:10441–5.
116. Grompe, M. Liver stem cells, where art thou? Cell Stem Cell, 144. Wang, X., Ge, S., McNamara, G. et al. Albumin‐expressing hepatocyte‐like
2014;15:257–8. cells develop in the livers of immune‐deficient mice that received trans-
117. Schaub, J.R., Huppert, K.A., Kurial, S.N.T. et al. De novo formation of the plants of highly purified human hematopoietic stem cells. Blood,
biliary system by TGFbeta‐mediated hepatocyte transdifferentiation. 2003;101:4201–8.
Nature, 2018;557:247–51. 145. Kakinuma, S., Tanaka, Y., Chinzei, R. et al. Human umbilical cord blood as
118. Michalopoulos, G.K., Barua, L., and Bowen, W.C. Transdifferentiation of a source of transplantable hepatic progenitor cells. Stem Cells, 2003;
rat hepatocytes into biliary cells after bile duct ligation and toxic biliary 21:217–27.
injury. Hepatology, 2005;41:535–44. 146. Kollet, O., Shivtiel, S., Chen, Y.Q. et  al. HGF, SDF‐1, and MMP‐9 are
119. Yovchev, M.I., Locker, J., and Oertel, M. Biliary fibrosis drives liver repop- involved in stress‐induced human CD34+ stem cell recruitment to the liver.
ulation and phenotype transition of transplanted hepatocytes. J Hepatol, J Clin Invest, 2003;112:160–9.
2016;64:1348–57. 147. Jiang, Y., Jahagirdar, B.N., Reinhardt, R.L. et al. Pluripotency of mesen-
120. Katsuda, T., Kawamata, M., Hagiwara, K. et al. Conversion of terminally chymal stem cells derived from adult marrow. Nature, 2002;418:41–9.
committed hepatocytes to culturable bipotent progenitor cells with regen- 148. Schwartz, R.E., Reyes, M., Koodie, L. et al. Multipotent adult progenitor
erative capacity. Cell Stem Cell, 2017;20:41–55. cells from bone marrow differentiate into functional hepatocyte‐like cells. J
121. Raven, A., Lu, W.Y., Man, T.Y. et al. Cholangiocytes act as facultative liver Clin Invest, 2002;109:1291–302.
stem cells during impaired hepatocyte regeneration. Nature, 2017; 149. Lee, O.K., Kuo, T.K., Chen, W.M. et al. Isolation of multipotent mesenchy-
547:350–4. mal stem cells from umbilical cord blood. Blood, 2004;103:1669–75.
122. Deng, X., Zhang, X., Li, W. et al. Chronic liver injury induces conversion 150. Anjos‐Afonso, F., Siapati, E.K., and Bonnet, D. In vivo contribution of
of biliary epithelial cells into hepatocytes. Cell Stem Cell, 2018;23: murine mesenchymal stem cells into multiple cell‐types under minimal
114–22. damage conditions. J Cell Sci, 2004;117:5655–64.
123. Wang, B., Zhao, L., and Fish, M. et al. Self‐renewing diploid Axin2(+) cells 151. Kogler, G., Sensken, S., Airey, J.A. et al. A new human somatic stem cell
fuel homeostatic renewal of the liver. Nature, 2015;524:180–5. from placental cord blood with intrinsic pluripotent differentiation poten-
124. Lin, S., Nascimento, E.M., Gajera, C.R. et  al. Distributed hepatocytes tial. J Exp Med, 2004;200:123–35.
expressing telomerase repopulate the liver in homeostasis and injury. 152. Lee, K.D., Kuo, T.K., Whang‐Peng, J. et al. In vitro hepatic differentiation
Nature, 2018;556:244–8. of human mesenchymal stem cells. Hepatology, 2004;40:1275–84.
125. Li, B., Dorrell, C., Canaday, P.S. et al. Adult mouse liver contains two dis- 153. Aurich, I., Mueller, L.P., Aurich, H. et al. Functional integration of hepato-
tinct populations of cholangiocytes. Stem Cell Reports, 2017;9:478–89. cytes derived from human mesenchymal stem cells into mouse livers. Gut,
126. Goodell, M.A. Stem‐cell “plasticity”: befuddled by the muddle. Curr Opin 2007;56:405–15.
Hematol, 2003;10:208–13. 154. Sato, Y., Araki, H., Kato, J. et al. Human mesenchymal stem cells xeno-
127. Wagers, A.J. and Weissman, I.L. Plasticity of adult stem cells. Cell, grafted directly to rat liver are differentiated into human hepatocytes with-
2004;116:639–48. out fusion. Blood, 2005;106:756–63.
128. Thorgeirsson, S.S. and Grisham, J.W. Hematopoietic cells as hepatocyte 155. Banas, A., Teratani, T., Yamamoto, Y. et al. Adipose tissue‐derived mesen-
stem cells: a critical review of the evidence. Hepatology, 2006;43:2–8. chymal stem cells as a source of human hepatocytes. Hepatology,
129. Petersen, B.E., Bowen, W.C., Patrene, K.D. et al. Bone marrow as a poten- 2007;46:219–28.
tial source of hepatic oval cells. Science, 1999;284:1168–70. 156. Sgodda, M., Aurich, H., Kleist, S. et al. Hepatocyte differentiation of mes-
130. Menthena, A., Deb, N., Oertel, M. et al. Bone marrow progenitors are not the enchymal stem cells from rat peritoneal adipose tissue in vitro and in vivo.
source of expanding oval cells in injured liver. Stem Cells, 2004;22:1049–61. Exp Cell Res, 2007;313:2875–86.
44:  Liver Repopulation by Cell Transplantation and the Role of Stem Cells in Liver Biology 565

157. Takahashi, K., Tanabe, K., Ohnuki, M. et al. Induction of pluripotent stem 170. Huang, P., Zhang, L., Gao, Y. et  al. Direct reprogramming of human
cells from adult human fibroblasts by defined factors. Cell, fibroblasts to functional and expandable hepatocytes. Cell Stem Cell,
­
2007;131:861–72. 2014;14:370–84.
158. Yu, J., Vodyanik, M.A., Smuga‐Otto, K. et al. Induced pluripotent stem cell 171. Sekiya, S. and Suzuki, A. Direct conversion of mouse fibroblasts to
lines derived from human somatic cells. Science, 2007;318:1917–20. ­hepatocyte‐like cells by defined factors. Nature, 2011;475:390–3.
159. Hamazaki, T., Iiboshi, Y., Oka, M. et al. Hepatic maturation in differentiat- 172. Zhu, S., Rezvani, M., Harbell, J. et  al. Mouse liver repopulation with
ing embryonic stem cells in vitro. FEBS Lett, 2001;497:15–9. ­hepatocytes generated from human fibroblasts. Nature, 2014;508:93–7.
160. Jones, E.A., Tosh, D., Wilson, D.I. et al. Hepatic differentiation of murine 173. Vierbuchen, T., Ostermeier, A., Pang, Z.P. et  al. Direct conversion of
embryonic stem cells. Exp Cell Res, 2002;272:15–22. ­fibroblasts to functional neurons by defined factors. Nature, 2010;463:
161. Yamada, T., Yoshikawa, M., Kanda, S. et  al. In vitro differentiation of 1035–41.
embryonic stem cells into hepatocyte‐like cells identified by cellular uptake 174. Roy‐Chowdhury, N., Wang, X., Guha, C. et  al. Hepatocyte‐like cells
of indocyanine green. Stem Cells, 2002;20:146–54. derived from induced pluripotent stem cells. Hepatol Int, 2017;11:54–69.
162. Yamamoto, H., Quinn, G., Asari, A. et  al. Differentiation of embryonic 175. Sakaida, I., Terai, S., Yamamoto, N. et al. Transplantion of bone marrow
stem cells into hepatocytes: biological functions and therapeutic applica- cells reduces CCl4‐induced liver fibrosis in mice. Hepatology,
tion. Hepatology, 2003;37:983–93. 2004;40:1304–11.
163. Rambhatla, L., Chiu, C.P., Kundu, P. et al. Generation of hepatocyte‐like 176. Ueno, T., Nakamura, T., Torimura, T. et  al. Angiogenic cell therapy for
cells from human embryonic stem cells. Cell Transplant, 2003;12:1–11. hepatic fibrosis. Med Mol Morphol, 2006;39:16–21.
164. Kubo, A., Shinozaki, K., Shannon, J.M. et al. Development of definitive endo- 177. Higashiyama, R., Inagaki, Y., Hong, Y.Y. et al. Bone marrow‐derived cells
derm from embryonic stem cells in culture. Development, 2004;131:1651–62. express matrix metalloproteinases and contribute to regression of liver
165. Gouon‐Evans, V., Boussemart, L., Gadue, P. et al. BMP‐4 is required for fibrosis in mice. Hepatology, 2007;45:213–22.
hepatic specification of mouse embryonic stem cell‐derived definitive 178. Taniguchi, E., Kin, M., Torimura, T. et  al. Endothelial progenitor cell
endoderm. Nat Biotechnol, 2006;24:1402–11. transplantation improves the survival following liver injury in mice.
­
166. Heo, J., Factor, V.M., Uren, T. et  al. Hepatic precursors derived from Gastroenterology, 2006;130:521–31.
murine embryonic stem cells contribute to regeneration of injured liver. 179. Yoshida, Y., Tokusashi, Y., Lee, G.H. et al. Intrahepatic transplantation of
Hepatology, 2006;44:1478–86. normal hepatocytes prevents Wilson’s disease in Long‐Evans cinnamon
167. Basma, H., Soto‐Gutierrez, A., Yannam, G.R. et  al. Differentiation and rats. Gastroenterology, 1996;111:1654–60.
transplantation of human embryonic stem cell‐derived hepatocytes. 180. Peterson, E.A., Polgar, Z., Devakanmalai, G.S. et al. Genes and pathways
Gastroenterology, 2009;136:990–9. promoting long‐term liver repopulation by ex vivo hYAP‐ERT2 transduced
168. Suzuki, A. Artificial induction and disease‐related conversion of the hepatic hepatocytes and treatment of jaundice in Gunn rats. Hepatol Commun,
fate. Curr Opin Genet Dev, 2013;23:579–84. 2018;3:129–46.
169. Du, Y., Wang, J., Jia, J. et al. Human hepatocytes with drug metabolic func-
tion induced from fibroblasts by lineage reprogramming. Cell Stem Cell,
2014;14:394–403.
Liver Regeneration
45 George K. Michalopoulos
Department of Pathology, University of Pittsburgh School of Medicine, Pittsburgh, PA, USA

INTRODUCTION in mass if a big portion of the organ (one kidney, one lung, part
of pancreas) is lost, but the residual organ tissue will not attain
Liver is the largest organ of the body. It functions as a portal of the exact total weight that existed prior to the tissue loss. This is
entry and metabolic processing for all substances absorbed not the case for liver! If a major portion of liver tissue is acutely
through the gastrointestinal tract. It is the major biochemical lost, liver will enter into a regenerative process so that the total
transducer for body homeostasis, triaging the absorbed sub- hepatostat‐driven LWBR is precisely reestablished in the status
stances from the gut for storage, elimination, transformation quo ante. This is the process of liver regeneration (LR) that will
into other types of organic chemicals, or packaged transport to be majorly examined in this chapter.
the bloodstream for whole body utilization. To the extent that In a clinical disease setting, regeneration is manifested in con-
muscles work mostly on fatty acids and brain primarily on glu- ditions leading to severe loss of hepatocytes. Chronic loss of
cose, liver controls the availability of both the nutrients. The key hepatocytes is seen in infectious diseases (e.g. HCV and HBV
role of liver as the “provider” for metabolism of all body tissues viruses), chronic toxic conditions (e.g. alcohol, metabolic dis-
is perhaps the reason for an established tight relationship eases, non‐alcoholic steatohepatitis (NASH), storage diseases,
between liver weight and body weight. Changes in body status hemochromatosis, alpha‐1 antitrypsin deficiency), ischemia‐
(cachexia, puberty, pregnancy, chronic disease, etc.) typically reperfusion injury (most often following liver transplantation),
alter the algorithm of the “liver to body weight ratio” (LBWR). or chronic immune attacks (autoimmune hepatitis). Chronic loss
Whatever the operative algorithm of LBWR may be for a given of hepatocytes is accompanied by compensatory proliferation of
body at any given time, the LBWR is tightly maintained, by the surviving hepatocytes. As we will address below in this chap-
mechanisms not fully understood. The rates of proliferation of ter, this occurs in potentially genotoxic environments and the
hepatic cells under normal conditions are typically very low (for chronic compensatory proliferation of the surviving hepatocytes
hepatocytes, less than 0.2%), but they exceed zero. Recent stud- may lead to development of neoplasia. Acute loss of hepatocytes
ies have demonstrated that under the placid calmness of the is rarer, often caused by acute ingestion of toxins (e.g. attempted
hepatic capsule, there are slow proliferative events of cells from suicide by acetaminophen), trauma, or a short course of an acute
different zones of the lobule, aiming to keep liver cell numbers hepatitis. In all these conditions, death of hepatocytes and com-
and organ weight steady. The precise mechanisms controlling pensatory proliferation proceed in tandem with inflammatory
these processes are not fully understood, but for operative processes which aim to remove dead hepatocytes and provide
reasons, we need to accept the existence of a set of controls cytokines for tissue repair. Overall, the regenerative processes in
comprising a “hepatostat”, a set of processes that ascertain acute loss of hepatocytes operate very efficiently. Studies of
maintenance of liver weight to where it needs to be [1]. Such patients with Dubin–Johnson (pigment similar to melanin accu-
demands for maintenance of steady weight exist in endocrine mulating in normal hepatocytes) have shown that approximately
glands controlled by the pituitary, and the feedbacks controlling 90% of hepatocytes die in an acute hepatitis, but they are restored
those homeostatic events are better understood. Other organs in proper numbers by the end of the disease [2]. It is highly likely
(pancreas, kidneys, intestine, lungs, etc.) will increase slightly that in a clinical setting, the inflammatory processes and the

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
45:  Liver Regeneration 567

regenerative events cooperate synergistically. It is difficult, Table 45.1  Mitogenic signals associated with liver regeneration after
however, to design experiments that would cleanly dissect the partial hepatectomy
regulatory signals regulating only the compensatory prolifera- Complete mitogens
tion of hepatocytes, from the comingled inflammatory events. 1. Mitogenic in hepatocyte cultures in chemically defined (serum‐free)
media
Experimental models in rodents associated with acute adminis- 2. Cause liver enlargement and hepatocyte DNA synthesis when injected
tration of a liver toxin (typically CCl4 or acetaminophen) are into whole animals:
more relevant to the ­clinical setting. Hepatocyte Growth Factor (HGF) and receptor (c‐Met)
The study of signals exclusively associated with regulation of Ligands of the EGF R (EGF, TGFα, HB‐EGF, amphiregulin)
Combined elimination of these two signaling receptors abolishes liver
hepatic regeneration, not affected by inflammatory processes, regeneration
has been carried mostly in rodent livers. Each rodent liver is
Auxiliary mitogens
composed of several lobes, and each lobe is supported by its 1. Ablation of their signaling pathways causes delay but does not abolish
own vasculature (arterial and portal vein) and separate branch of liver regeneration.
the common bile duct. An easy surgical procedure, first 2. They are not mitogenic in hepatocyte cultures and when injected in
vivo do not cause hepatocyte DNA synthesis and liver enlargement
described by Higgins and Anderson in 1933 [3], allows clean Norepinephrine and the α1 adrenergic receptor
removal of the two larger lobes, comprising around two‐thirds TNF and TNFR1
of the liver. The non‐resected small lobes remain intact, with no IL6
damage or associated inflammatory response. In rats, this pro- Notch and Jagged (recombinant Jagged causes DNA synthesis in
hepatocyte cultures)
cedure allows removal of about 68% of the liver mass, whereas VEGF and receptors I and II
in mice, due to the presence of a gallbladder, typically about Bile acids
60% of liver mass can be easily removed. The procedure is Serotonin
Complement proteins
known as “2/3 partial hepatectomy” (PHx) and the results Leptin
obtained by using this procedure will be the basis for discussion Insulin
of regulatory signals associated with LR. At the end of the PPAR gamma
regenerative process, the non‐resected lobes become larger and, Very important and yet to be defined
in aggregate, restore the hepatic mass that was present prior to Elimination of these signaling molecules does not abolish liver
regeneration. Their effects by exogenous administration to intact animals
PHx. The regenerated liver has a different shape; it now has or hepatocyte cultures have not been fully tested
fewer lobes (two or three, based on anatomic conventions), and Hedgehog
the resected lobes are not restored. Wnt/beta‐catenin
Signals controlling Hippo pathway and Yap

SIGNALS REGULATING LR Hepatocyte growth factor (HGF) and its


Signals related to LR have been identified by triggering prolifera-
receptor (c‐Met)
tion in hepatocyte cultures, delay of regeneration associated with HGF was first isolated on the basis of its capacity to induce
signal ablation in genomically modified mice, or response of DNA synthesis in primary cultures of hepatocytes [7].
intact liver in vivo by injecting specific signaling substances [4, Subsequent cloning and sequencing revealed an unusual struc-
5]. Many of these signals appear not only in the regenerating liver, ture, composed of four Kringle domains and a pseudo‐protease
but also in the peripheral blood. Table  45.1 shows a list of the domain [8, 9]. There is no active protease function in HGF how-
most studied mitogenic signals associated with LR. Of these sig- ever, despite its otherwise strong homology to plasminogen. It is
nals, epidermal growth factor receptor (EGFR) and associated synthesized as a single polypeptide, cleaved to its active heter-
ligands (epidermal growth factor [EGF], transforming growth odimeric form in an Arg‐Val‐Val site, distal to the Kringle
factor [TGFα], heparin binding epidermal growth factor domains (identical to plasminogen). Urokinase plasminogen
[HBEGF], amphiregulin) and hepatocyte growth factor (HGF) activator (uPA), free or bound to its receptor (uPAR) directly
and its receptor (MET) are capable of inducing proliferation of activates HGF; tissue plasminogen activator (TPA) can also
hepatocytes in serum‐free cultures [6]. They also induce hepato- activate HGF, but not as effectively [10]. HGF can also be acti-
cyte DNA synthesis when injected into the portal vein of normal vated by a soluble protein with high homology to Factor XII,
rodent livers. They should be viewed as “direct mitogens”. Others known as HGF activator (HGFA). Hepsin and matriptase have
show their relevance by delays in LR following elimination of the also been shown to activate HGF. In normal liver, HGF is pro-
signaling molecule or its cognate receptor. They are not directly duced by hepatic stellate cells and it is deposited in large
mitogenic for hepatocytes in culture or in vivo. They should be amounts in the extracellular matrix (ECM), mostly in the peri-
viewed as “indirect mitogens”. Some more recently discovered portal region of the lobule, bound to glycosaminoglycans and
signals (e.g. Wnt and Hedgehog) have not been tested in vivo or specific collagens [11]. HGF production by stellate cells is con-
in vitro but they do have strong effects on LR and homeostasis. In trolled by the neurotrophin receptor p75NTR [12]. At later
this chapter, we will discuss these signals with emphasis on their stages of LR, it is also produced by sinusoidal endothelial cells
effects on LR. It should be noted that LR always finds pathways and bone marrow endothelial progenitor cells homing to the
to bypass elimination of any single mitogenic signal, direct or liver [13, 14]. HGF is produced in most peripheral tissues by
indirect, and proceeds to completion by following alternate path- mesenchymal cells; it is also produced by specific populations
ways overcoming the signaling obstacle. of neurons and glial cells in both brain (hippocampus, frontal
568 THE LIVER:  SIGNALS REGULATING LR

and parietal cortex) and spinal cord [15] and also in the target of Yap protein, is also minimally expressed in normal
­placental syncytiotrophoblast [16]. The HGF receptor, MET, is liver, but its expression in hepatocytes rapidly increases after
the product of the gene c‐Met. It is also synthesized as a long PHx. Germline elimination of amphiregulin delays liver genera-
polypeptide and activated into its dimeric form at the Golgi tion [24]. Heparin binding EGF (HB‐EGF) is produced by mac-
apparatus. It resides as a transmembrane protein in the plasma rophages and endothelial cells but not in hepatocytes [25]. It is
membrane of most epithelial cells, selective groups of neurons also produced as a transmembrane protein, and the extracellular
and glial cells, as well macrophages and endothelial cells. domain is released by the action of metalloproteinases as the
Both hepatocytes and cholangiocytes express MET. Activation mature form of HB‐EGF. It rises rapidly after PHx, prior to the
of MET is typically the result of ligation with active HGF. This appearance of TGFα. Germline elimination of HB‐EGF delays
results in formation of HGF–MET dimers, based on the dimeri- LR and the progression of the steps through the cell cycle [26].
zation domain present on HGF. The formation of the HGF– In the normal liver, EGF is the major interacting ligand with
MET dimers is associated with cross‐phosphorylation of EGFR. TGFα, amphiregulin, and HB‐EGF are rapidly mobi-
tyrosine amino acids at sites 1234 and 1235. This is followed lized after PHx and they have complementary functions, with
by  phosphorylation of tyrosine in position 1349, which then neither one individually being critical for regeneration, but hav-
becomes a docking site for attachment of other proteins (includ- ing effects related both to hepatocyte proliferation as well as to
ing GAB1, GRB2), responsible for a multiplicity of downstream proliferation of cholangiocytes and endothelial cells, all of
signals, which eventually cause activation of PIPK3, AKT (pro- which abundantly express EGFR.
tein kinase B), and mammalian target of rapamycin (mTOR),
rearrangement of cell polarity and enhancement of motility and/
or stimulation of cell proliferation [17]. The “signalosome” for- Fibroblast growth factors (FGF)
mation around Gab1 is already assembled within 30 minutes
and their receptors
after PHx [18]. Hepatocyte MET appears always activated in the
liver, presumably due to availability of HGF in abundance in the FGF is a large family of signaling proteins, composed of 23 mem-
hepatic ECM [19]. Incomplete forms of HGF containing only bers with diverse structures and cells or tissues of origin. Their
one or two Kringles (NK1, NK2) can also ligate MET, but since common feature is their capacity to activate some or all of four
they do not contain the dimerization domain of complete HGF, receptors (FGFR1–FGR4). Mature hepatocytes express only
dimers can only form and MET activation occurs only in the FGFR4 [27], with the other receptors present in non‐epithelial
presence of glycosaminoglycans. hepatic cells. FGF1/2 are expressed by hepatocytes during LR
[5]. They are slightly mitogenic in hepatocyte cultures (less than
20% of the effect of HGF or EGF). It is not clear whether, during
Epidermal growth factor receptor (EGFR) LR, FGF1/2 are exerting autocrine or paracrine effects, especially
on endothelial cells expressing FGFR1/2. In hepatocytes, FGFR4
and associated ligands is co‐expressed and functioning in conjunction with βKlotho.
EGFR is expressed in both hepatocytes and cholangiocytes. It is Elimination of either βKlotho or FGFR4 s had no effect on hepat-
a member of the ErbB family of receptors. ERB1 (EGFR) and ocyte proliferation but it disrupted hepatic cholesterol, bile acid,
ERB3 are expressed in adult hepatocytes and cholangiocytes, and lipid metabolism [28]. In another study, however, knockdown
whereas fetal hepatocytes express the above as well as ErbB2 of FGFR4 was associated with increased necrosis of non‐paren-
(Her2/Neu) [20]. ERB3 does not have a kinase domain and chymal cells and 25% mortality after PHx; liver weight was
ErbB4 is not expressed in liver. In contrast to HGF/MET, EGFR restored by hepatocyte hypertrophy [29]. The role of FGFR4 is
and ERB3 can form heterodimers with themselves or with each crucial in hepatocyte biology. It is the receptor ligated by FGF15
other without being ligated. There are multiple ligands of EGFR. (human FGF19), produced in the intestine following stimulation
The ones studied in liver and in the context of LR, primarily by bile acids binding to the intestinal farnesoid X receptor (FXR)
include EGF, TGFα, amphiregulin, and HB‐EGF. EGF is pro- [30]. Functioning as an endocrine signal, FGF15 enters through
duced in secretions of exocrine glands, including salivary the portal circulation and regulates cholesterol metabolism and
glands. Brunner’s glands, with histology similar to salivary suppresses hepatocyte bile acid synthesis via FGFR4, by down-
glands, exist in the submucosa of the duodenum. They secrete regulating CYP7a1, the rate limiting enzyme for bile acid synthe-
EGF in the intestinal lumen. A portion of the secreted EGF is sis. Elimination of FGF15 caused massive increase of hepatic bile
absorbed intact from the lumen of the duodenum and trans- acids and increased mortality after PHx with delayed LR [31].
ported to the liver via the portal circulation [21]. Thus, hepato- FGF21 is another endocrine signal, often viewed as a “hepa-
cytes are constantly exposed to EGF, and EGFR is found tokine”, produced by hepatocytes. Its synthesis is controlled by
activated in normal liver [19]. EGF secretion from Brunner’s PPAR’α and PGC‐1a. It is also produced in testes, pancreas, and
glands is enhanced by norepinephrine [22]. TGFα is minimally adipose tissue. It has insulin‐mimetic effects and exerts general
expressed in resting normal liver but its expression is maximally metabolic effects in a variety of tissues, including adipose tissue
enhanced following PHx [23]. It is produced as a transmem- [32]. Stellate cells both produce and respond to FGF; evidence
brane protein and the mature form is the extracellular domain, suggests that FGF regulate synthesis of ECM components by stel-
released by proteolytic action by TACE/ADAM17 protease. late cells [33]. All evidence suggests that FGF and their receptors
Germline elimination of TGFα does not affect liver generation, are crucial for regulation of hepatic function, proliferation of non‐
presumably due to complementary effects by the other EGFR parenchymal cells, and regulation of key metabolic functions in
ligands involved in the process. Amphiregulin, a downstream normal and regenerating liver.
45:  Liver Regeneration 569

Transforming growth factor beta (TGFβ) Interleukin 6 (IL6)


There are three different forms of TGFβ, all present in the liver. IL6 has been studied extensively in hepatocytes, as the major
TGFβ receptor is the same for all three forms, composed of ligand triggering the “acute phase response”, a rapid increase
three components (TGFβr I, II, and III), present in all hepatic and secretion of many proteins associated with innate immunity.
cell types. TGFβR II is the site in which TGFβ binds directly. Circulating IL6 binds to a soluble receptor, and the complex
In liver, TGFβ1 has been the most studied and it is the one to be binds to a hepatocyte receptor known as gp130, which is shared
referred to in this segment. It is produced by stellate cells and with oncostatin M, LIF, CNTF, and so on. This trimeric com-
Kupffer cell macrophages. In normal liver, TGFβ is bound to its plex dimerizes to form a hexameric complex and is subsequently
receptor and also to decorin, a GPI‐linked protein on plasma phosphorylated in Tyr residues, becoming a docking site for
membrane. Following PHx, TGFβ is gradually removed from activation of JAK tyrosine kinases and STAT transcription fac-
hepatocytes as regeneration proceeds from periportal to peri- tors. In the context of LR, IL6 is the major regulator of activa-
central areas of the lobule [34]. It is also massively increasing tion of STAT3 transcription factor. Germline elimination of IL6
in the plasma, probably released from decorin as a result of the is associated with delayed regeneration due to delayed STAT3
early remodeling of ECM following PHx (see below, section on activation. LR however proceeds and completes normally, as
Urokinase plasminogen activator (uPA) and ECM remodeling) other signals (MET, EGFR) can also phosphorylate and activate
[4]. Circulating alpha‐2 macroglobulin binds TGFβ and trans- STAT3 [43–45]. IL6 is produced by Kupffer cells, but also by
ports it to hepatocytes where it is inactivated [4, 5, 35]. TGFβ is hepatocytes [46]. It rises in plasma after PHx, following the rise
mitoinhibitory in hepatocyte cultures, and its removal from in circulating tumor necrosis factor (TNF), a major stimulus for
liver soon after PHx would be a rational expectation, in order to IL6 secretion [47].
allow hepatocytes to proliferate. Administration of exogenous
TGFβ immediately after PHx delays initiation of hepatocyte
Tumor necrosis factor alpha (TNF)
DNA synthesis. Surprisingly, however, TGFβ expression
increases at about 3 hours after PHx and reaches a plateau at 72 TNF is produced primarily by macrophages and binds to two
hours, at the same time as hepatocytes are at their maximal receptors (TNFR1, TNFR2). Direct administration of TNF to
state of proliferation [36]. The successful completion of hepat- normal animals leads to liver damage. TNF, however, rapidly
ocyte proliferation as TGFβ levels rise, is accomplished by the rises in the plasma after PHx, with no apparent damage to the
simultaneous decrease in expression of hepatocyte TGFβ liver [48]. Studies by Fausto et al. have shown that germline
receptors at the time of hepatocyte proliferation [37]. The con- elimination of either TNFR1 or TNFR2 leads to delayed LR
temporaneous rise in plasma norepinephrine likely also con- and decreased production of IL6 [47]. Perhaps the most
tributes to the hepatocyte “resistance” to TGFβ; as evidenced important function of TNF in LR is the activation of the tran-
by studies in hepatocyte cultures, norepinephrine attenuates the scription factor NFκB, which is delayed in TNFR knockout
effect of TGFβ and suppresses its mitoinhibitory activity [38]. (KO) mice [49]. TNF also regulates expression of TACE/
Studies in hepatic organoid cultures show that the increase in ADAMS17, the plasma membrane protease associated with
TGFβ expression after PHx is linked to activation of EGFR and secretion of the mature (extracellular) form of TGF [50]. It is
MET [39]. TGFβ plays an important role at the later stages of of interest that either TNFR1 or FAS activation can induce
regeneration, though not by affecting termination of liver complete liver failure, independent of each other, and yet,
regeneration (see below, section on Termination of liver regen- similar to TNFR, germ line elimination of FAS also leads to
eration). TGFβ is important in angiogenesis, inducing forma- delayed LR [51].
tion of tubules by endothelial cells [35]. It also likely stimulates
production of ECM by stellate cells, a phenomenon that occurs
Norepinephrine (NE)
at the later stages of LR, restoring ECM after its remodeling
immediately following PHx [40]. All evidence suggests that This neurotransmitter is produced at the synaptic endings and,
TGFβ expression during LR is a constructive event and impor- to a lesser extent, by adrenal medulla. Stellate cells also produce
tant for the completion of LR by restoration of intact tissue and respond to NE [52]. Addition of NE in serum‐free hepato-
histology. This is also bolstered by the finding that loss of cyte cultures dramatically enhances the mitogenic effects of
beta‐2 spectrin, a component of the TGFβ signaling pathway, is EGF and HGF [53]. In addition, it downregulates the mitoin-
associated with suppressed and delayed liver regeneration [41]. hibitory effects of TGFβ [38]. In “balanced” concentrations of
Also of interest, is a previous study in which normal rats were mitogen EGF and mito‐inhibitor TGFβ, addition of NE leads to
injected with a dominant negative construct against TGFβ hepatocyte DNA synthesis [38]. NE acts by binding to the
receptor II. This, unexpectedly, resulted in initiation of DNA G‐protein coupled alpha‐1 adrenergic receptor (A1AR). Its
synthesis in the normal liver. The results suggest that under effects on hepatocyte proliferation are mediated, at least in part,
normal conditions, hepatocytes are under opposite “tonic” by A1AR inducing phosphorylation of EGFR and activation of
influences between ambient mitogens HGF, EGF, and so on, STAT3 via Src kinase [54]. NE rises rapidly in plasma after PHx
and the mitoinhibitory effect of TGFβ, with the combined and exerts the above effects directly on hepatocytes at the time
effect resulting in maintaining hepatocytes in G0 state and in when both EGFR and MET are becoming increasingly activated
a stable state of differentiation; acute removal of TGFβ signal- [55]. NE, however, has effects on LR from extrahepatic sites.
ing creates an imbalance driving hepatocytes toward DNA NE increases availability of both EGF and HGF to the regener-
­synthesis [42]. ating liver by directly enhancing secretion of EGF from
570 THE LIVER:  SIGNALS REGULATING LR

Brunner’s duodenal glands [22] and production of HGF by mes- proteins. (For more information on this important group of
enchymal cells [56]. Given the high levels of NE in plasma after hepatocyte signaling molecules, please see Chapter 46.)
PHx, it may be the mediating factor for the observed increased
in HGF production in extrahepatic sites (lungs, kidney) after
PHx [57]. Administration of prazosin, a specific A1AR inhibi-
Hedgehog (Hh) signaling
tor, has long‐lasting (three days) effects of suppression of hepat- Emerging literature has demonstrated multiple roles of the Hh
ocyte DNA synthesis after PHx. Similar effects were seen after signaling pathway in liver pathobiology. In relation to LR, all
surgical sympathectomy of the liver prior to PHx [55]. known hepatic cell types produce Hh and respond to Hh signal-
ing. Inhibition of Hh signaling by cyclopamine delayed liver
regeneration by 48 hours. There was suppression of hepatocyte
Bile acids and cholangiocyte proliferation and decreased activation of stel-
Bile acids increase in circulating blood after PHx and depletion late cells [64]. Glypican 3 (GPC3), a GPI‐linked protein on
of bile acids leads to decreased regeneration [58]. The rise of hepatocyte plasma membrane binding to tetraspanin CD81,
bile acids in plasma, however, occurs several hours after PHx, binds and retains Hh proteins in normal liver. Following PHx,
suggesting that bile acids act after LR has already been initiated. GPC3 ceases binding to CD81 and releases Hh proteins, while
Bile acids bind to the transcription factor FXR, resulting in a Gli1, a transcription factor controlled by Hh, appears promi-
feedback suppression of bile acid synthesis by downregulation nently in hepatocyte nuclei at day 2 after PHx [65]. Hh signaling
of CYP7a1, the first enzyme involved in bile acid synthesis. regulates expression of Yap in stellate cells, and Yap itself is
Mice with genetic elimination of FXR have defective regenera- involved in stellate cell activation and production of ECM [66].
tion [59]. Despite the evidence of suppressed regeneration fol- Elimination of Smoothened, the cytoplasmic signaling mediator
lowing bile acid depletion or germline elimination of FXR, the of Hh, in stellate cells also dramatically decreased expression of
mediating pathways are not clear. It was originally thought that Yap in hepatocytes after PHx [67]. These results demonstrate
the effects of bile acids on LR were mediated by the production that much is yet to be discovered on the regulation of LR by this
of FGF15 in the intestine (mediated by bile acids binding to very complex system of important signaling regulators.
intestinal FXR). As mentioned above, FGF15 acts as an endo-
crine FGF and regulates many aspects of lipid metabolism,
including synthesis of bile acids, via receptor FGR4. Germline
Insulin
elimination of FGF15, however, has minimal effects on LR. Insulin can best be described as an enabler of all hepatocyte
FXR elimination and its effects on LR are probably mediated by functions, including proliferation and differentiation. Produced
the uncontrolled biosynthesis of bile acids observed in the FXR by the beta cells of pancreatic islets, it is supplied directly to the
KO mice, associated with cholestatic damage of hepatic tissue. liver via portal vein, before it goes to any other organ in the
In FXR KO mice, bile acid synthesis is not inhibited neither via body. In the absence of insulin, mitogenic effects of HGF and
FXR, nor by FGF15 acting on FGFR4, because synthesis of EGF in hepatocyte cultures are abolished and the viability of
FGF15 is also suppressed in FXR KO mice. There is recent evi- hepatocytes is severely curtailed [6]. Diversion of the portal cir-
dence that bile acids may play a direct role in cholangiocyte culation directly to the vena cava (portacaval shunt) causes liver
proliferation via the G‐protein coupled receptor TGR5 [60]. atrophy. Administration of insulin in animals with portacaval
shunt restores liver weight, and this process is mediated by pro-
liferation of hepatocytes [68]. Insulin, however, is not by itself
Serotonin mitogenic in hepatocyte cultures [53] and does not cause hepat-
Thrombocytopenic mice have deficient LR and this is partially ocyte proliferation when administered to normal animals.
corrected by administration of serotonin [61]. Administration
of serotonin to normal liver or hepatocytes in culture does not
have direct mitogenic effects on hepatocytes. Mice with low
Growth hormone (GH)
levels of serotonin (deficient in tyrosine hydroxylase) also have There are isolated studies on effects of GH in liver regeneration
deficient LR [61]. On the other hand, mice with germline elimi- but there is no clear evidence that GH is a major regulator. It
nation of the serotonin transporter resulting in severe decrease controls production of “somatomedins” insulin‐like growth fac-
of serotonin in platelets and peripheral blood, do not have defi- tor 1 (IGF1) and IGF2 by hepatocytes, which play important
cient LR [62]. Serotonin has multiple beneficial effects on liver roles in liver development. GH has been linked to proper timing
in clinical studies and stimulation of VEGF production may of hepatocyte DNA synthesis and to activation of EGFR [69]
play a role [63]. and enhances liver regeneration in aged animals [70].

Wnt/beta‐catenin Purinergic signals and NK cells


The family of Wnt proteins has emerged as a critical signaling There is evidence that ATP is rapidly released immediately after
pathway for tissue repair and regeneration in most tissues. They PHx due to increase in portal vein pressure, derived from hepato-
operate through the Frizzled family of receptors, with LRP5/6 cyte and macrophage lysosomes. The phenomenon is of short
acting as coreceptors. There are multiple members to the Wnt duration. The released ATP is rapidly hydrolyzed by ecto‐ATPases
family and they are expressed in different proportions from all present on plasma membrane of most cells. Adenosine, the final
hepatic cell types. Beta‐catenin is the signaling vehicle for Wnt derivative of ATP hydrolysis, can interact with purinergic
45:  Liver Regeneration 571

receptors. Blockade of P2Y2 purinergic receptor by a specific Table 45.2  Activation of signaling pathways and changes occurring
inhibitor delayed hepatocyte proliferation [71]. Other studies early after partial hepatectomy
showed that the stimulation of purinergic receptors primarily • Increase in urokinase activity (first 5 minutes)
affects hepatic natural killer (NK) cells and interference with NK • Translocation of N(otch)ICD to the nucleus (15 minutes)
• Translocation of beta‐catenin to the nucleus (5–10 minutes to 6 hours)
cell function also delays liver regeneration [72]. Hepatocytes • Decrease in HGF biomatrix stores (30 minutes to 3 hours)
express two main types and multiple subtypes of purinergic • Activation of the HGF receptor (within 30–60 minutes)
receptors, and their stimulation regulates multiple signaling path- • Activation of the EGF receptor (within 30–60 minutes)
ways associated with hepatocyte proliferation, including phos- • Increase of HGF, norepinephrine, IL6, TNF, TGFb1, serotonin, and
hyaluronic acid in the plasma within 1–2 hours.
phorylation of ERK1 [71]. • Activation of AP1, NFkB, and STAT3 within 2–3 hours.
• Extensive gene expression reprogramming of hepatocytes within 30
minutes after PHx
Immediate biochemical or physical signals • Remodeling of extracellular matrix within the first 2–3 hours
associated with initiation of liver regeneration
There are two immediate consequences of PHx: this event, since anti‐uPA antibodies inhibit activation of the
1. Immediate increase of portal flow to the remaining liver single peptide, pro‐ HGF in whole liver homogenates taken
lobes (approximately one‐third of the original liver mass). within 5 minutes after PHx [78]. uPA is also involved with the
2. Acute changes in peripheral blood and tissue constituents remodeling of ECM observed within the first 2 hours after PHx.
that inform on the required, “hepatostat” driven, liver to body The sequential steps (conversion of plasminogen to plasmin,
weight ratio. subsequent degradation of fibrinogen, and activation of metal-
loproteinases MMP2 and MMP9 by plasmin) become noted
The identity of the ultimate early signals that trigger LR must be shortly after PHx, and are eventually downregulated by a subse-
sought in these two fundamental and rapid changes. The early quent elevation of TIMP1 by 6–18 hours [5, 35, 79]. There are
observation by Moolten and Bucher that PHx in one member of changes in multiple proteins of ECM (fibronectin prominently
a pair of parabiotic rats stimulated liver regeneration in the decreasing in periportal area within 5 minutes after PHx) fol-
unoperated member of the pair [73], gave rise to all subsequent lowed by an increase in mRNA of several of these proteins,
studies that resulted in isolation of HGF and documentation of probably aiming to resynthesize them as a compensatory mech-
PHx‐related changes in plasma concentrations of IL6, TNF, NE, anism [80]. Hyaluronic acid, also an ECM component, is rap-
bile acids, and so on. idly released in the peripheral blood and TGFβ1, present in the
In addition to the effects of specific molecular/biochemical extracellular space bound to hepatocyte decorin, is also rapidly
signals, the very physical forces generated by the increased por- released with the same kinetics [5, 35]. HGF, embedded in the
tal flow through the remnant liver may also play a role. extracellular matrix as pro‐HGF form, is released to the plasma
Preventing rise in portal vein pressure after PHx results in as active HGF and rises very rapidly to more than tenfold
reduced early events of LR and deficient activation of HGF [74]. increase within 1 hour after PHx [81]. New HGF synthesis and
Fluid shear stress due to acute increase in flow over endothelial the appearance of new HGF mRNA starts at 3 hours after PHx
monolayers is associated with production of specific signals. and reaches plateau at 24 hours. Increased expression of HGF
Stability of urokinase plasminogen activator (uPA) mRNA and mRNA is also seen in lung and kidneys after PHx. The rise of
increase in the amount of uPA protein occurs by fluid shear HGF in the plasma occurs 3 hours before any rise in HGF
stress in endothelial monolayers [75]. uPA activity increases as mRNA and is a result of release of HGF stored in hepatic ECM
the first, so far, detected signal, within 1 minute after PHx [10]. [4, 5, 35, 82, 83]. It should be noted that HGF is heavily embed-
uPA, however, is produced by both endothelial cells and hepato- ded in hepatic ECM; following whole‐body elimination of the
cytes and both cells may be the source of very early rise of uPA HGF gene, the hepatic ECM concentration of HGF did not
[76]. Wnt proteins are also released by endothelial cells, and this decrease until after two sequential episodes of LR triggered by
may relate to the appearance of beta‐catenin in hepatocyte CCl4 [84]. The overall changes of ECM remodeling in LR are
nuclei within 5 minutes after PHx [77]. There is no study as yet, quite complex with multiple factors involved, including plasmi-
however, to assess the immediate impact of PHx on Wnt release nogen activator inhibitors (PAI), tissue inhibitors of metallopro-
by sinusoidal endothelial cells. teinases (TIMP), regulation of HGF activation, and so on. [85].
ECM proteins and associated glycosaminoglycans act as co‐
Immediate and early LR‐related signals receptors for many LR associated ligands, and also directly
following PHx (Table 45.2) transmit not fully understood altered signals through integrins
[86–88]. This will be discussed in relation to signals controlling
Urokinase plasminogen activator (uPA) termination of liver regeneration (see below, Termination of
and ECM remodeling liver regeneration).
uPA activity in liver increases within 1 minute after PHx [10]. In
Mobilization of Wnt/beta‐catenin and Notch signaling
addition to its recognized roles in initiating ECM remodeling
and activation of plasminogen to plasmin (leading to a cascade There is increase in beta‐catenin in hepatocyte nuclei within 5
of activation of metalloproteinases), uPA also activates HGF by minutes after PHx and lasting beyond 6 hours [89]. This is pre-
converting the ECM‐embedded, single peptide HGF to its heter- sumably driven by Wnt derived from sinusoidal endothelial
odimeric form. There are no other agents that can be invoked for cells, although that linkage in terms of chronology is not directly
572 THE LIVER:  SIGNALS REGULATING LR

established. Notch is expressed in hepatocytes, cholangiocytes, changes in gene expression of hepatocytes during LR, the gene
and endothelial cells. Its ligand, Jagged‐1, is expressed in hepat- expression patterns associated with LR are different than those
ocytes and cholangiocytes. The active proteolytic fragment of operating during liver embryogenesis [97]. Following PHx,
Notch, NICD, appears in hepatocyte nuclei within 15 minutes hepatocytes are the first cells to exhibit rapid changes and
after PHx, and is associated with subsequent expression of responses to mitogenic signals. As mentioned above, there is
Notch gene targets [90]. This event may be connected to the rapid migration of beta‐catenin and Notch (NICD) to hepatocyte
immediate early production by stellate cells of high levels of the nuclei within minutes after PHx [89, 90]. Taub et  al. have
non‐canonical Notch ligand DLK1 within minutes after PHx reported rapid changes in hepatocyte gene expression, within 30
[91]. Both Notch and its ligand Jagged remain upregulated until minutes [98]. In preparation for mitosis, the structure of canali-
at least day 4 after PHx. The Notch changes are clearly associ- culi becomes temporarily simplified. There are changes in gap
ated with hepatocyte proliferation; addition of Jagged‐1 protein and tight junctions from 24–72 hours, regulated by p38‐MAP
in hepatocyte cultures induced DNA synthesis and silencing kinase [99]. MET and EGFR are activated within 30 minutes
RNA against Notch or Jagged decreased hepatocyte prolifera- [100]. STAT3, regulated by IL6 and JAK kinases [101] and
tion until day 4 after PHx [90]. NFκB, regulated by TNF [102] become activated in the first 3–5
hours. IL6 also regulates enhanced expression and action of
Changes of LR‐related signals in the peripheral blood C/EBPβ and has reverse effects on C/EBPα [103]. N‐terminal
Earlier studies by Moolten and Bucher [73] employed parabiotic truncated isoforms of these factors (LIP and LAP, each other’s
rats and demonstrated that PHx in one of the members of the antagonists) are generated during LR and regulate many of the
parabiotic pair led to hepatocyte DNA synthesis not only in the process of the mature C/EBP isoforms [104]. Hepatocyte
operated rat but also in the liver of the non‐operated one. Similar FoxM1b plays a key role for all events associated with progres-
findings were also noted in hepatocytes transplanted to adipose sion through the S‐phase and mitosis, and its expression
tissue following PHx of the in situ normal liver [92]. These find- increases at 24 hours, prior to the G1/S interphase. In mice defi-
ings clearly demonstrated the rise of regenerative signals in the cient in FoxM1b, there is delay of DNA synthesis associated
peripheral blood following PHx. Following several decades of with increase in p21 [105, 106]. Genes associated with genera-
studies, we now know that many signals associated with LR rise tion of IPS cells, including Oct3/4, Nanog, and Myc, increase
in the peripheral blood after PHx. HGF and NE rise within less very significantly after PHx [107]. Critical signals affecting
than an hour, with overlapping but subsequent rise of IL6, TNF, hepatocyte entry into S‐phase are regulated by cyclins D1 and
and TGFβ, serotonin, and so on. (see above sections describing D2. They both turn on expression of multiple genes associated
changes related to the specific signaling molecules) [4, 5, 35, 82, with cell proliferation. In addition, cyclin D1 regulates tran-
83]. An often‐overlooked consequence of PHx is the acute drop scription of enzymes in pathways of metabolism of carbohy-
of glucose concentration in the peripheral blood. The importance drates, lipids, and amino acids [108]. Cyclin D1 expression is
of this for stimulation of LR was clearly demonstrated by admin- regulated at the translational level by mir‐21 [109]. Cyclins D1
istration of dextrose following PHx; this resulted in temporary and D2 regulate activity of cyclin dependent kinases (CDK),
suppression of DNA synthesis at day 2 in mice. Surprisingly, which mediate many of the cyclin D1/2 events. CDK activities
most of the changes in hepatocyte transcription factors, activa- are inhibited by the p53‐regulated protein p21 [110]. The
tion of HGF, and so on, were not affected; there was, however, a expression of p21, however, rapidly increases within few hours
rise in the mitoinhibitory signaling proteins p21 and p27 and after PHx [111], suggesting a complex time‐dependent interplay
suppression of FoxM1, a major regulator of cyclin D1 and many of regulatory events, guaranteeing precision of chronology in
other proteins associated with hepatocyte DNA synthesis [93]. In the progression of the various steps associated with hepatocytes
addition to specific blood signals, platelets attach to the sinusoi- in G1. The chronology of these events varies between species.
dal endothelial cells and become activated at the early stages of Peak of hepatocyte DNA synthesis is seen at 24 hours in the rat
LR [94]. Platelets contain a mixture of stimulatory (HGF, seroto- but at 36–48 hours in the mouse. However, not all hepatocytes
nin) and mitoinhibitory (TGFβ) signals. enter into DNA synthesis at the same time. LR and DNA syn-
thesis proceed from the periportal to the pericentral areas of the
lobules, with time kinetics also varying between species [112].
Cell proliferation kinetics Klochendler et al. were able to isolate cells from different stages
and intercellular signaling of the cell cycle and documented decreased expression of genes
characteristic of mature hepatocytes during DNA synthesis
Hepatocytes
[113]. This suppression is in part explained by the decreased
LR is not dependent on stem cells. The vast majority of existing expression and suppression of action of HNF4α very early in
hepatocytes participate in LR and undergo DNA synthesis. The LR [114]. The waveform transition of hepatocyte replication is
percent of hepatocyte participation in DNA synthesis and cell thus guaranteeing that not all hepatocytes in the liver will
proliferation varies with age. LR, however, is by no means undergo decrease in essential gene expression patterns and pre-
impaired in older mice. More than 95% of hepatocytes partici- serves sufficient homeostatic liver function during LR. A
pate in LR in rats younger than twenty months of age and the marked evidence of altered gene expression patterns in hepato-
percent decreases only to 75% for older animals [95]. LR is also cytes is the accumulation of triglyceride droplets during days 2
remarkable in that it can be repeated multiple times (up to 12 and 3 of LR. This is dependent on signaling changes associated
times reported) in the same animal and the performance of with blood‐borne signals, as it can be induced in hepatocyte cul-
regeneration is not affected [96]. While there are profound tures exposed to serum of partially hepatectomized rats [115].
45:  Liver Regeneration 573

Lipid accumulation in hepatocytes in G1 is associated with PHx in mice with germline deletion of TGR5 is associated with
induction of lipogenic enzymes and regulated by leptin [116] severe jaundice, increased mortality, and hepatocyte necrosis.
and by EGFR but not by MET [19]. Caveolin is present at the Melatonin has overall inhibitory effects on cholangiocyte prolif-
interface between lipid droplets and cytoplasm and regulates eration [126] whereas histamine released by mast cells has stim-
lipid processing and metabolomics of regeneration [117]. ulatory effects [127]. While these effects are well documented,
Altered gene expression patterns are also affected by expression there is no detailed analysis of their impact on cholangiocyte
of multiple miRNA at different phases of the cycle, operating at proliferation during LR. Cholangiocytes also produce platelet
the translational level of gene expression [118]. Changes in derived growth factor (PDGF), which is mitogenic for stellate
Hippo pathway kinases and Yap also occur during LR. The cells [128]. It should be noted that even though then term “chol-
Hippo pathway kinases MST1/2 and LATS1/2 phosphorylate angiocytes” is used to characterize the entire biliary epithelial
and deactivate nuclear Yap, whereas nuclear levels of Yap are compartment, there is evidence from multiple sources that there
associated with regulation of liver size [119]. Yap controls are many subtypes, characterized by size and/or expression of
expression of amphiregulin, an EGFR ligand involved in LR specific receptors and the overall regenerative biology of these
regulation (see above section on Epidermal growth factor recep- subtypes may differ [129].
tor). In normal liver, Yap is expressed in cholangiocytes, with
little or no detectable expression in hepatocytes [120]. Yap Endothelial cells
increases in hepatocyte nuclei within 24 hours, associated with
Sinusoidal endothelial cells enter into proliferation later than
decreased activity of the Hippo kinases. Levels of Yap and
hepatocytes and cholangiocytes, with active DNA synthesis
enhanced activity of Hippo pathway return to the lower normal
peaking at days 4–7 after PHx. Their proliferation is associated
by day 7 [119]. Through currently unexplained pathways,
with activation of VEGF receptors 1, 2, angiopoietin receptors
nuclear Yap in hepatocytes during LR is regulated by Hedgehog‐
Tie 1, 2, PDGFRβ, EGFR, and MET [130]. Since hepatocytes
mediated actions on stellate cells [67]. Autophagy regulation is
are the first to enter into proliferation, they form avascular small
also an important part of LR. Elimination of Atg5, an important
clusters and start synthesizing VEGF, which attracts and stimu-
component of the autophagic pathway, has severe suppressive
lates proliferation of endothelial cells penetrating the clusters
effects on DNA synthesis and liver weight is restored mostly by
and establishing formation of sinusoids [131]. This activity is
hepatocyte hypertrophy in ATG5 KO mice [121]. Hepatocytes
associated with stimulation of VEGFR2. Simultaneous stimula-
in the cell cycle produce and receive mitogenic signals from the
tion of endothelial VEGF receptor 1, however, induces produc-
other hepatic cells types. TGFα, FGF1/2, angiopoietins 1 and 2,
tion of HGF by the endothelial cells, which contributes toward
VEGF, and GM‐CSF expressed and secreted by hepatocytes are
stimulation of hepatocyte proliferation [13]. The endothelial
known to be mitogenic and probably have paracrine effects on
cells in the newly formed capillaries gradually transform into
stellate cells, endothelial cells, and Kupffer macrophages, con-
fenestrated endothelial cells through a complex process [132].
tributing to formation of proper lobular histologic microarchi-
Recent studies have shown that LR is also associated with
tecture [4, 5, 35, 82, 83]. Hepatocytes also receive newly
migration of endothelial precursors from bone marrow to the
synthesized HGF from stellate and endothelial cells throughout
liver, which convert from typical endothelial cells into fenes-
LR [13]. The net result is precise alignment of newly synthe-
trated endothelial cells and participate in the restructuring of the
sized hepatocytes along the orientation of the closest sinusoid in
hepatic sinusoids in the enlarged lobules that result from LR
a very precise manner [122]. Though the above events related to
[14]. The migrating sinusoidal progenitor cells (“sprocs”)
hepatocyte proliferation during LR, several studies suggest that
actively produce HGF. Their migration is controlled by hepatic
even after 2/3 PHx, and more definitively after hepatectomies
production of circulating VEGF and mediated via stromal cell
less than 2/3, the pathways involving restoration of liver weight
derived factor 1 (SDF1) [133]. HGF and TGFβ1 are known to
involve not only hepatocyte proliferation but also hepatocyte
stimulate tubule formation in endothelial cells and are very
hypertrophy, with hepatocyte enlargement being a solid compo-
likely play a role in the restructuring of the sinusoidal network
nent of the overall LR strategy [123, 124]. The signaling path-
[4, 5, 35]. Despite complex pathways of cell migration, endothe-
ways controlling regulation of the precise percent of involvement
lial cells regulate LR in many ways by producing cytokines
of hepatocyte proliferation versus mere hypertrophy are not
favoring regenerative response. These “angiocrine” effects were
understood at this point.
described recently by Ding et al. [134].

Cholangiocytes Stellate cells


Proliferation of cholangiocytes in portal ductules follows the These cells exist in multiple organs of endodermal origin (lungs,
same time frame as hepatocytes. Cells of the biliary system pancreas, etc.). Hepatic stellate cells store vitamin A in lipid
respond to the same tyrosine receptor kinases (MET and EGFR) droplets, and they express many genes in common with glial
as hepatocytes. LR proceeds by enlargement of existing lobules; cells of the brain, even though stellate cells are derived from
there is no formation of new portal triads following PHx. The nestin‐positive cells of the cardiac mesenchyme and not the neu-
receptor TGR5, a G‐protein couple receptor, also regulates pro- ral crest. There is strong evidence that stellate cells function as
liferation of cholangiocytes in culture in response to bile acids, part of a neuroendocrine control of normal liver and respond to
with some of the bile acids having stimulatory or inhibitory parasympathetic and sympathetic innervation [135]. They also
effects [60]. TGR5 has complex effects during LR, protecting function as antigen presenting cells for NK cells [136]. Most of
hepatocytes from increase in biliary flux from the intestine [125]. the studies related to stellate cells focus on pathways resulting
574 THE LIVER:  SIGNALS REGULATING LR

in stellate cell “activation” during chronic liver injury, and their regenerated lobules at day 7 of LR, in parallel with the sinusoi-
contribution to liver fibrosis and cirrhosis [137]. Their participa- dal endothelial cells. Increase in desmin‐negative cells resem-
tion in LR after PHx has also focused mostly on their contribu- bling immature stellate cells is seen at day 1 of LR [138].
tion of signaling molecules (including HGF, TGFβ, epimorphin,
pleiotropin, etc.) and the production of most of ECM proteins in Kupffer cells (hepatic macrophages)
different stages of LR. Despite their vital role in normal and
regenerating liver, there is no detailed study of their prolifera- Several studies have documented that the Kupffer cell mac-
tion rates during LR. In normal liver, stellate cells make contact rophages lining the hepatic sinusoids, in addition to standard
with both hepatocytes and sinusoidal endothelial cells by very macrophage functions (phagocytosis of particles, participation
long processes (Figure  45.1). This relationship suggests that in eliminating tissue debris after local damage, etc.) also have
there are regulatory pathways operating by direct communica- some unique properties, related to aspects of immunity in which
tion between stellate cells and the cell they contact; these inter- liver is an active participator. During LR, there is evidence that
actions need to be explored. Histologic observations demonstrate many Kupffer cells proliferate locally [139]. There is evidence,
that stellate cell numbers increase in parallel with the restoration however, that mononuclear cells deriving from bone marrow
of the sinusoidal network. Stellate cell numbers peak in also enter the liver and become typical Kupffer cells during LR
[140]. Kupffer cells are recognized as F4/80+ and there distin-
guished into two categories. The CD11b+ Kupffer cells appear
to be derived from bone marrow and their proportion increases
at day 3, after the peak of hepatocyte proliferation. The CD68+
Kupffer cells do not seem to change in numbers during LR
[141]. Proliferating hepatocytes produce GM‐CSF, a mitogenic
signal for Kupffer cells; stellate cells produce M‐CSF, and
Kupffer cells produce IL6, TNF, TGFβ, and TGFα during LR
[4, 5, 35]. Overall, there are no precise measurements of popu-
lation changes of Kupffer cells during LR but their numbers
vary along with their participation in restoration of sinusoidal
vascular network.

Termination of liver regeneration


Initiation of LR can be precisely timed by the performance of
PHx. Termination of LR, however, is more difficult to define. If
the liver weight to body ratio is to be used as a criterion, normal
ratios (depending on species) are approached by end of the sec-
ond week. In terms of hepatocyte proliferation, a return of nuclear
DNA synthesis to normal levels is approached by days 5–6 [4, 5,
35]. Analysis of liver transcriptomics, however shows a slower
return to normal gene expression values by some time beyond day
14 after PHx [88] (Figure 45.2). Several signaling molecules have
been considered as potentially associated with “termination” of
LR. Hepatocyte‐specific TGFβ1 transgenic mice have normal
liver regeneration, even though TGFβ is mito‐inhibitory for
hepatocytes [142]. Hepatocyte targeted elimination of TGFβ
receptors or activin receptors does not prolong liver regeneration;
extended liver regeneration is seen only when both receptors are
eliminated [143]. An important regulator associated with termina-
tion of liver regeneration and restoration of normal liver function
is likely to be ECM. Addition of ECM preparations (collagen
gels, Matrigel) stabilizes differentiation but inhibits proliferation
of hepatocytes in culture [6]. ECM is partially degraded and
remodeled at the beginning of LR, but it is restored toward the
end of LR [40, 144, 145]. Signaling between ECM and hepato-
Figure 45.1  Intricate connections between stellate cells (SC), sinu- cytes is mediated by integrin α3β1; the β1 intracellular domain
soidal endothelial cells (SEC) and hepatocytes (Hep). A single stellate makes contact with integrin linked kinase (ILK). The latter trans-
cell (green) is making contacts with multiple other cells. The nature mits growth suppressor and differentiation enhancement signals
and purpose of these contacts is not fully understood. They are, how-
ever, likely to be involved in bilateral communications regulating
through multiple pathways [87]. Hepatocyte‐specific elimination
extracellular matrix deposition and exchange of growth‐regulatory of ILK prolongs liver regeneration and results in livers of size
signaling molecules. Image courtesy of Dr. Donna Stolz, University of 158% of the ­original. All protein partners associated with ILK
Pittsburgh. increase during LR until the end of hepatocyte proliferation, by
45:  Liver Regeneration 575

Figure 45.2  Expression of the top 250 genes expressed in normal mouse liver at different times after 2/3 partial hepatectomy (PHx). There is rapid
up and down change in expression of most of the 250 genes within 1 day after PHx, with changes in both direction continuing until beyond day 14
of regeneration. This makes it difficult to determine a precise point at which the processes of LR should be considered as “terminated”. A few
specific genes are massively overexpressed during LR, exceeding expression values of any gene seen in normal resting liver. The reason for the
selective massive overexpression of a few genes is not clear. Mup1 is not present in humans. Selenoprotein P1 is the major vehicle by which liver
sends selenium to peripheral tissues, essential for regulation of redox functions in most cells of body. Transthyretin is the major transport vehicle
in plasma for thyroxin (T4) and retinol bound to the retinol binding protein. Cyp2c9 protein is approximately 18% of all proteins of members of the
liver Cyp family and is associated with metabolism of most xenobiotics and many endogenous metabolic compounds including arachidonic acid.
Jak3 gene encodes JAK3 kinase protein, associated with gp130 receptor of IL6 and activating STAT3 transcription factor. Psg28 is a glycoprotein
produced in liver during pregnancy. There is no known function for this protein in liver regeneration.

day 6 [88]. These results suggest that ECM signaling is important elimination of single receptor tyrosine kinases (MET and
for regulation of termination of liver regeneration, but the path- EGFR). The flexibility of cybernetics of hepatocyte transcrip-
ways remain to be further understood. Glypican 3 (GPC3), a GPI‐ tomics, which allows overcoming such obstacles, is remarkable.
linked plasma membrane protein over‐expressed in liver cancer Absence of many of the extracellular signals normally involved
[146], is nonetheless a growth suppressor for most organs, as to initiate hepatocyte DNA synthesis (e.g. IL6 and STAT3; TNF
demonstrated by the Simpson–Golabi–Bechmel syndrome in and NFκB) can be corrected by MET and EGFR [45]. Systemic,
humans with loss of function of GPC3 [147]. Hepatocyte‐specific whole‐body, elimination of the HGF gene does not affect LR,
transgenic expression of GPC3 results in substantially suppressed because the heavily embedded HGF in ECM is sufficient in
LR [148]. GPC3, normally interacting with tetraspanin CD81, is quantity to support several LR episodes prior to be depleted
involved during LR with many complex interactions with differ- [84]. The only extracellular signaling intervention that com-
ent pathways, including Hedgehog, Wnt, and Hippo signaling. pletely abolishes LR is the combined elimination of signaling of
GPC3 dissociates from CD81 during LR but associates with it both MET and EGFR [19]. When signaling of both of these
again at days 6–7 [65]. Recent evidence has shown that GPC3 receptors is eliminated 5 days prior to PHx, there is a small
bound to CD81 enhances activity of the Hippo pathway and amount of hepatocyte proliferation at post‐PHx day 2; prolifera-
results in decreased levels of nuclear Yap. This may be a very tion, however, stops entirely after that. This is followed by
important regulating signal by which GPC3, via Yap, contributes decrease in hepatocyte size to 35% of normal and decrease in
to the termination of LR [149]. LSP‐1 is another protein, deleted the size of hepatic lobules. There is no net gain in liver size by
in about 50% of HCC, and regulating the RAF‐MEK‐ERK path- the end of day 14. There is significant decrease in hepatocyte‐
way. Deletion of hepatocyte LSP‐1 prolongs LR and transgenic specific gene expression, especially for genes involved in lipid
expression suppresses LR [150]. Beta‐catenin is an important and sterol biosynthesis, glycogen synthesis, and urea cycle.
mitogenic signal contributing to LR; beta‐catenin signaling is There is no activation of mTOR and AKT, enhanced activation
downregulated by Wnt5a whose expression also increases toward of PTEN and AMPKα, and inactivation of enzymes related to
the end of LR and restricts further beta‐catenin signaling [151]. fatty acid synthesis and redox regulation. Despite these pro-
found changes, there is dramatic upregulation of expression of
Overcoming deficiencies of extracellular genes involved in plasma protein synthesis and all members of
signals and the critical role of receptor CYP family genes (with the exception of Cyp7a1, the first
enzyme involved in bile acid synthesis). Mice die by day 14
tyrosine kinases
after PHx, with low glucose, marked ascites, and high ammonia
Elimination of single extracellular signals (IL6, TNF, Wnt/beta‐ levels in peripheral blood. Remarkably, there is no increase
catenin, Hedgehog, bile acids, etc.) involved in regulation of in  hepatocyte death, but there is increased apoptosis of non‐
LR, delays but does not abolish LR. This also applies to parenchymal cells, especially stellate cells. Mice die not because
576 THE LIVER:  SIGNALS REGULATING LR

liver dies, but because liver ceases to function in the absence of other signaling pathways occurs when these two receptor tyros-
activation of EGFR and MET. The same phenomena, with some ine kinases (RTK) are simultaneously abolished. The findings
variations in activation pathways, were observed and resulted may have implications for pathogenesis in human liver failure.
in death of normal, non‐hepatectomized mice [152]. In both Plasma HGF in fulminant hepatitis rises to high levels, but is not
of those studies, MET and EGFR signaling was eliminated activated [153], in contrast to the active HGF released in plasma
throughout the body. Yet no abnormalities were detected in after PHx [154]. There is also closure of the fenestrae of the
other tissues. Proliferation of stem cells in intestinal crypts was sinusoidal endothelial cells in fulminant hepatic failure, poten-
unaffected. Why is liver so uniquely dependent on activation of tially interfering with EGF availability [155].
EGFR and MET? (Figure 45.3). Both receptors are activated by
Tyr phosphorylation in normal resting liver [19]. Activation of Facultative stem cell relations between
MET can be explained by the heavy presence of HGF in hepatic
hepatocytes and cholangiocytes
ECM [84]. The story for EGFR is more complex. As mentioned
above, EGF is synthesized by the Brunner’s glands residing Hepatocytes and cholangiocytes are the two epithelial cell types
under the duodenal mucosa [21]. EGF protein is directly in the liver. They typically enter into proliferation to restore the
secreted in the intestinal lumen, and a portion is absorbed intact proper cell numbers in their own compartment, as required for
and provided to hepatocytes continually through the portal cir- normal liver function. This is done without participation of stem
culation [21]. The observations for constant supply of HGF and cells. There are situations, however, in which either hepatocytes
EGF to hepatocytes suggest that the combined signaling of HGF or cholangiocytes cannot enter into proliferation to repair defi-
and EGF generates a fundamental cybernetic platform required ciencies in their own compartment. In such situations, hepato-
for the proper coordination of all other signaling pathways char- cytes and cholangiocytes follow steps of transdifferentiation to
acteristic of normal hepatocytes. Collapse of coordination of the transform themselves into each other [156]. This plasticity of

Figure 45.3  Dependence of hepatocytes on HGF/MET and EGFR signaling. Hepatocytes are continually exposed to HGF and EGF. HGF is
produced by stellate cells and deposited as inactive single peptide form in the extracellular matrix. EGF is produced by Brunner’s glands of the
duodenum, shown by black arrows in the histologic picture (hematoxylin eosin stain; magnification 100×). The histology of Brunner’s glands is
similar to salivary glands, which produce and secrete EGF in saliva. EGF is produced and secreted in the intestinal lumen, and then a portion is
absorbed intact and enters the liver through the portal circulation. HGF receptor MET and EGFR are activated in normal, resting liver. For details
and references, see text.
45:  Liver Regeneration 577

phenotype confers regenerative advantages to the liver, espe- centrilobular damage induced by CCL4 than after regeneration
cially in situations of liver failure. It is much faster to restore a following PHx. In AAF‐suppressed centrilobular damage, the
hepatocyte or cholangiocyte compartment by a rapid transdif- progenitor cells are likely to appear from the cholangioles that
ferentiation of one type of cell to another, than to rely on a slow penetrate deep into the hepatic lobule (canals of Hering) [169].
growth and expansion of a true stem cell compartment. There Lemaigre has traced the histology of formation of the ductal
are now numerous studies that have documented the capacity of plate during embryogenesis and demonstrated that most newly
either cell type to enter into proliferation, transdifferentiate, and formed ductal cells migrate into the space next the portal and
restore the numbers of cells in the other compartment suffering arterial branches to form the ductules of the portal triads. Cells
with defective proliferation [156]. This has been shown in sev- that fail to migrate, revert to hepatocytes and remain immedi-
eral experimental models in rats and mice. Blockade of rat liver ately proximal to the portal triad [170]. Studies in the rat dem-
regeneration by acetylaminofluorene (AAF), resulting in onstrated that these hepatocytes preferentially contribute to
expression of DNA adducts and expression of p21 in hepato- formation of biliary ductules [163]. In hepatic organoids in
cytes [157], causes the appearance of a transitory population of roller bottle cultures maintained in the presence of insulin, HGF,
“oval” cells with morphologic properties intermediate between and EGF, epithelial cells retain a hepatobiliary phenotype with
hepatocytes and cholangiocytes [158]. These cells eventually mixed gene expression patterns. Addition of corticosteroids
evolve into mature hepatocytes [159]. The emergence of the induces the separation of the mixed lineage and the emergence
oval cells from cholangiocytes in rats is supported by the expres- of mature biliary cells immediately under the culture medium
sion of hepatocyte‐associated transcription factors in cholangio- and mature hepatocytes in the tissue underlying the biliary cells.
cytes immediately after AAF‐PHx [160] and the lack of Conversion of hepatocytes to biliary cells in that system cannot
emergence of oval cells by pretreatment with the cholangiocyte‐ occur in the absence of HGF and EGF (either one alone can
specific toxin DAPM (4,4’‐diaminodiphenylmethane) prior to sustain the conversion) [39, 171]. In the same system, either
AAF‐PHx [161]. Mice cannot activate AAF, and the evidence HGF or EGF induce TGFβ expression. Given the above recent
for cholangiocyte conversion to hepatocytes had to exploit more findings by Schaub et al. [166] on the role of TGFβ involved in
difficult pathways that cause enhanced expression of p21 in this transdifferentiation, it is likely that not only HGF and EGF,
hepatocytes. This was achieved by Forbes et  al. in mice with but also TGFβ, are involved in the conversion of hepatocytes to
hepatocyte specific deletion of Mdm2, an E3 ubiquitin‐protein cholangiocytes in the organoid cultures. It should be noted that
ligase which specifically degrades p53. In the absence of whereas cells of mixed hepatobiliary phenotype are rarely seen
MDM2, there is enhanced expression of p53 which subse- in normal human liver, they become the predominant type in
quently induces p21. There is increased hepatocyte death and acute fulminant hepatitis, regardless of etiology [172]. It is not
senescence followed by the appearance of “progenitor” cells clear whether this is a pathway to healing of fulminant hepatitis
with properties similar to the oval cells seen in rats. The pro- and restoration of normal structure, because the evolution of the
genitor cells converted into hepatocytes and restored hepatocyte hepatobiliary cells has not been traced in the few cases of fulmi-
populations. The biliary origin of these progenitor cells was nant hepatitis that heal spontaneously.
established by detecting the newly emergent hepatocytes as car-
rying lineage tagging performed on cholangiocytes [162]. Regeneration following xenobiotic‐induced
Extended studies have also documented the transformation of
hepatocytes to cholangiocytes in settings where cholangiocyte
centrilobular liver injury
proliferation is expected but cannot occur, as, for example, in Hepatocytes in the central portion of hepatic lobules express the
rats pretreated with DAPM and subjected to bile duct ligation (phase I) CYP family and (phase II) other enzymes involved
(BDL). Up to 50% of the newly produced biliary ductules in the with metabolism and processing of xenobiotics. In most
DAPM–BDL model carry markers specific for hepatocytes instances, xenobiotics are appropriately processed for elimina-
[163]. Studies with hepatic organoids demonstrated that this tion through blood (kidneys) or bile (feces). In rare instances,
conversion is dependent on EGFR and MET [164]. Similar however, the processing of some xenobiotics results in genera-
results have been shown in mice. Conversion of hepatocytes to tion of reactive electrophiles, capable of reacting with the nucle-
cholangiocytes was shown in mice fed diethyldithiocarbamate ophile residues of proteins and nucleic acids, resulting in death
(DDC) diet [165]. The most recent evidence for hepatocyte of centrilobular hepatocytes. Experimental models to study this
transdifferentiation to cholangiocytes was recently presented by injury and its repair pathways have most often utilized carbon
Huppert and Willenbring, who demonstrated a complete de tetrachloride (CCl4) or acetaminophen as the damaging agents,
novo generation of a biliary system from hepatocytes by path- in rats and mice [173, 174]. The extent of centrilobular necrosis
ways controlled by TGFβ [166]. The pathways mediating this and the lethality depend on the dose of the xenobiotic, and the
interconversion are complex and, in addition to TGFβ, also lethality of doses vary with the strain of the animals, other die-
involve Hippo/Yap [120] and Notch [167]. It is not clear whether tetic components, and so on, as these parameters affect gene
all hepatocytes, anywhere in the lobule, can participate in these expression patterns in the centrilobular regions [175]. Overall,
interconversions. Demetris et  al. demonstrated cells with this type of injury is more representative of the type of damage
mixed  expression of hepatocyte and cholangiocyte transcrip- seen in human livers, especially given the fact that inappropri-
tion factors at the terminal end of the canals of Hering, suggest- ately high doses of acetaminophen are often used for suicide,
ing a population of pre‐existing cells in normal liver with resulting in massive liver necrosis, salvaged, whenever possible,
progenitor cell phenotypes [168]. Also, surprisingly, in LR sup- with liver transplantation. The first evidence of this type of
pressed by AAF, there are more progenitor cells appearing after injury is a hepatocyte necrosis in the centrilobular areas. This is
578 THE LIVER:  SIGNALS REGULATING LR

followed by invasion of monocytes, which proliferate very continuous DNA labeling of hepatocytes in normal liver claimed
actively in the site of damage and produce active macrophages. a slow streaming of proliferating hepatocytes arising in peripor-
The latter proceed to remove the area of necrosis. There is rapid tal areas and descending to pericentral areas of the lobule [178].
activation of EGFR within minutes after acetaminophen dose Subsequent studies with pulse labeling of DNA, however,
[176]. Hepatocytes enter into the cell cycle (proliferating cell showed that replicating hepatocytes could be found in all areas
nuclear antigen [PCNA]‐positive) by the end of day 1, primarily of the hepatic lobule, not just in periportal, and that “streaming”
around the area of necrosis. Most of the cells of the rest of the of hepatocytes was not the mechanism for maintaining liver size
lobule, however, become PCNA‐positive by the end of day 2. [179]. There have been several major recent studies, that have
Hepatocytes positive for Ki‐67, a biomarker for active DNA provided new information about the role of different hepatocyte
synthesis, appear at day 2 and increase to cover the entire lob- subpopulations in maintenance of the “hepatostat”. Studies by
ule. HGF levels increase within the liver and in the peripheral Nusse et  al. documented a unique population of hepatocytes
blood [81]. mRNA for TGFα and HGF increased in two peaks, immediately around the central venules. These cells express
at 12 and 48 hours [177]. In addition to contributions of growth Axin2, a protein which is part of the beta‐catenin ubiquitination
factors and stimuli seen in LR, macrophages may also contrib- complex. These cells express Tbx3, a biomarker for fetal hepa-
ute mitogenic stimuli, as they are known to produce HGF, toblasts, and are diploid. They are also positive for the beta‐
TGFα, TGFβ, PDGF, IL6, and TNF. The chronology of expres- catenin dependent glutamine synthetase. Wang et  al. tagged
sion of these signals from macrophages in the context of repair these cells and documented that over time they generate a line-
of this type of injury has not been assessed. The fact that the first age of cells which expand around the central area and lead to
PCNA‐positive hepatocytes at day 1 first emerge around the progeny covering up to 40% of the cells of the lobule [180]. The
area of necrosis suggests that mitogenic stimuli produced by results were impressive, though there is a concern that beta‐
macrophages, actively infiltrating and proliferating in that area, catenin levels may be elevated in these cells. There is only one
must play a very early, perhaps the first, role. The rate of removal of the Axin2 genes active in this model, and the results of poten-
of the necrotic hepatocytes and the restoration of normal tial Axin2 haplo‐insufficiency may cause elevated levels of
­histology depends on the dose of the offending agent. Normal beta‐catenin, enhancing the proliferative rate of the affected
histology, with moderate experimental doses, is usually seen hepatocytes. From the other pole of the lobule, the periportal
within one week. The mechanistic details of repair of this type area, Furuyama et al. provided evidence that Sox9+ lineage pre-
of injury, though clinically relevant, are less defined compared cursors in liver and pancreas generate progeny of cells that grad-
to LR. This is due to the comingling of inflammatory events at ually expand toward the other areas of the lobule, differentiating
the earliest stages, making it difficult to distinguish signals into Sox− hepatocytes and pancreatic acinar cells [181]. Sox9 is
strictly related to cell proliferation versus inflammation. An expressed in cholangiocytes, and in that study, it was thought
additional issue is the variation of the response based on the that cholangiocytes continually differentiate into hepatocytes
dose of the offending agent [175]. It should be stated, however, maintaining liver mass homeostasis. Other studies, however,
that all mitogenic signals identified in LR have also been demonstrated that immediate periportal hepatocytes also express
identified in this type of injury, though their sources and time‐ low levels of Sox9 and they also proliferate to establish expanded
kinetics are not be identical. progeny migrating toward the midzonal and centrilobular areas,
especially after chronic liver injury [182]. These cells are the
Maintenance of steady mass in normal same hepatocytes earlier identified as being responsible for gen-
erating biliary ductules in inhibited biliary repair [163], previ-
resting livers
ously identified by Lemaigre et al. as remnant cells of the ductal
In all previous sections of this chapter we discussed the mito- plate which failed to enter into the portal triad and returned to
genic signals and complex responses triggered by massive loss hepatocyte differentiation [170]. Thus, the findings of Furuyama
of liver parenchyma. The elicited hepatocyte proliferation is a et al. could be explained in a different way; the lineage described
compensatory response, aimed to restore the full functional by Furuyama did not originate from cholangiocytes, but from
capacity of liver. The loss of hepatocytes is the primary reason Sox9+ immediate periportal hepatocytes. Given, however, the
for the emergence of signaling pathways driving the compensa- earlier studies with pulse labeling which demonstrated that
tory response. In a normal resting rodent liver, however, at any hepatocytes can be tagged in DNA synthesis in all areas of the
given moment, the percent of dead or proliferating hepatocytes lobule, it is logical to consider that both the centrilobular and
typically does not exceed 0.2% of the total number of cells. periportal hepatocyte lineages operate, proceeding in opposite
Even though this would involve very small numbers of cells, directions, and replenishing hepatocytes. In this scenario, DNA
there must be pathways that trigger compensatory proliferation pulse labeling would tend to label hepatocytes in all areas of the
at a micro‐scale, so that the total loss over time does not result lobule. Supporting evidence from a pan‐lobular hepatocyte par-
in decreased liver mass. The mechanisms underlying this “hepa- ticipation in the process of continual renewal came from another
tostat” are not well understood. It is reasonable to postulate that study. Tchorz et al. demonstrated that liver zonation depends on
the signals regulating this process may be different than the sig- a Wnt/beta‐catenin gradient with intensity increasing from peri-
nals elicited by a large hepatocyte loss. As mentioned above, portal to centrilobular areas, and that the formation of this gradi-
based on constant availability of HGF (from ECM) and EGF ent is dependent on angiocrine signals mediated through
(from the Brunner glands of duodenum), MET and EGFR are R‐spondin (RSPO) ligands and their receptors (Lgr4/5). Lgr4+
always activated in normal liver, though we do not know if this hepatocytes had enhanced proliferative capacities during LR
applies to all or a fraction of hepatocytes. Early studies using and they were located throughout the lobule [183]. Finally, in a
45:  Liver Regeneration 579

very recent study by Artandi et al., lineage tagging of subset of functions of such missing alleles are provided by the presence
hepatocytes expressing high levels of telomerase demonstrated of several copies of the single missing chromosome, present in
that they exist in all zones of the lobule and they give rise to the polyploid status. As hepatocytes revert to diploidy, however,
tagged progeny that can over time cover the entire lobule [184]. some of the diploid cells are likely to randomly have only single
It is very likely that all these pathways (periportal, pericentral, copies of some chromosomes. For example, an 8n polyploid
pan‐lobular) operate at the same time to give progeny that main- hepatocyte missing a single copy of chromosome 13 would gen-
tains a sufficient population of hepatocytes to guarantee ade- erate four diploid hepatocytes of which one would only have a
quate mass for liver function. What is not identified however, is single copy of chromosome 13. This has implications for devel-
perhaps the most crucial function of the hepatostat. What are the opment of liver cancer. In chronic liver disease, hepatocyte pro-
stimuli that regulate all these processes and guarantee that liver liferation takes place in an inflammatory environment with
size does not exceed the required size? associated genotoxic products, including reactive electrophiles,
lipid peroxides, O2− radicals, and so on. Any inflicted genotoxic
damage on the single copy chromosome at the diploid stage
Chronic regeneration in chronic liver disease would not be balanced by the genes of the missing chromosome.
Most chronic liver diseases are associated with attrition of In this setting, otherwise non‐dominant mutations in the exist-
hepatocytes, to a variable extent. This occurs regardless of etiol- ing single chromosome copy may function as dominant (driver)
ogy, be that viral, toxic (including alcohol), metabolic, steato- mutations. This enhances the risk of developing neoplasia. In
hepatitis, and so on. Loss of hepatocytes triggers regenerative view of this, it should not be a surprise that all types of chronic
responses. These responses can be documented by enhanced liver disease are associated with increased risk for development
immunohistochemistry for biomarkers associated with hepato- of hepatocellular carcinoma [1]. The risk for developing liver
cyte proliferation (PCNA, Ki67), as well as biomarkers associ- cancer is also enhanced in situations in which proliferation of
ated with hepatocyte death (TUNEL assays, expression of hepatocytes is overall impaired. The classic experiment by Solt
activated caspases). The precise signaling pathways triggering et al. demonstrated that “resistant hepatocytes” against any par-
the regenerative response are likely to be similar with those dis- ticular mito‐inhibitory block will rapidly develop into nodules
covered following acute injury, discussed above. But, there also of monoclonal hepatocyte growth when challenged with acute
likely to be peculiarities related to the injurious agent, probably liver injury (partial hepatectomy), and the resultant nodules will
affecting regenerative signaling in different ways. The supreme allow restoration of the hepatostat (liver to body weight ratio)
“prerogative” of loss of hepatocytes is to cause compensatory [189]. These “resistant hepatocytes” are more exposed to the
hepatocyte proliferation, in order to maintain the required liver risks of developing aneuploid diploidy due to their enhanced
to body weight ratio, best thought of as a “hepatostat”. In con- rates of proliferation. The few “resistant cells” are the only ones
trast to any other organ in the body, a required and fixed amount that can proliferate to meet hepatostat demands for the whole
of liver tissue is needed for body homeostasis, and on that regu- liver. The phenotypes of liver cancer in chronic liver disease due
latory principle, loss of liver “mass” is not tolerated. This phe- to errors of metabolism associated with chronic inflammation,
nomenon is unique to the liver; it is not the case with other are a clear example of this situation. Most cancers developing in
organs (with the exception of endocrine glands under control by hemochromatosis do not store excessive iron; cancers emerging
pituitary). Loss of half of pancreas, or one of the kidneys, is not in glycogen storage diseases, lipid storage diseases, and so on,
associated with regeneration to increase the remnant tissue to in their majority do not store the abnormal metabolic product;
the original size of the total pancreas, and the residual kidney most cancers seen in patients with alpha‐1 antitrypsin deficiency
becomes slightly larger but not twice the weight [185]. As a do not carry the characteristic PAS‐diastase positive globules
consequence of the expectations of the hepatostat, there is characteristic of the misfolded ATZ protein [1]. In all of these
enhanced proliferation of the surviving hepatocytes at any time situations, hepatocytes that for random reasons do not carry the
there is continuous hepatocyte loss. This chronic compensatory metabolic abnormality are the “resistant hepatocytes” of Solt
proliferation of surviving hepatocytes to make up for the loss of et al. In their effort to maintain the hepatostat, their proliferation
hepatocytes induced by the chronic disease has several adverse is faster than the vast majority of hepatocytes affected with the
consequences. Several studies have shown that continuous pro- metabolic disease. In a recent study, the same phenomenon
liferation in chronic liver disease is associated with decrease in appears to apply to hepatocellular carcinomas associated with
hepatocyte ploidy. Hepatocytes in both in humans and rodents HCV infection. HCV impairs hepatocyte proliferation by inter-
are typically polyploid. The average ploidy of human hepato- acting with tetraspanin CD81 and activating the Hippo pathway,
cytes is 4n binucleate. In mice, ploidy levels reaching 8n or 16n thus causing decrease in nuclear Yap [149]. Most hepatocellular
are commonly seen. Experimental studies in rodents have shown carcinomas do not express CD81 in plasma membrane and are
that during chronic hepatocyte proliferation, the “ploidy con- thus uninfectable by HCV [190]. This makes them resistant to
veyor” reverses course and most polyploid hepatocytes revert to the mito‐inhibitory effects of Hippo activation and Yap decrease
diploid status [186]. Several studies in human liver have docu- affecting the normal hepatocytes infected by HCV. The “HCV
mented increased diploidy levels in most chronic liver diseases, resistant” CD81‐negative hepatocytes would proliferate faster
including cirrhosis [187]. Polyploidy reversal to diploidy, in and to maintain the hepatostat, and compensate for the loss of hepat-
by itself, should not be expected to have negative consequences. ocytes caused by HCV infection. In this situation, HCV func-
Unfortunately, however, a large percentage of human and rodent tions as a “promoter” for enhanced development of hepatocellular
polyploid hepatocytes are also aneuploid, randomly missing carcinomas, with most of the latter being composed of “HCV
single copies of chromosomes [188]. In a polyploidy status, the resistant” neoplastic hepatocytes.
580 THE LIVER:  REFERENCES

CONCLUDING REMARKS 13. LeCouter, J., Moritz, D.R., Li, B. et al. Angiogenesis‐independent endothe-
lial protection of liver: role of VEGFR‐1. Science, 2003;299(5608):890–3.
14. Wang, L., Wang, X., Xie, G., Wang, L., Hill, C.K., and DeLeve, L.D. Liver
Even though LR after partial hepatectomy is not representative sinusoidal endothelial cell progenitor cells promote liver regeneration in rats.
of types of liver injury most commonly seen in liver disease, it J Clin Invest, 2012;122(4):1567–73.
has been a very useful model for understanding the fundamental 15. Achim, C.L., Katyal, S., Wiley, C.A. et al. Expression of HGF and cMet in
signaling mechanisms controlling liver regeneration. It is not the developing and adult brain. Brain Res Dev Brain Res 1997;102(2):
299–303.
complicated by considerations of inflammatory phenomena 16. Wolf, H.K., Zarnegar, R., Oliver, L., and Michalopoulos, G.K. Hepatocyte
associated with acute hepatocyte necrosis. Much recent knowl- growth factor in human placenta and trophoblastic disease. Am J Pathol,
edge has surfaced for homeostatic events related to liver mass 1991;138(4):1035–43.
maintenance under normal conditions. The signaling pathways, 17. Trusolino, L., Bertotti, A., and Comoglio, P.M. MET signalling: principles
and functions in development, organ regeneration and cancer. Nat Rev Mol
however, regulating the hepatostat are not yet well understood
Cell Cell Biol, 2010;11(12):834–48.
[1]. We understand that nuclear Yap plays a role in regulating 18. Bard‐Chapeau, E.A., Yuan, J., Droin, N. et al. Concerted functions of Gab1
liver size and that combined signaling of HGF/MET and EGF/ and Shp2 in liver regeneration and hepatoprotection. Mol Cell Cell Biol,
EGFR are essential for maintaining the hepatostat; combined 2006;26(12):4664–74.
elimination of these RTK signals causes liver decompensation 19. Paranjpe, S., Bowen, W.C., Mars, W.M. et al. Combined systemic elimina-
tion of MET and epidermal growth factor receptor signaling completely
and functional collapse. It should not be forgotten at the end
abolishes liver regeneration and leads to liver decompensation. Hepatology,
of this chapter, that hepatocytes have very high, almost forever, 2016;64(5):1711–24.
proliferative capacity in liver recolonization models. In the 20. Carver, R.S., Stevenson, M.C., Scheving, L.A., and Russell, W.E. Diverse
Fah−/− model, after ten sequential recolonization events in which expression of ErbB receptor proteins during rat liver development and regen-
it was shown that both polyploid and diploid hepatocytes equally eration. Gastroenterology, 2002;123(6):2017–27.
21. Skov Olsen, P., Boesby, S., Kirkegaard, P. et  al. Influence of epidermal
participate, calculations demonstrated that 1 hepatocyte could, growth factor on liver regeneration after partial hepatectomy in rats.
in mathematic principle, generate 50 mouse livers [191]. In that Hepatology, 1988;8(5):992–6.
sense, hepatocytes are unique amongst the epithelial cells in 22. Olsen, P.S., Poulsen, S.S., and Kirkegaard, P. Adrenergic effects on secretion
our  body, and much more needs to be learned not only about of epidermal growth factor from Brunner’s glands. Gut, 1985;26(9):920–7.
what turns on their proliferation, but also what harnesses that 23. Mead, J.E. and Fausto, N. Transforming growth factor alpha may be a physi-
ological regulator of liver regeneration by means of an autocrine mechanism.
unlimited proliferative capacity and channels it to meaningful, Proc Natl Acad Sci USA, 1989;86(5):1558–62.
hepatostat‐defined, boundaries, so the liver continues to ­function 24. Berasain, C., Garcia‐Trevijano, E.R., Castillo, J. et  al. Amphiregulin: an
normally. early trigger of liver regeneration in mice.[see comment]. Gastroenterology,
2005;128(2):424–32.
25. Khai, N.C., Takahashi, T., Ushikoshi, H. et  al. In vivo hepatic HB‐EGF
gene transduction inhibits Fas‐induced liver injury and induces liver regen-
eration in mice: a comparative study to HGF. J Hepatol, 2006;44(6):
REFERENCES 1046–54.
26. Mitchell, C., Nivison, M., Jackson, L.F. et  al. Heparin‐binding epidermal
1. Michalopoulos, G.K. Hepatostat: liver regeneration and normal liver tissue growth factor‐like growth factor links hepatocyte priming with cell cycle
maintenance. Hepatology, 2017;65(4):1384–92. progression during liver regeneration. J Biol Chem, 2005;280(4):2562–8.
2. Ware, A.J., Eigenbrodt, E.H., Shorey, J., and Combes, B. Viral hepatitis compli- 27. Huang, X., Yu, C., Jin, C. et al. Ectopic activity of fibroblast growth factor
cating the Dubin–Johnson syndrome. Gastroenterology, 1972;63(2):331–9. receptor 1 in hepatocytes accelerates hepatocarcinogenesis by driving prolif-
3. Higgins, G.M. and Anderson, R.M. Experimental pathology of the liver, 1: eration and vascular endothelial growth factor‐induced angiogenesis. Cancer
Restoration of the liver of the white rat following partial surgical removal. Res, 2006;66(3):1481–90.
Arch Pathol, 1931;12:186–202. 28. Luo, Y., Yang, C., Lu, W. et al. Metabolic regulator betaKlotho interacts with
4. Michalopoulos, G.K. and DeFrances, M.C. Liver regeneration. Science, fibroblast growth factor receptor 4 (FGFR4) to induce apoptosis and inhibit
1997;276(5309):60–6. tumor Cell, proliferation. J Biol Chem, 2010;285(39):30069–78.
5. Michalopoulos, G.K. Liver regeneration. J Cell Physiol, 2007;213(2):286–300. 29. Padrissa‐Altes, S., Bachofner, M., Bogorad, R.L. et al. Control of hepatocyte
6. Block, G.D., Locker, J., Bowen, W.C. et  al. Population expansion, clonal proliferation and survival by Fgf receptors is essential for liver regeneration
growth, and specific differentiation patterns in primary cultures of hepato- in mice. Gut, 2015;64(9):1444–53.
cytes induced by HGF/SF, EGF and TGF alpha in a chemically defined 30. Cicione, C., Degirolamo, C., and Moschetta, A. Emerging role of fibroblast
(HGM) medium. J Cell Biol, 1996;132(6):1133–49. growth factors 15/19 and 21 as metabolic integrators in the liver. Hepatology,
7. Zarnegar, R. and Michalopoulos, G. Purification and biological characteriza- 2012;56(6):2404–11.
tion of human hepatopoietin A., a polypeptide growth factor for hepatocytes. 31. Kong, B., Huang, J., Zhu, Y. et  al. Fibroblast growth factor 15 deficiency
Cancer Res, 1989;49(12):3314–20. impairs liver regeneration in mice. Am J Physiol Gastrointest Liver Physiol,
8. Nakamura, T., Nishizawa, T., Hagiya, M. et  al. Molecular cloning and 2014;306(10):G893–902.
expression of human hepatocyte growth factor. Nature, 1989;342(6248): 32. Yang, C., Lu, W., Lin, T. et al. Activation of Liver FGF21 in hepatocarcino-
440–3. genesis and during hepatic stress. BMC Gastroenterol, 2013;13:67.
9. Liu, Y., Michalopoulos, G.K., and Zarnegar, R. Molecular cloning and char- 33. Schumacher, J.D. and Guo, G.L. Regulation of hepatic stellate cells and
acterization of cDNA encoding mouse hepatocyte growth factor. Biochim fibrogenesis by fibroblast growth factors. Biomed Res Int, 2016;2016:
Biophys Acta, 1993;1216(2):299–303. 8323747.
10. Mars, W.M., Liu, M.L., Kitson, R.P., Goldfarb, R.H., Gabauer, M.K., and 34. Jirtle, R.L., Carr, B.I., and Scott, C.D. Modulation of insulin‐like growth
Michalopoulos, G.K. Immediate early detection of urokinase receptor after factor‐II/mannose 6‐phosphate receptors and transforming growth factor‐
partial hepatectomy and its implications for initiation of liver regeneration. beta 1 during liver regeneration (published erratum appears in J Biol Chem,
Hepatology, 1995;21(6):1695–701. 1991;266(36):24860). J Biol Chem, 1991;266(3):22444–50.
11. Schuppan, D., Schmid, M., Somasundaram, R. et  al. Collagens in the 35. Michalopoulos, G.K. Principles of liver regeneration and growth homeostasis.
liver  extraular matrix bind hepatocyte growth factor. Gastroenterology,. Compr Physiol, 2013;3(1):485–513.
1998;114(1):139–52. 36. Jakowlew, S.B., Mead, J.E., Danielpour, D., Wu, J., Roberts, A.B., and
12. Passino, M.A., Adams, R.A., Sikorski, S.L., and Akassoglou, K. Regulation Fausto, N. Transforming growth factor‐beta (TGF‐beta) isoforms in rat liver
of hepatic stellate cell differentiation by the neurotrophin receptor p75NTR. regeneration: messenger RNA expression and activation of latent TGF‐ beta.
Science, 2007;315(5820):1853–6. Cell Regul, 1991;2(7):535–48.
45:  Liver Regeneration 581

37. Chari, R.S., Price, D.T., Sue, S.R., Meyers, W.C., and Jirtle, R.L. Down‐ 62. Matondo, R.B., Punt, C., Homberg, J. et  al. Deletion of the serotonin
regulation of transforming growth factor beta receptor type I., I.I., and III transporter in rats disturbs serotonin homeostasis without impairing liver
during liver regeneration. Am J Surg, 1995;169(1):126–31. regeneration. Am J Physiol Gastrointest Liver Physiol, 2009;296(4):
38. Houck, K.A., Cruise, J.L., and Michalopoulos, G. Norepinephrine modulates G963–8.
the growth‐inhibitory effect of transforming growth factor‐beta in primary 63. Furrer, K., Rickenbacher, A., Tian, Y. et al. Serotonin reverts age‐related cap-
rat hepatocyte cultures. J Cell Physiol, 1988;135(3):551–5. illarization and failure of regeneration in the liver through a VEGF‐depend-
39. Michalopoulos, G.K., Bowen, W.C., Mule, K., and Stolz, D.B. Histological ent pathway. Proc Natl Acad Sci USA, 2011;108(7):2945–50.
organization in hepatocyte organoid cultures. Am J Pathol, 2001;159(5): 64. Ochoa, B., Syn, W.K., Delgado, I. et al. Hedgehog signaling is critical for
1877–87. normal liver regeneration after partial hepatectomy in mice. Hepatology,
40. Rudolph, K.L., Trautwein, C., Kubicka, S. et al. Differential regulation of 2010;51(5):1712–23.
extracellular matrix synthesis during liver regeneration after partial hepatec- 65. Bhave, V.S., Mars, W., Donthamsetty, S. et al. Regulation of liver growth by
tomy in rats. Hepatology, 1999;30(5):1159–66. glypican 3, CD81, hedgehog, and Hhex. Am J Pathol, 2013;183(1):153–9.
41. Thenappan, A., Shukla, V., Abdul Khalek, F.J. et  al. Loss of transforming 66. Sicklick, J.K., Li, Y.X., Choi, S.S. et  al. Role for hedgehog signaling in
growth factor beta adaptor protein beta‐2 spectrin leads to delayed liver hepatic stellate cell activation and viability. Lab Invest, 2005;85(1):1368–80.
regeneration in mice. Hepatology, 2011;53(5):1641–50. 67. Swiderska‐Syn, M., Xie, G., Michelotti, G.A. et al. Hedgehog regulates yes‐
42. Ichikawa, T., Zhang, Y.Q., Kogure, K. et al. Transforming growth factor beta associated protein 1 in regenerating mouse liver. Hepatology, 2016;64(1):
and activin tonically inhibit DNA synthesis in the rat liver. Hepatology, 232–44.
2001;34(5):918–25. 68. Francavilla, A., Starzl, T.E., Porter, K. et al. Screening for candidate hepatic
43. Cressman, D.E., Greenbaum, L.E., DeAngelis, R.A. et al. Liver failure and growth factors by selective portal infusion after canine Eck’s fistula.
defective hepatocyte regeneration in interleukin‐6‐deficient mice. Science, Hepatology, 1991;14(4 Pt 1):665–70.
1996;274(5291):1379–83. 69. Zerrad‐Saadi, A., Lambert‐Blot, M., Mitchell, C. et al. GH receptor plays a
44. Sakamoto, T., Liu, Z., Murase, N. et al. Mitosis and apoptosis in the liver major role in liver regeneration through the control of EGFR and ERK1/2
of  interleukin‐6‐deficient mice after partial hepatectomy. Hepatology, activation. Endocrinology, 2011;152(7):2731–41.
1999;29(2):403–11. 70. Jin, J., Wang, G.L., Shi, X., Darlington, G.J.,and Timchenko, N.A. The age‐
45. Runge, D.M., Runge, D., Foth, H., Strom, S.C., and Michalopoulos, G.K. associated decline of glycogen synthase kinase 3beta plays a critical role in
STAT 1alpha/1beta, STAT 3 and STAT 5: expression and association with the inhibition of liver regeneration. Mol Cell Biol, 2009;29(14):3867–80.
c‐ MET and EGF‐receptor in long‐term cultures of human hepatocytes. 71. Tackett, B.C., Sun, H., Mei, Y. et al. P2Y2 purinergic receptor activation is
Biochem Biophys Res Commun, 1999;265(2):376–81. essential for efficient hepatocyte proliferation in response to partial hepatec-
46. Norris, C.A., He, M., Kang, L.I. et al. Synthesis of IL‐6 by hepatocytes is a tomy. Am J Physiol Gastrointest Liver Physiol, 2014;307(1):G1073–87.
normal response to common hepatic stimuli. PLoS One, 2014;9(4):e96053. 72. Graubardt, N., Fahrner, R., Trochsler, M, et al. Promotion of liver regeneration
47. Fausto, N. Involvement of the innate immune system in liver regeneration by natural killer cells in a murine model is dependent on extracellular adeno-
and injury. J Hepatol, 2006;45(3):347–9. sine triphosphate phosphohydrolysis. Hepatology, 2013;57(5):1969–79.
48. Chaisson, M.L., Brooling, J.T., Ladiges, W., Tsai, S., and Fausto, N. 73. Moolten, F.L. and Bucher, N.L. Regeneration of rat liver: transfer of humoral
Hepatocyte‐specific inhibition of NF‐kappaB leads to apoptosis after TNF agent by cross circulation. Science, 1967;158(798):272–4.
treatment, but not after partial hepatectomy. J Clin Invest, 2002;110(2): 74. Marubashi, S., Sakon, M., Nagano, H. et al. Effect of portal hemodynamics
193–202. on liver regeneration studied in a novel portohepatic shunt rat model.
49. Yamada, Y., Webber, E.M., Kirillova, I., Peschon, J.J., and Fausto, N. Surgery, 2004;136(5):1028–37.
Analysis of liver regeneration in mice lacking type 1 or type 2 tumor necrosis 75. Sokabe, T., Yamamoto, K., Ohura, N. et al. Differential regulation of uroki-
factor receptor: requirement for type 1 but not type 2 receptor [comment]. nase‐type plasminogen activator expression by fluid shear stress in human
Hepatology, 1998;28(4):959–70. coronary artery endothelial cells. Am J Physiol, 2004;287(5):H2027–34.
50. Argast, G.M., Campbell, J.S., Brooling, J.T., and Fausto, N. Epidermal 76. Mars, W.M., Kim, T.H., Stolz, D.B., Liu, M.L., and Michalopoulos, G.K.
growth factor receptor transactivation mediates tumor necrosis factor‐ Presence of urokinase in serum‐free primary rat hepatocyte cultures and
induced hepatocyte replication. J Biol Chem, 2004;279(3):34530–6. its  role in activating hepatocyte growth factor. Cancer Res, 1996;56(12):
51. Desbarats, J. and Newell, M.K. Fas engagement accelerates liver regeneration 2837–43.
after partial hepatectomy. Nat Med, 2000;6(8):920–3. 77. Nejak‐Bowen, K., Moghe, A., Cornuet, P., Preziosi, M., Nagarajan, S., and
52. Oben, J.A., Roskams, T., Yang, S. et  al. Hepatic fibrogenesis requires Monga, S.P. Role and regulation of p65/beta‐catenin association during liver
­sympathetic neurotransmitters. Gut, 2004;53(3):438–45. injury and regeneration: a "complex" relationship. Gene Expr, 2017;17(3):
53. Cruise, J.L., Houck, K.A., and Michalopoulos, G.K. Induction of DNA 219–35.
­synthesis in cultured rat hepatocytes through stimulation of alpha 1 adreno- 78. Mars, W.M., Zarnegar, R., and Michalopoulos, G.K. Activation of hepato-
receptor by norepinephrine. Science, 1985;227(4688):749–51. cyte growth factor by the plasminogen activators uPA and tPA. Am J Pathol,
54. Han, C., Bowen, W.C., Michalopoulos, G.K., and Wu, T. Alpha‐1 adrenergic 1993;143(3):949–58.
receptor transactivates signal transducer and activator of transcription‐3 79. Mohammed, F.F., Pennington, C.J., Kassiri, Z. et  al. Metalloproteinase
(Stat3) through activation of Src and epidermal growth factor receptor inhibitor TIMP‐1 affects hepatocyte cell cycle via HGF activation in murine
(EGFR) in hepatocytes. J Cell Physiol, 2008;216(2):486–97. liver regeneration. Hepatology, 2005;41(4):857–67.
55. Cruise, J.L., Knechtle, S.J., Bollinger, R.R., Kuhn, C., and Michalopoulos, 80. Kim, T.H., Bowen, W.C., Stolz, D.B., Runge, D., Mars, W.M., and
G. Alpha 1‐adrenergic effects and liver regeneration. Hepatology, 1987;7(6): Michalopoulos, G.K. Differential expression and distribution of focal adhe-
1189–94. sion and cell adhesion molecules in rat hepatocyte differentiation. Exp Cell
56. Broten, J., Michalopoulos, G., Petersen, B., and Cruise, J. Adrenergic Res, 1998;244(1):93–104.
­stimulation of hepatocyte growth factor expression. Biochem Biophys Res 81. Lindroos, P.M., Zarnegar, R., and Michalopoulos, G.K. Hepatocyte growth
Commun, 1999;262(1):76–9. factor (hepatopoietin A) rapidly increases in plasma before DNA synthesis
57. Yanagita, K., Nagaike, M., Ishibashi, H., Niho, Y., Matsumoto, K., and and liver regeneration stimulated by partial hepatectomy and carbon tetra-
Nakamura, T. Lung may have an endocrine function producing hepatocyte chloride administration. Hepatology, 1991;13(4):743–50.
growth factor in response to injury of distal organs. Biochem Biophys Res 82. Michalopoulos, G.K. Liver regeneration after partial hepatectomy: critical
Commun, 1992;182(2):802–9. analysis of mechanistic dilemmas. Am J Pathol, 2010;176(1):2–13.
58. Huang, W., Ma, K. Zhang, J. et al. Nuclear receptor‐dependent bile acid sign- 83. Michalopoulos, G.K. Advances in liver regeneration. Exp Rev Gastroenterol
aling is required for normal liver regeneration. Science, 2006;312:233–6. Hepatol, 2014:1–11.
59. Borude, P., Edwards, G., Walesky, C. et al. Hepatocyte‐specific deletion of 84. Nejak‐Bowen, K., Orr, A., Bowen, W.C., Jr., and Michalopoulos, G.K.
farnesoid X receptor delays but does not inhibit liver regeneration after par- Conditional genetic elimination of hepatocyte growth factor in mice compro-
tial hepatectomy in mice. Hepatology, 2012;56(6):2344–52. mises liver regeneration after partial hepatectomy. PLoS One, 2013;8(3):
60. Keitel, V. and Haussinger, D. TGR5 in the biliary tree. Dig Dis, e59836.
2011;29(1):45–7. 85. Hojilla, C.V., Mohammed, F.F., and Khokha, R. Matrix metalloproteinases
61. Lesurtel, M., Graf, R., Aleil, B. et  al. Platelet‐derived serotonin mediates and their tissue inhibitors direct Cell, fate during cancer development. Br J
liver regeneration. Science, 2006;312:104–7. Cancer, 2003;89(10):1817–21.
582 THE LIVER:  REFERENCES

86. Gkretsi, V., Bowen, W.C., Yang, Y., Wu, C., and Michalopoulos, G.K. 108. Mullany, L.K., White, P., Hanse, E.A. et al. Distinct proliferative and tran-
Integrin‐linked kinase is involved in matrix‐induced hepatocyte differentia- scriptional effects of the D‐type cyclins in vivo. Cell Cycle, 2008;7(14):
tion. Biochem Biophys Res Commun, 2007;353(3):638–43. 2215–24.
87. Gkretsi, V., Apte, U., Mars, W.M. et al. Liver‐specific ablation of integrin‐ 109. Ng, R., Song, G., Roll, G.R., Frandsen, N.M., and Willenbring, H. A
linked kinase in mice results in abnormal histology, enhanced cell prolifera- microRNA‐21 surge facilitates rapid cyclin D1 translation and Cell,
tion, and hepatomegaly. Hepatology, 2008;48(6):1932–41. cycle progression in mouse liver regeneration. J Clin Invest, 2012;122(3):
88. Apte, U., Gkretsi, V., Bowen, W.C. et al. Enhanced liver regeneration fol- 1097–108.
lowing changes induced by hepatocyte‐specific genetic ablation of integrin‐ 110. Albrecht, J.H., Poon, R.Y., Ahonen, C.L., Rieland, B.M., Deng, C., and
linked kinase. Hepatology, 2009;50(3):844–51. Crary, G.S. Involvement of p21 and p27 in the regulation of CDK activity
89. Monga, S.P., Pediaditakis, P., Mule, K., Stolz, D.B., and Michalopoulos, and Cell, cycle progression in the regenerating liver. Oncogene,
G.K. Changes in WNT/beta‐catenin pathway during regulated growth in rat 1998;16(16):2141–50.
liver regeneration. Hepatology, 2001;33(5):1098–109. 111. Fausto, N. Protooncogenes and growth factors associated with normal and
90. Kohler, C., Bell, A.W., Bowen, W.C., Monga, S.P., Fleig, W., and abnormal liver growth. Dig Dis Sci, 1991;36(5):653–8.
Michalopoulos, G.K. Expression of Notch‐1 and its ligand Jagged‐1 in rat 112. Rabes, H.M. Kinetics of hepatocellular proliferation as a function of the
liver during liver regeneration. Hepatology, 2004;39(4):1056–65. microvascular structure and functional state of the liver. Ciba Found Symp,
91. Zhu, N.L., Asahina, K., Wang, J. et al. Hepatic stellate cell‐derived delta‐ 1977(5):31–53.
like  homolog 1 (DLK1) protein in liver regeneration. J Biol Chem, 113. Klochendler, A., Weinberg‐Corem, N., Moran, M. et al. A transgenic mouse
2012;287(13):10355–67. marking live replicating Cell,s reveals in vivo transcriptional program of
92. Jirtle, R.L. and Michalopoulos, G. Effects of partial hepatectomy on proliferation. Dev Cell Cell, 2012;23(4):681–90.
­transplanted hepatocytes. Cancer Res, 1982;42(8):3000–4. 114. Jiao, H., Zhu, Y., Lu, S., Zheng, Y., and Chen, H. An integrated approach
93. Weymann, A., Hartman, E., Gazit, V. et  al. p21 is required for dextrose‐ for  the identification of HNF4alpha‐centered transcriptional regulatory
mediated inhibition of mouse liver regeneration. Hepatology, 2009;50(1): ­networks during early liver regeneration. Cell Physiol Biochem, 2015;36(6):
207–15. 2317–26.
94. Shido, K., Chavez, D., Cao, Z., Ko, J., Rafii, S., and Ding, B.S. Platelets 115. Michalopoulos, G., Cianciulli, H.D., Novotny, A.R., Kligerman, A.D.,
prime hematopoietic and vascular niche to drive angiocrine‐mediated liver Strom, S.C., and Jirtle, R.L. Liver regeneration studies with rat hepatocytes
regeneration. Signal Transduct Target Ther, 2017;2. in primary culture. Cancer Res, 1982;42(1):4673–82.
95. Stocker, E. and Heine, W.D. Regeneration of liver parenchyma under nor- 116. Shteyer, E., Liao, Y., Muglia, L.J., Hruz, P.W., and Rudnick, D.A. Disruption
mal and pathological conditions. Beitr Pathol, 1971;144(4):400–8. of hepatic adipogenesis is associated with impaired liver regeneration in
96. Stocker, E., Wullstein, H.K., and Brau, G. Capacity of regeneration in mice (see comment). Hepatology, 2004;40(6):1322–32.
liver  epithelia of juvenile, repeated partially hepatectomized rats. 117. Fernandez‐Rojo, M.A., Restall, C., Ferguson, C. et al. Caveolin‐1 orches-
Autoradiographic studies after continous infusion of 3H‐thymidine trates the balance between glucose and lipid‐dependent energy metabolism:
(author’s transl). Virchows Arch B Cell Pathol, 1973;14(2):93–103. implications for liver regeneration. Hepatology, 2012;55(5):1574–84.
97. Kelley‐Loughnane, N., Sabla, G.E., Ley‐Ebert, C., Aronow, B.J., and 118. Shu, J., Kren, B.T., Xia, Z. et al. Genomewide microRNA down‐regulation
Bezerra, J.A. Independent and overlapping transcriptional activation during as a negative feedback mechanism in the early phases of liver regeneration.
liver development and regeneration in mice. Hepatology, 2002;35(3): Hepatology, 2011;54(2):609–19.
525–34. 119. Grijalva, J.L., Huizenga, M., Mueller, K. et  al. Dynamic alterations in
98. Taub, R., Greenbaum, L.E., and Peng, Y. Transcriptional regulatory signals Hippo signaling pathway and YAP activation during liver regeneration. Am
define cytokine‐dependent and ‐independent pathways in liver regenera- J Physiol Gastrointest Liver Physiol, 2014;307(2):G196–204.
tion. Semin Liver Dis, 1999;19(2):117–27. 120. Yimlamai, D., Christodoulou, C., Galli, G.G. et al. Hippo pathway activity
99. Yamamoto, T., Kojima, T., Murata, M. et  al. p38 MAP‐kinase regulates influences liver cell fate. Cell, 2014;157(6):1324–38.
function of gap and tight junctions during regeneration of rat hepatocytes. J 121. Toshima, T., Shirabe, K., Fukuhara, T. et al. Suppression of autophagy dur-
Hepatol, 2005;42(5):707–18. ing liver regeneration impairs energy charge and hepatocyte senescence in
100. Stolz, D.B., Mars, W.M., Petersen, B.E., Kim, T.H., and Michalopoulos, mice. Hepatology, 2014;60(1):290–300.
G.K. Growth factor signal transduction immediately after two‐thirds partial 122. Hoehme, S., Brulport, M., Bauer, A. et al. Prediction and validation of cell
hepatectomy in the rat. Cancer Res, 1999;59(16):3954–60. alignment along microvessels as order principle to restore tissue architec-
101. Li, W., Liang, X., Kellendonk, C., Poli, V.,and Taub, R. STAT3 contributes ture in liver regeneration. Proc Nat Acad Sci USA, 2010;107(23):10371–6.
to the mitogenic response of hepatocytes during liver regeneration. J Biol 123. Miyaoka, Y. and Miyajima, A. To divide or not to divide: revisiting liver
Chem, 2002;277(32):28411–7. regeneration. Cell Div, 2013;8(1):8.
102. Chaisson, M.L., Brooling, J.T., Ladiges, W., Tsai, S., and Fausto, N. 124. Meier, M., Knudsen, A.R., Andersen, K.J., Bjerregaard, N.C., Jensen, U.B.,
Hepatocyte‐specific inhibition of NF‐kappaB leads to apoptosis after TNF and Mortensen, F.V. Gene expression in the liver remnant is significantly
treatment, but not after partial hepatectomy. J Clin Invest, 2002;110(2): affected by the size of partial hepatectomy: an experimental rat study. Gene
193–202. Expr, 2017;17(4):289–99.
103. Greenbaum, L.E., Cressman, D.E., Haber, B.A., and Taub, R. Coexistence 125. Pean, N., Doignon, I., Garcin, I. et al. The receptor TGR5 protects the liver
of C/EBP alpha, beta, growth‐induced proteins and DNA synthesis in hepat- from bile acid overload during liver regeneration in mice. Hepatology,
ocytes during liver regeneration. Implications for maintenance of the differ- 2013;58(4):1451–60.
entiated state during liver growth. J Clin Invest, 1995;96(3):1351–65. 126. Glaser, S., Han, Y., Francis, H., and Alpini, G. Melatonin regulation of bil-
104. Luedde, T., Duderstadt, M., Streetz, K.L. et al. C/EBP beta isoforms LIP iary functions. Hepatobiliary Surg Nutr, 2014;3(1):35–43.
and LAP modulate progression of the cell cycle in the regenerating mouse 127. Kennedy, L., Hargrove, L., Demieville, J. et al. Blocking H1/H2 histamine
liver. Hepatology, 2004;40(2):356–65. receptors inhibits damage/fibrosis in Mdr2(‐/‐) mice and human cholangio-
105. Wang, X., Quail, E., Hung, N.J., Tan, Y., Ye, H., and Costa, R.H. Increased carcinoma tumorigenesis. Hepatology, 2018.
levels of forkhead box M1B transcription factor in transgenic mouse hepat- 128. Grappone, C., Pinzani, M., Parola, M. et al. Expression of platelet‐derived
ocytes prevent age‐related proliferation defects in regenerating liver. Proc growth factor in newly formed cholangiocytes during experimental biliary
Natl Acad Sci USA, 2001;98(20):11468–73. fibrosis in rats. J Hepatol, 1999;31(1):100–9.
106. Wang, X., Bhattacharyya, D., Dennewitz, M.B. et  al. Rapid hepatocyte 129. Li, B., Dorrell, C., Canaday, P.S. et al. Adult mouse liver contains two dis-
nuclear translocation of the Forkhead Box M1B (FoxM1B) transcription tinct populations of cholangiocytes. Stem Cell Reports, 2017;9(2):478–89.
factor caused a transient increase in size of regenerating transgenic hepato- 130. Ross, M.A., Sander, C.M., Kleeb, T.B., Watkins, S.C., and Stolz, D.B.
cytes. Gene Expr, 2003;11(3–4):149–62. Spatiotemporal expression of angiogenesis growth factor receptors during the
107. Bhave, V.S., Paranjpe, S., Bowen, W.C. et al. Genes inducing iPS pheno- revascularization of regenerating rat liver. Hepatology, 2001;34(6):1135–48.
type play a role in hepatocyte survival and proliferation in vitro and liver 131. Modis, L. and Martinez‐Hernandez, A. Hepatocytes modulate the hepatic
regeneration in vivo. Hepatology, 2011;54(4):1360–70. microvascular phenotype. Lab Invest, 1991;65(6):661–70.
45:  Liver Regeneration 583

132. Wack, K.E., Ross, M.A., Zegarra, V., Sysko, L.R., Watkins, S.C., and Stolz, 155. Le Bail, B., Bioulac‐Sage, P., Senuita, R., Quinton, A., Saric, J., and
D.B. Sinusoidal ultrastructure evaluated during the revascularization of Balabaud, C. Fine structure of hepatic sinusoids and sinusoidal cells in dis-
regenerating rat liver. Hepatology, 2001;33(2):363–78. ease. J Electron Microsc Tech, 1990;14(3):257–82.
133. DeLeve, L.D., Wang, X., and Wang, L. VEGF‐sdf1 recruitment of CXCR7+ 156. Michalopoulos, G.K. and Khan, Z. Liver stem cells: experimental findings and
bone marrow progenitors of liver sinusoidal endothelial cells promotes rat implications for human liver disease. Gastroenterology, 2015;149(4):876–82.
liver regeneration. Am J Physiol Gastrointest Liver Physiol< 2016;310(9): 157. Trautwein, C., Will, M., Kubicka, S., Rakemann, T., Flemming, P., and
G739–46. Manns, M.P. 2‐acetaminofluorene blocks cell cycle progression after hepa-
134. Ding, B.S., Nolan, D.J., Butler, J.M. et  al. Inductive angiocrine signals tectomy by p21 induction and lack of cyclin E expression. Oncogene,
from  sinusoidal endothelium are required for liver regeneration. Nature, 1999;18(47):6443–53.
2010;468(7321):310–5. 158. Tatematsu, M., Ho, R.H., Kaku, T., Ekem, J.K., and Farber, E. Studies on
135. Roskams, T., Cassiman, D., De Vos, R., and Libbrecht, L. Neuroregulation the proliferation and fate of oval cells in the liver of rats treated with 2‐
of the neuroendocrine compartment of the liver. Anat Rec A Discov Mol acetylaminofluorene and partial hepatectomy. Am J Pathol, 1984;114(3):
Cell Cell Evol Biol, 2004;280(1):910–23. 418–30.
136. Unanue, E.R. Ito cells, stellate cells, and myofibroblasts: new actors in anti- 159. Evarts, R.P., Nagy, P., Nakatsukasa, H., Marsden, E., and Thorgeirsson,
gen presentation. Immunity, 2007;26(1):9–10. S.S. In vivo differentiation of rat liver oval cells into hepatocytes. Cancer
137. Gandhi, C.R. Hepatic stellate cell activation and pro‐fibrogenic signals. Res, 1989;49(6):1541–7.
J Hepatol, 2017;67(5):1104–5. 160. Bisgaard, H.C., Nagy, P., Santoni‐Rugiu, E., and Thorgeirsson, S.S.
138. Zimmermann, A. Liver regeneration: the emergence of new pathways. Med Proliferation, apoptosis, and induction of hepatic transcription factors are
Sci Monit, 2002;8(3):RA53–63. characteristics of the early response of biliary epithelial (oval) cells to
139. Widmann, J.J. and Fahimi, H.D. Proliferation of mononuclear phagocytes chemical carcinogens. Hepatology, 1996;23(1):62–70.
(Kupffer cells) and endothelial cells in regenerating rat liver. A light and 161. Petersen, B.E., Zajac, V.F., and Michalopoulos, G.K. Bile ductular damage
electron microscopic cytochemical study. Am J Pathol, 1975;80(3): induced by methylene dianiline inhibits oval cell activation. J Pathol,
349–66. 1997;151(4):905–9.
140. Fujii, H., Hirose, T., Oe, S. et al. Contribution of bone marrow cells to liver 162. Lu, W.Y., Bird, T.G., Boulter, L. et al. Hepatic progenitor cells of biliary
regeneration after partial hepatectomy in mice. J Hepatol, 2002;36(5):653–9. origin with liver repopulation capacity. Nat Cell Biol, 2015.
141. Ikarashi, M., Nakashima, H., Kinoshita, M. et  al. Distinct development 163. Michalopoulos, G.K., Barua, L., and Bowen, W.C. Transdifferentiation of
and  functions of resident and recruited liver Kupffer cells/macrophages. rat hepatocytes into biliary cells after bile duct ligation and toxic biliary
J Leukoc Biol, 2013;94(6):1325–36. injury. Hepatology, 2005;41(3):535–44.
142. Sanderson, N., Factor, V., Nagy, P. et  al. Hepatic expression of mature 164. Limaye, P.B., Bowen, W.C., Orr, A.V., Luo, J., Tseng, G.C., and
­transforming growth factor beta 1 in transgenic mice results in multiple Michalopoulos, G.K. Mechanisms of hepatocyte growth factor‐mediated
tissue lesions. Proc Nat Acad Sci USA, 1995;92(7):2572–6. and epidermal growth factor‐mediated signaling in transdifferentiation of
143. Oe, S., Lemmer, E.R., Conner, E.A. et al. Intact signaling by transforming rat hepatocytes to biliary epithelium. Hepatology, 2008;47(5):1702–13.
growth factor beta is not required for termination of liver regeneration in 165. Sekiya, S., and Suzuki, A. Hepatocytes, rather than cholangiocytes, can be
mice. Hepatology, 2004;40(5):1098–105. the major source of primitive ductules in the chronically injured mouse
144. Kim, T.H., Mars, W.M., Stolz, D.B., and Michalopoulos, G.K. Expression liver. Am J Pathol, 2014;184(5):1468–78.
and activation of pro‐MMP‐2 and pro‐MMP‐9 during rat liver regeneration. 166. Schaub, J.R., Huppert, K.A., Kurial, S.N.T. et al. De novo formation of the
Hepatology, 2000;31(1):75–82. biliary system by TGFbeta‐mediated hepatocyte transdifferentiation.
145. Kim, T.H., Mars, W.M., Stolz, D.B., Petersen, B.E., and Michalopoulos, Nature, 2018;557(7704):247–51.
G.K. Extracellular matrix remodeling at the early stages of liver regenera- 167. Sparks, E.E., Huppert, K.A., Brown, M.A., Washington, M.K., and Huppert,
tion in the rat. Hepatology, 1997;26(4):896–904. S.S. Notch signaling regulates formation of the three‐dimensional architec-
146. Luo, J.H., Ren, B., Keryanov, S. et al. Transcriptomic and genomic analysis ture of intrahepatic bile ducts in mice. Hepatology, 2010;51(4):1391–400.
of human hepatocellular carcinomas and hepatoblastomas. Hepatology, 168. Isse, K., Lesniak, A., Grama, K. et  al. Preexisting epithelial diversity in
2006;44(4):1012–24. normal human livers: a tissue‐tethered cytometric analysis in portal/peri-
147. Pilia, G., Hughes‐Benzie, R.M., MacKenzie, A. et al. Mutations in GPC3, a portal epithelial cells. Hepatology, 2013;57(4):1632–43.
glypican gene, cause the Simpson‐Golabi‐Behmel overgrowth syndrome. 169. Petersen, B.E., Zajac, V.F., and Michalopoulos, G.K. Hepatic oval cell acti-
Nat Genet, 1996;12(3):241–7. vation in response to injury following chemically induced periportal or
148. Liu, B., Bell, A.W., Paranjpe, S. et al. Suppression of liver regeneration and pericentral damage in rats. Hepatology, 1998;27(4):1030–8.
hepatocyte proliferation in hepatocyte‐targeted glypican 3 transgenic mice. 170. Carpentier, R., Suner, R.E., van Hul, N. et al. Embryonic ductal plate cells
Hepatology, 2010;52(3):1060–7. give rise to cholangiocytes, periportal hepatocytes, and adult liver progeni-
149. Xue, Y., Mars, W.M., Bowen, W., Singhi, A.D., Stoops, J., and tor cells. Gastroenterology,. 2011;141(4):1432–8, 8 e1–4.
Michalopoulos, G.K. Hepatitis C virus mimics effects of Glypican‐3 on 171. Michalopoulos, G.K., Bowen, W.C., Mule, K., and Luo, J. HGF‐, EGF‐,
CD81 and promotes development of hepatocellular carcinomas via activa- and dexamethasone‐induced gene expression patterns during formation of
tion of hippo pathway in hepatocytes. Am J Pathol, 2018;188(6):1469–77. tissue in hepatic organoid cultures. Gene Expr, 2003;11(2):55–75.
150. Koral, K., Paranjpe, S., Bowen, W.C., Mars, W., Luo, J., and Michalopoulos, 172. Hattoum, A., Rubin, E., Orr, A., and Michalopoulos, G.K. Expression of
G.K. Leukocyte‐specific protein 1: a novel regulator of hepatocellular pro- hepatocyte epidermal growth factor receptor, FAS and glypican 3 in
liferation and migration deleted in human hepatocellular carcinoma. EpCAM‐positive regenerative clusters of hepatocytes, cholangiocytes, and
Hepatology, 2015;61(2):537–47. progenitor cells in human liver failure. Hum Pathol, 2013;44(5):743–9.
151. Yang, J., Cusimano, A., Monga, J.K. et al. WNT5A inhibits hepatocyte pro- 173. Jaeschke, H. and Bajt, M.L. Intracellular signaling mechanisms of acetami-
liferation and concludes beta‐catenin signaling in liver regeneration. Am J nophen‐induced liver cell death. Toxicol Sci, 2006;89(1):31–41.
Pathol, 2015;185(8):2194–205. 174. Slater, T.F., Cheeseman, K.H., and Ingold, K.U. Carbon tetrachloride toxic-
152. Tsagianni, A., Mars, W.M., Bhushan, B. et al. Combined systemic disrup- ity as a model for studying free‐radical mediated liver injury. Phil Trans R
tion of MET and epidermal growth factor receptor signaling causes liver Soc Lond B Biol Sci, 1985;311(1152):633–45.
failure in normal mice. Am J Pathol, 2018;188(10):2223–35. 175. Bhushan, B., Walesky, C., Manley, M. et al. Pro‐regenerative signaling after
153. Arakaki, N., Kawakami, S., Nakamura, O. et al. Evidence for the presence acetaminophen‐induced acute liver injury in mice identified using a novel
of an inactive precursor of human hepatocyte growth factor in plasma and incremental dose model. Am J Pathol, 2014;184(1):3013–25.
sera of patients with liver diseases. Hepatology, 1995;22(6):1728–34. 176. Bhushan, B., Chavan, H., Borude, P. et al. Dual role of epidermal growth
154. Pediaditakis, P., Lopez‐Talavera, J.C., Petersen, B., Monga, S.P., and factor receptor in liver injury and regeneration after acetaminophen over-
Michalopoulos, G.K. The processing and utilization of hepatocyte growth dose in mice. Toxicol Sci, 2017;155(2):363–78.
factor/scatter factor following partial hepatectomy in the rat. Hepatology, 177. Webber, E.M., FitzGerald, M.J., Brown, P.I., Bartlett, M.H., and Fausto, N.
2001;34(4 Pt 1):688–93. Transforming growth factor‐alpha expression during liver regeneration
584 THE LIVER:  REFERENCES

after partial hepatectomy and toxic injury, and potential interactions 184. Lin, S., Nascimento, E.M., Gajera, C.R. et  al. Distributed hepatocytes
between transforming growth factor‐alpha and hepatocyte growth factor. expressing telomerase repopulate the liver in homeostasis and injury.
Hepatology, 1993;18(6):1422–31. Nature, 2018;556(7700):244–8.
178. Lamers, W.H. The streaming liver: can the age of a hepatocyte be 185. Prassopoulos, P., Cavouras, D., and Gourtsoyiannis, N. Pre‐ and post‐
determined from its position on the portohepatic radius? Hepatology,
­ nephrectomy kidney enlargement in patients with contralateral renal
1990;12(2):372–4. ­cancer. Eur Urol, 1993;24(1):58–61.
179. Kennedy, S., Rettinger, S., Flye, M.W., and Ponder, K.P. Experiments in 186. Duncan, A.W., Taylor, M.H., Hickey, R.D. et al. The ploidy conveyor of mature
transgenic mice show that hepatocytes are the source for postnatal liver hepatocytes as a source of genetic variation. Nature, 2010;467(7316):707–10.
growth and do not stream. Hepatology, 1995;22(1):160–8. 187. Anti, M., Marra, G., Rapaccini, G.L. et al. DNA ploidy pattern in human
180. Wang, B., Zhao, L., Fish, M., Logan, C.Y., and Nusse, R. Self‐renewing chronic liver diseases and hepatic nodular lesions. Flow cytometric analysis
diploid Axin2(+) cells fuel homeostatic renewal of the liver. Nature, on echo‐guided needle liver biopsy. Cancer, 1994;73(2):281–8.
2015;524(7564):180–5. 188. Duncan, A.W., Hanlon Newell, A.E., Smith, L. et al. Frequent aneuploidy
181. Furuyama, K., Kawaguchi, Y., Akiyama, H. et al. Continuous cell supply among normal human hepatocytes. Gastroenterology, 2012;142(1):25–8.
from a Sox9‐expressing progenitor zone in adult liver, exocrine pancreas 189. Solt, D.B., Medline, A.,and Farber, E. Rapid emergence of carcinogen‐
and intestine. Nat Genet, 2011;43(1):34–41. induced hyperplastic lesions in a new model for the sequential analysis of
182. Font‐Burgada, J., Shalapour, S., Ramaswamy, S. et  al. Hybrid periportal liver carcinogenesis. Am J Pathol, 1977;88(3):595–618.
hepatocytes regenerate the injured liver without giving rise to cancer. Cell, 190. Zhang, Y.Y., Zhang, B.H., Ishii, K., and Liang, T.J. Novel function of CD81
2015;162(4):766–79. in controlling hepatitis C virus replication. J Virol, 2010;84(7):3396–407.
183. Planas‐Paz, L., Orsini, V., Boulter, L. et  al. The RSPO‐LGR4/5‐ZNRF3/ 191. Overturf, K., al‐Dhalimy, M., Ou, C.N., Finegold, M.,and Grompe, M.
RNF43 module controls liver zonation and size. Nat Cell Biol, Serial transplantation reveals the stem‐cell‐like regenerative potential of
2016;18(5):467–79. adult mouse hepatocytes. Am J Pathol, 1997;151(5):1273–80.
β‐Catenin Signaling
46 Satdarshan P.S. Monga
Pittsburgh Liver Research Center, University of Pittsburgh, School of Medicine and UPMC, Pittsburgh,
PA, USA

BACKGROUND lymphoid enhancement factor (TCF/LEF) family of transcrip-


tion factors, and their interactions were identified that were
Genetic studies in species such as Xenopus, Drosophila, and directly influenced by the Wnt signaling. Many new compo-
Caenorhabditis have lent themselves well to improve our under- nents and interactions as well as expanding list of target genes
standing of the molecular basis of human diseases. A classic are being identified. Also, research is focused on their role in
example is the identification and characterization of the Wnt/β‐ regulation of this pathway in health and disease. Furthermore,
catenin pathway that is crucial in normal development including crosstalk has now been established between the Wnt pathway
embryogenesis and organogenesis, and at the same time its and other prominent pathways such as the Hippo, Hedgehog,
deregulation is implicated in disorders such as cancers (reviewed Jagged/Notch, HGF, EGF, and TGFβ signaling that could have
in [1–3]). This pathway has remained conserved through the additional implications.
evolutionary process. In Drosophila, the role of Wnt or Wingless Presently, the role of Wnt/β‐catenin pathway is well estab-
(Wg) was initially identified in normal wing development, how- lished in vertebrates in development, tissue homeostasis, and
ever it was later recognized for multiple functions such as induc- carcinogenesis [11, 12]. β‐catenin knockout yields an embryonic
ing segment polarity and anterior–posterior patterning that were lethal phenotype in mice due to a defect in gastrulation [13].
imperative for a viable embryo [4–6]. As the importance of Wnt Availability of conditional knockouts to overcome embryonic
emerged, several key components of this pathway were identi- lethality has been key to understanding a more ubiquitous role of
fied. The discovery of armadillo (or β‐catenin) added a signifi- β‐catenin and other Wnt components in the development of
cant player to this orchestra and although circumstantial many organs such as kidneys, lungs, brain, limbs, muscles, and
evidence suggesting such a relationship existed earlier, it was a skin [14–19]. The role of Wnt/β‐catenin signaling in liver patho-
few years later that β‐catenin was positively identified as a cen- biology has gained significant importance in the last 15–20
tral component of the canonical Wg pathway [4, 7–9]. These years. This chapter will highlight the role of the Wnt/β‐catenin
studies led to the emergence of a model system for cell adhesion signaling pathway and of β‐catenin independent of the canonical
and signal transduction [10]. This was also the beginning of the Wnt pathway in liver physiology and pathobiology (Table 46.1).
understanding of the Wnt/β‐catenin pathway and its role in
complex cellular processes such as cell–cell adhesion, mitogen-
esis, motogenesis and morphogenesis in the vertebrates.
The next several years focused on the discovery of various THE WNT SIGNAL TRANSDUCTION
components of this pathway that improved our understanding PATHWAY
of  the regulation of this complex signaling cascade in normal
physiology and disease. Several important players such as the The binding of an extracellular secreted glycoprotein Wnt to
Wnt receptor frizzled (Fzd), zeste‐white 3 kinase or glycogen its  cell surface receptor Fzd or alternate receptors induces
synthase kinase 3β (GSK3β), adenomatous polyposis gene ­specific downstream cascades with unique biological functions.
product (APC), axin, dishevelled (Dsh), and T cell factor/ Although the most characterized signaling downstream of the

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
586 THE LIVER:  ALTERNATE WNT SIGNALING PATHWAYS

Table 46.1  Selected Wnt/β‐catenin pathway targets in liver


Target genes Context Change Reference
Axin‐2 Liver tumors Upregulated [212]
Claudin‐2 Liver baseline/injury Downregulated [109, 112,
(β‐catenin knockout) 132]
Cyclin‐D1 Liver tumors (hepatoblastoma) Upregulated [101, 216,
Liver regeneration 217]
Cyclooxygenase‐2 Liver tumors (cell line) Upregulated [218]
Cyp1a2 Liver tumors Upregulated [219]
Cyp2e1 Liver tumors Upregulated [219]
Cyp7a1 Liver baseline/injury (β‐catenin knockout) Downregulated [112, 132,
133]
Cyp27 Liver baseline/injury (β‐catenin knockout) Downregulated [112, 132,
133]
Epidermal growth factor receptor Transgenic liver (β‐catenin overexpression) Upregulated [119]
Fibronectin Liver tumors (hepatoblastoma) Upregulated [216]
G‐protein‐coupled receptor 49 Hepatocellular cancer Upregulated [220]
(Gpr49 or Lgr5)
Glutamate transporter‐1(GLT‐1) Transgenic liver (mutant β‐catenin overexpression) Upregulated [99]
Glutamine synthetase (GS, Glul) Transgenic liver (mutant β‐catenin overexpression) Upregulated [99]
Lect2 Transgenic liver (mutant β‐catenin overexpression) Upregulated [221]
Ornithine aminotransferase Transgenic liver (mutant β‐catenin overexpression) Upregulated [99]
Regenerating iselet‐derived‐3α Liver tumors (HCC and hepatoblastoma Upregulated [222]
Regucalcin Baseline liver and liver tumors (β‐catenin transgenic and Upregulated (transgenic) [107]
knockout) Downregulated (knockout)
Tbx3 Liver tumors (hepatoblastoma) Upregulated [223]
TGFα Liver regeneration Upregulated [217]
VEGF Liver tumors (cell line) Upregulated [224]

Wnt is via the β‐catenin‐TCF‐dependent gene expression, the Wnt signaling through binding to the leucine‐rich repeat‐con-
signaling can occur through other effectors. In this, a broad taining G‐protein coupled receptor‐4 (LGR4) and LGR5 recep-
overview of the various signaling cascades activated down- tors, which in turn increase Wnt‐dependent phosphorylation of
stream of Wnt are discussed. LRP6 [30]. The recruitment of Axin to the plasma membrane
disrupts the β‐catenin destruction complex, promoting its cyto-
plasmic stabilization and nuclear translocation where it binds
and transactivates the TCF/LEF family of transcription factors
THE WNT/β‐CATENIN SIGNALING to mediate target gene expression [31]. Several target genes of
the pathway have now been identified, which are stage‐ and tis-
In a normal steady state, in the absence of a Wnt signal, the free sue‐specific, and the liver related targets will be discussed
monomeric form of β‐catenin in the cytoplasm is actively targeted throughout the text. It is important to note that several Wnt path-
for degradation by ubiquitination. In this “off” state, β‐catenin is way components such as AXIN, DKK, dFzd7, Fzd2, FRP2,
phosphorylated at serine45 (Ser45), Ser33, Ser33, and threo- WISP, βTrCP, and TCF are themselves targets, suggesting exist-
nine‐41 (Thr41) by casein kinase Iα (CKIα) and GSK3β [20, 21]. ence of several regulatory loops within this pathway.
CK and GSK3β are part of a larger multiprotein degradation com-
plex that includes axin and APC. Once phosphorylated this larger
complex enables recognition and ubiquitination of β‐catenin by
β‐transducin repeat‐containing protein (βTrCP) and its ensuing ALTERNATE WNT SIGNALING
proteosomal degradation (Figure 46.1a, left panel) [22]. PATHWAYS
Wnt proteins usually act on a cell in a paracrine or autocrine
manner. However, Wnt proteins, in order to be biologically There are 19 Wnts and 10 Frizzled receptors in the mammalian
active, need to undergo specific post‐translational modifica- genome. Understandably, not all Wnt signaling converges on
tions. Porcupine, located in the endoplasmic reticulum of a cell, β‐catenin and various Wnt‐Fzd or Wnt‐non‐Fzd interactions can
is essential for glycosylation and acylation of Wnts (Figure 46.1a, lead to activation of alternate signaling pathways that are briefly
right panel) [23]. Following acylation, Wnts become hydropho- discussed here. It should be noted that some Wnts that were clas-
bic and require a specific cargo receptor, Wntless (Wls) or sically dubbed to be non‐canonical have now been shown to acti-
Evenness Interrupted (Evi), to transport Wnts from Golgi to vate or inactivate β‐catenin based on interaction with specific
membrane for secretion (Figure 46.1a, right panel) [24]. receptors and the final outcome of a particular signaling is con-
Once secreted, Wnts bind to cell surface Fzd receptor and text‐dependent. The Wnt signaling usually activates but rarely
low‐density lipoprotein receptor‐related protein (LRP) 5/6 co‐ can inhibit β‐catenin‐TCF transactivation. An example is that of
receptors [25–27]. This triggers phosphorylation of LRP5/6, Wnt5a, which can activate or inhibit β‐catenin based on the recep-
recruitment of Dishevelled (Dvl), and recruitment of Axin tor it binds [32]. It was shown that Wnt5a activated β‐catenin
to  the  plasma membrane (Figure  46.1, right panel) [28, 29]. signaling in the presence of Fz4 while it inhibited Wnt3a‐depend-
Interestingly, the family of R‐spondin secreted proteins enhance ent β‐catenin activation in the presence of Ror2 (Figure 46.1b).
β-CATENIN SIGNALING
46:  587

(a) WNT
WNT
sFRP
WNT
WIF
LRP5/6
WNT Fzd
LRP5/6
Fzd Wls Dsh
Dkk APC
GSK3 Axin
Axin
WNT CK β-catenin
APC P Porcn
β-catenin
GSK3
CK TCF/LEF β-catenin
P
β-catenin
Target
genes
TCF/LEF

Serine 33 37 45
β-Catenin HGF EGF Fer Src (pp60)

Threonine 41 Y654 Y654 Y142 Y86 β-Catenin


Y670 Y654

(b) (c)
β-Catenin AJ
β-Catenin
WNT5a WNT5a WNT5a Actin cytoskeleton

Fzd4 Fzd2 Ror2 E-Cadherin


Tyrosine-
Occludin Phospho-
JAM-A β-Catenin
Activate Inactivate
β-catenin β-catenin Claudin-2
?
Claudin-1
TJ TCF

Figure 46.1  β‐catenin as a component of the Wnt pathway and cell–cell junctions. (a) On the left, in the absence of Wnt binding to its receptor
frizzled (Fzd) and co‐receptor LDL related protein 5/6 (LRP5/6) due to either sequestration of Wnt by soluble Frizzled related protein (sFRP), or
in the presence of Wnt inhibitory factor (WIF) or inhibition of Fzd‐LRP5/6 interactions by Dikkopf (Dkk), β‐catenin in cytoplasm is bound to its
degradation complex composed of Axin, APC, GSK3, and casein kinase (CK), which leads to its phosphorylation sequentially at Ser45 and Sr33,
Ser37 and Thr41. After phosphorylation, β‐catenin is ubiquitinated by βTrCP and degraded. On the right, Wnt is glycosylated and palmitoylated by
Porcupine (Porcn) in the endoplasmic reticulum and is bound to Wntless (Wls) in the Golgi apparatus to be transported to the membrane for secre-
tion. Wnt binds to receptor Fzd and co‐receptor LRP5/6, which recruits the β‐catenin degradation complex through Dishevelled (Dsh) to phosphor‐
LRP5/6, inactivating the degradation complex, leading to hypophosphorylation of β‐catenin, which is released and eventually translocates to the
nucleus to interact with T cell factor (TCF)/lymphoid enhancer factor (LEF) family of transcription factors to induce target genes. (b) Based on the
receptor binding, Wnt5a can activate or inactivate β‐catenin. If it binds to Fzd4, it can activate β‐catenin while if it binds to Ror2 or Fzd2, it can
inactivate β‐catenin through Wnt/calcium pathway or other mechanisms. (c) β‐Catenin is a component of adherens junctions (AJ) where it links the
cytoplasmic tail of E‐cadherin to α‐catenin and actin cytoskeleton. AJ in the liver are next to tight junctions (TJ) which are known for their barrier
function and are located close to the apical surface of hepatocytes preventing leakage of bile along the lateral surface of hepatocytes into the blood.
TJ are composed of proteins like occludin, junctional adhesion molecule‐A (JAM‐A) and claudins. Claudin‐2 is also a Wnt/β‐catenin target and may
be one basis of junctional crosstalk between the components of the AJ and TJ. Growth factors like HGF, EGF, Fer kinase, and Src kinase can phos-
phorylate β‐catenin at various tyrosine residues located at the C‐terminal of β‐catenin and disrupt β‐catenin‐E‐cadherin interactions and may also
induce β‐catenin nuclear translocation and activation.

The Wnt/Ca2+ pathway


4,5 biphosphate (PIP2) into diacylglycerol (DAG) and inositol
Wnt ligands such as Wnt5a can bind to Fzd2 or Fzd7 or the tyros- 1,4,5‐triphosphate (IP3). DAG activates protein kinase C
ine kinase‐like orphan receptor 2 (Ror2) to activate this pathway (PKC) and IP3 promotes the levels of intracellular calcium,
[33]. Binding of Wnt leads to formation of a complex between which in turn activates calcium‐calmodulin dependent kinase II
Fzd, Dishevelled, and G proteins, promoting the ­activation of (CaMKII) and calcineurin (CaN) to regulate cell migration and
phospholipase C (PLC), which cleaves phosphatidylinositol proliferation [34].
588 THE LIVER:  β-CATENIN AND PROTEIN KINASE A (PKA)

The planar cell polarity pathway well, however, this remains controversial (Figure 46.1c). The inter-
actions between β‐catenin and cadherins are regulated by tyrosine
In the PCP pathway, binding of Wnt ligands to the Ror2/Fzd/ phosphorylation at the C‐terminal of β‐catenin (reviewed in [44]).
Dishevelled complex triggers the activation of Rho‐family Phosphorylation of β‐catenin destabilizes cadherin‐β‐catenin bond,
small GTPases including RhoA and Rac to then activate the α‐catenin‐β‐catenin complex, uncouples cadherin from actin
Rho‐associated protein kinase (ROCK) and c‐Jun N‐terminal cytoskeleton, and promotes loss of intracellular adhesion [45, 46].
kinase (JNK) with an eventual impact on cell polarity and Conversely, dephosphorylating β‐catenin at tyrosine residues
migration [34, 35]. enhanced E‐cadherin, β‐catenin, and α‐catenin reassembly [47].
Following tyrosine phosphorylation of β‐catenin, its cytosolic pool
The Wnt/STOP pathway is greatly increased, as is its ability to bind to TATA‐box binding
protein (TBP) to eventually increase transcriptional activity of the
This pathway leads to Wnt‐dependent stabilization of proteins β‐catenin/TCF complex [48]. This has also been narrowed down to
[36]. GSK3β is the central mediator of this pathway as it can tyrosine residue 654. Another important ramification of tyrosine
phosphorylate many additional proteins besides β‐catenin to tar- phosphorylation of β‐catenin and dissociation of β‐catenin‐E‐cad-
get them for proteasomal degradation [37]. Wnt binding to its herin complex is that it leaves the cytoplasmic domain of E‐cadherin
co‐receptors can trigger the sequestration of GSK3β in multive- unstructured and vulnerable to degradation [42].
sicular bodies, allowing the cytoplasmic accumulation of Several factors can regulate tyrosine phosphorylation of
GSK3β‐target proteins [38]. The biological effects of this path- β‐catenin including: (i) nonreceptor kinases, Src and Fer [49,
way can range from effect on cell cycle, cell division, regulation 50]; (ii) transmembrane kinases, EGF receptor (EGFR) and Met
of cytoskeleton, cell size regulation, and DNA remodeling [39]. (HGF receptor) [51–55]; and (iii) various protein tyrosine phos-
phatases (Figure 46.1c).
Wnt/TOR signaling We reported a novel Met‐β‐catenin complex at the hepatocyte
membrane [56]. HGF induced tyrosine phosphorylation‐
In this pathway GSK3β phosphorylates and activates tuberous dependent nuclear translocation of β‐catenin. In a follow‐up
sclerosis complex 2 (TSC2), which in turn inhibits the function study, we identified tyrosine residues 654 and 670 as targets of
of mTOR complex 1 (mTORC1). The presence of Wnt ligands HGF‐induced β‐catenin phosphorylation (Figure  46.1c) [57].
inhibits GSK3β‐mediated phosphorylation and degradation of Other reports had identified a similar effect of HGF on posi-
TSC2, which activates the mTORC1 signaling pathway and tively regulating β‐catenin/TCF transactivation, albeit via other
stimulates protein translation [40]. mechanisms [58–60]. These observations are relevant as high
levels of HGF have been observed in patients with liver patholo-
gies that might be influencing β‐catenin redistribution and alter-
ing the disease course [61, 62].
WNT INDEPENDENT ROLES Similarly, direct interactions of β‐catenin with EGFR have
OF β‐CATENIN been reported as well (Figure  46.1c) [52]. In fact, ErbB2 has
also been shown to be associated with β‐catenin [53]. Although
Apart from playing a central role in the canonical Wnt pathway, the fate of β‐catenin and nuclear signaling in this context is
β‐catenin can interact with a number of proteins at membrane unclear its effect on cell–cell adhesion is well defined [63].
and in the cytoplasm or nuclei to modulate their signaling and
biological function. A complete discussion on such interactions
is out of the scope of the current chapter but some key interac-
tions relevant in liver pathophysiology and pathobiology are β‐CATENIN AND PROTEIN
briefly discussed here. KINASE A (PKA)
β‐catenin can also be phosphorylated by cAMP/PKA in vitro and
in vivo at Ser552 and Ser675, which leads to increased β‐catenin‐
β‐CATENIN AT ADHERENS JUNCTIONS TCF activation independent of any input from GSK3β [64]. This
has also been shown to occur in liver pathophysiology. In evalu-
A crucial function of β‐catenin is to act as a bridge between the cyto- ating ways to activate the Wnt/β‐catenin signaling to promote
plasmic domain of the cadherins and the actin‐containing‐cytoskel- liver regeneration after hepatectomy, the effect of triiodothyro-
eton as component of the adherens junctions (AJ) [41]. Cadherins nine (T3) or another thyroid hormone receptor‐β agonist GC‐1
consist of an extracellular domain, a transmembrane domain, and a (Sobetirome) was identified [65, 66]. In fact, we showed that
cytoplasmic tail that is the most conserved region among various β‐catenin activation after T3/GC‐1 was at least partially depend-
subtypes. Structurally, the cytoplasmic tails of cadherins show ent on PKA‐mediated phosphorylation on Ser552 and Ser675.
dimerization and connect to the actin‐cytoskeleton via p120, Owing to the increased levels of cAMP observed in various
β‐catenin, and α‐catenin. Specific β‐catenin‐binding sites on the genetic diseases associated with the loss of fibrocystin function
cytoplasmic domain of cadherins have been characterized [42, 43]. such as congenital hepatic fibrosis and Caroli disease, there was
The significance of regulation of β‐catenin‐cadherin interactions at increased Ser675‐β‐catenin evident in the cholangiocytes [67].
AJ is not only important in modulating cell–cell adhesion but has This contributed to the increased motility of Pkhd1‐deficient
been extended to transcriptional activation function of β‐catenin as cholangiocytes and thus the cAMP/PKA/β‐catenin/TCF axis
β-CATENIN SIGNALING
46:  589

may be playing an important result in the disease pathogenesis cultures and a comprehensive ontogenic analysis, the role of
and represent a therapeutic target [67]. β‐catenin was demonstrated in early liver development [73,
74]. Several groups have now complemented and furthered the
understanding of how extremely tight regulation of Wnt/β‐
catenin signaling is essential at multiple steps during hepatic
β‐CATENIN AND NF‐κB development [75].
Wnt signaling promotes the posterior endodermal fate and
β‐catenin and p65 subunit of NF‐κB were shown to complex in
suppresses the anterior endodermal fate during gastrulation and
breast cancer [68]. We also discovered this complex in the liver, in
early somitogenesis [76]. At the same time suppression of Wnt
hepatocytes [69]. This complex seems to be inhibitory for p65
signaling by secreted frizzled‐related protein 5 (sFRP5) has
activation in hepatocytes. In fact, hepatocyte‐specific knockouts
been shown to maintain foregut fate in the anterior endoderm
of β‐catenin were protected from LPS injury due to increased
and allows for liver development to initiate [76, 77]. However,
NF‐κB activation resulting in pro‐survival gene expression. There
once anteroposterior endoderm patterning is established, Wnt
is also an increased baseline immune cell infiltration in the liver
signaling now positively regulates liver specification [76].
in the absence of β‐catenin and the significance of this is unclear.
These highly temporal and opposite roles of Wnt/β‐catenin
Recently, we have also shown that while β‐catenin and NF‐κB are
signaling in liver formation during early development are also
both activated after partial hepatectomy, these ensue in distinct
observed in zebrafish [78]. In fact, an unbiased forward‐genetic
cellular compartments with the former occurring in the hepato-
screen in zebrafish led to the identification of wnt2bb mutants
cytes while the latter observed in the non‐parenchymal cells [70].
that have very small or no liver buds [79]. Wnt2 knockdown in
Overexpression of β‐catenin has also been shown to decrease
wnt2bb mutants blocked liver recovery and importantly resulted
NF‐κB reporter activity [71]. Cancer cells harboring β‐catenin
in no hhex expression in the liver‐forming region [80], indicat-
mutations showed reduced NF‐κB activity while enhanced
ing the essential role of Wnt signaling in liver specification.
activity was seen in cells lacking basal β‐catenin activity.
Both wnt2 [80] and wnt2bb [79] are expressed in the lateral
Coimmunoprecipitation studies also confirmed formation of a
plate mesoderm adjacent to the liver‐forming region.
complex of β‐catenin with both p65 and p50 subunits of the
Wnt/β‐catenin signaling is not only necessary but also suffi-
NF‐κB protein. Expression of NF‐κB targets like iNOS and Fas
cient for liver specification. Gain‐of‐function studies in zebrafish
also inversely correlated with β‐catenin levels in hepatocellular
showed that overexpression of Wnt2bb [80] or Wnt8a [81] in
cancer (HCC) samples. These findings are highly relevant as
entire tissues induced ectopic hepatoblast and hepatocyte for-
majority of HCCs occur in livers with chronic injury and inflam-
mation in the posterior endoderm that normally gives rise to the
mation, which may in part be due to the complex interplay of
intestine. Wnt8a overexpression in non‐hepatic‐destined region
β‐catenin and NF‐κB at least in a subset of cases.
of dorsal endoderm in zebrafish was also shown to induce
ectopic hepatoblasts [82].
A mouse model in which Wnt/β‐catenin signaling is activated
β‐CATENIN AND FXR or inactivated in the foregut endoderm after anteroposterior
endoderm patterning but prior to liver specification, will be nec-
More recently, a complex between farnesoid X receptor (FXR) essary to define the role of Wnt/β‐catenin signaling in liver
and β‐catenin in liver cells has also been shown. Very similar to specification. We have performed foxa3‐cre driven β‐catenin
its interaction with NF‐κB, β‐catenin also complexes with FXR deletion that was evident at E9.5 in mouse hepatoblasts [83].
and suppresses it. This observation was made after β‐catenin However, this did not affect the hepatoblast compartment and
conditional knockout mice were shown to be less injured after HNF4α‐positive hepatoblasts were seen unequivocally in these
bile duct ligation as compared to the control littermates [72]. It conditional knockout embryos. While this may imply that
appears that β‐catenin functions as both an inhibitor of nuclear Wnt/β‐catenin signaling is dispensable for hepatic induction in
translocation and a nuclear corepressor through formation of a mice, it is a technical challenge to abolish β‐catenin expression
physical complex with FXR. Loss of β‐catenin expedited FXR at the right time and at the right place.
nuclear localization and FXR‐retinoic X receptor (RXR)‐α β‐catenin gene and protein expression peaks at E10–14 in
association, culminating in small heterodimer protein (SHP) mouse embryonic liver and during this time β‐catenin is local-
promoter occupancy and activation in response to bile acids ized in the nucleus, cytoplasm, and membrane in different epi-
(BA) or FXR agonist. Conversely, excessive β‐catenin seques- thelial cells and coincided well with ongoing cell proliferation
ters FXR, thus inhibiting its activation. This complex could be [74]. Mouse embryonic livers from E9.5–10 stages cultured in
playing a role in pathogenesis of cholestatic injury and may also the presence of a β‐catenin antisense oligonucleotides, showed
have therapeutic implications in various hepatic diseases. decreased proliferation and a simultaneous increase in apopto-
sis, two processes vital to hepatic morphogenesis that follow
hepatic specification and induction [73]. This correlated well
with a subsequent study that found overexpression of β‐catenin
WNT/β‐CATENIN SIGNALING IN LIVER in developing chicken livers leads to a threefold increase in liver
DEVELOPMENT size, which is due at least in part to an expanded hepatoblast
population [84].
In 2003, the first study identified the role of Wnt/β‐catenin Role of β‐catenin in bile duct morphogenesis during develop-
signaling in liver development [73]. Utilizing in vitro organ ment is also intriguing. Antisense against β‐catenin in
590 THE LIVER:  ROLE OF WNT/β-CATENIN AND CELL JUNCTIONS IN BLOOD BILE BARRIER

embryonic liver cultures led to decreased bile duct differentia- vein. While hepatocytes that are organized radially flanking
tion and the addition of Wnt3a to the embryonic liver cultures sinusoids along this porto‐central axis look grossly similar,
induced a biliary phenotype [85]. These observations were sup- these are destined to perform differing functions. This charac-
ported by in vivo studies that showed suboptimal biliary differ- teristic within a hepatic lobule is called metabolic zonation,
entiation in β‐catenin conditional null livers [83]. Conversely, which has been known for several decades [97]. Wnt/β‐catenin
enhanced biliary differentiation of hepatoblasts was observed in signaling has now been shown to be a key regulator of the gene
APC‐null livers that showed increased β‐catenin during prenatal expression in the pericentral or zone three hepatocytes [98]. In
development, and led to an untimely demise of the embryos fact, β‐catenin activation controls expression of several genes
[86]. More recent analysis shows β‐catenin to be dispensable for involved in glutamine metabolism and xenobiotic metabolism
differentiation of cholangiocytes to hepatocytes, although including glutamine synthetase (GS), ornithine aminotrans-
β‐catenin needs to be kept in tight control during development ferase, and the glutamate transporter GLT‐1 [99–101].
[87]. A cell nonautonomous role of Wnt/β‐catenin signaling in The overall basis of pericentral Wnt/β‐catenin signaling is a
biliary morphogenesis has also been reported whereby hepato- culmination of several factors. APC, the protein responsible for
cyte‐specific suppression of Wnt/β‐catenin signaling reduced β‐catenin degradation is expressed at highest levels in zones one
Notch activity in biliary cells by reducing expression of jag1ab and two and absent in zone three. Wnt2 and Wnt9b have now
and jag2b in hepatocytes [88]. Wnt5a was shown to inhibit bile been shown to be basally secreted from endothelial cells lining
duct differentiation of hepatoblasts [89]. Wnt5a deletion from the central veins [102] to activate β‐catenin in pericentral hepat-
mesenchymal cells in mid‐gestational liver led to increased ocytes through as yet unknown Fzd receptor but requiring Wnt
expression of Sox9 and the numbers of hepatocyte nuclear fac- co‐receptors LRP5/6 (Figure 46.2). This has been shown in vari-
tor (HNF)‐1β‐positive cholangiocyte precursors. In an in vitro ous genetic knockout mice. Mice lacking β‐catenin in hepato-
differentiation assay this effect was CaMKII‐dependent. The cytes [100, 101] or Wnt co‐receptors in hepatocytes [103], or
authors did not address the state of β‐catenin signaling in this lacking Wntless in endothelial cells lining central veins prevent-
study. Wnt5a treatment can activate NF‐AT to induce β‐catenin ing all Wnt secretion from these cells [102, 104, 105], all lack
degradation and thus may regulate biliary differentiation par- Wnt/β‐catenin targets in zone three hepatocytes. Similarly, the
tially through such a mechanism [90]. R‐Spondin/Lgr4/5 axis was shown to optimize Wnt signaling
The role of β‐catenin in hepatocyte maturation during devel- and also contribute to activation of β‐catenin in zone three
opment is important. Antisense‐mediated β‐catenin knockdown hepatocytes [106].
in embryonic liver cultures led to persistent expression of stem Nuclear regulation of β‐catenin‐TCF4 complex during zona-
cell markers in hepatocytes [73]. Lack of β‐catenin in hepato- tion is also relevant to discuss. It was shown that in the presence
blasts resulted in dramatic decreases in nuclear enriched tran- of β‐catenin, TCF4 is able to bind to Wnt‐responsive elements
scription factors CEBPα and HNF4α, impairing hepatocyte on target gene promoters, while in its absence TCF4 binds
maturation and fetal viability [83]. HNF4‐α responsive elements. This switch back and forth of
How is β‐catenin spatio‐temporally regulated during devel- TCF4 may eventually be key to regulation of pericentral versus
opment? Wnt2bb appears as the upstream effector of β‐catenin periportal gene expression but needs to be further ascertained.
during the earliest phases of hepatic morphogenesis [79]. Wnt9a Several known and new targets of β‐catenin/TCF4 signaling in
expression was reported in endothelial and stellate cells of the the liver revealed in the study include Axin2, GS, constitutive
embryonic sinusoidal wall in the developing liver [91]. This androstane receptor, CYP1A2, aryl hydrocarbon receptor,
report also showed presence of Frizzled 4, 7, and 9 on hepato- Mrp2, glutathione S‐transferases, and Cyp27A1. Other intrigu-
cytes. Wnt5a was shown to be expressed in mesenchymal cells ing targets of β‐catenin reported in the liver include regucalcin
during hepatic development although its effect on β‐catenin was or senescent marker protein‐30 and 4‐gulonolactonase, both of
not directly addressed [89]. Based on interaction of Met and β‐ which are essential for ascorbic acid biosynthesis in murine
catenin and the role of HGF/Met in liver development, HGF/ liver [107]. Lect2 is also an important target of β‐catenin with
Met/β‐catenin signaling may be relevant [56, 92–94]. Expression relevance in liver tumors and inflammation.
of FGF‐10 in the mouse liver correlates with peak β‐catenin
activation; moreover, release of FGF‐10 from stellate cells stim-
ulates β‐catenin expression in hepatoblasts [95]. FGF‐2, FGF‐4,
and FGF‐8 can impact β‐catenin activation in embryonic liver ROLE OF WNT/β‐CATENIN AND CELL
cultures [96]. JUNCTIONS IN BLOOD BILE BARRIER
Liver is a highly vascular organ. Likewise, it also is the largest
gland, and among other factors, one of its main functions is the
ROLE OF WNT/β‐CATENIN secretion of bile. Blood and bile are kept segregated at a micro-
IN METABOLIC ZONATION scopic level owing to the function of tight junctions (TJ) in
hepatocytes which allow bile to be secreted apically into biliary
Liver occupies a strategic location in the body and receives canaliculi moving from the pericentral to periportal zone even-
nutrient‐ and toxin‐enriched blood through portal circulation. tually to be carried into bile ductules, while blood traverses the
Histologically, this enters the hepatic lobule through the portal sinusoids at the basolateral surface of the hepatocytes.
vein located in the portal triad and mixes with blood from the Disruption of this barrier has been reported in a subset of cases
hepatic artery in sinusoids to eventually travel to the central with progressive familial intrahepatic cholestasis (PFIC) due to
β-CATENIN SIGNALING
46:  591

Figure 46.2  Cell–molecule circuitry of the β‐catenin activation in a baseline liver and after partial hepatectomy. At baseline, endothelial cells
lining the central vein constitutively release Wnt2 and Wnt9b, which act in a paracrine manner on proximal hepatocytes to activate β‐catenin‐TCF4
dependent expression of genes such as glutamine synthetase (GS), cytochrome P450 2e1 (Cyp2e1), and Cyp1a2. After partial hepatectomy, the
increased shear stress caused by sinusoidal blood flow on sinusoidal endothelial cells (red double sided arrows) likely triggers release of Wnt2 and
Wnt9b from these cells (1) to act on Fzd‐LRP5/6 receptors on hepatocytes in proximity (2) to induce β‐catenin stabilization, nuclear translocation,
and binding to TCF4 (3). This leads to increased transcription of Ccnd1 as early at 12 hours (12h) after PH (4) and can last up to 72h. The protein
increase of cyclin‐D1 is observed at 24 h which in turn starts to promote G1 to S phase transition of hepatocytes which peaks at 40h (5). Around
the same time, hepatocytes begin to express increased levels of Wnt5a (6) which is then released and through autocrine signaling is able to inactivate
β‐catenin signaling between 40–72h and thus contribute to termination of this mitogenic signal to eventually help in cessation of liver regeneration
when normal liver size is realized after partial hepatectomy.

loss‐of‐function mutation in the gene encoding for the TJ pro- less than 30 days after birth, the survivors progressed to severe
tein‐2 or ZO‐2 [108]. The role of β‐catenin as part of the AJ or fibrosis and HCC and succumbed by two and a half months of
of AJ in general in the barrier function within the liver is not age. The loss of the two catenins led to two major types of
well understood. molecular perturbations. Due to loss of β‐catenin and its role
Liver‐specific β‐catenin conditional knockout mice also within the Wnt signaling pathway, there was reduced expression
showed increase in γ‐catenin protein levels, which associated of TJ proteins such as claudin‐2 which is a known target of the
with E‐cadherin in lieu of β‐catenin to maintain cell–cell junc- Wnt pathway [111]. However, such loss was also evident in
tions [109]. A follow‐up study showed that while not optimal, β‐catenin single knockout which did not have this phenotype of
γ‐catenin compensation at AJ via its interaction with E‐cadherin the dual knockouts. Loss of γ‐catenin on top of β‐catenin loss,
did prevent E‐cadherin degradation from ubiquitin proteasome, we posit, caused a second hit deregulating junctional integrity
but E‐cadherin levels were still overall decreased in β‐catenin and decompensation. β‐catenin is known to interact with E‐
knockout livers despite γ‐catenin upregulation [110]. Further, cadherin and through this association, it masks the PEST domain
increased serine/threonine phosphorylation of γ‐catenin after in E‐cadherin, thus preventing its degradation [42]. β‐catenin loss
β‐catenin knockdown was identified and at least partially PKA‐ in the liver was compensated by increased γ‐catenin which
dependent [110]. Despite γ‐catenin increase, all pericentral bound to E‐cadherin and hence prevented its degradation
Wnt/β‐catenin targets in the liver such as GS, cyp2e1, and although not as effectively as β‐catenin as some decrease in
cyp1a2 were absent in the β‐catenin knockouts, showing lack of E‐cadherin was evident upon β‐catenin loss despite increased
compensation of γ‐catenin in the Wnt pathway [109, 110]. γ‐catenin [109, 110]. However, this prevented a major decom-
To conclusively address the functional relevance of γ‐catenin pensation and mice were able to survive and showed only mini-
at the AJ in the absence of β‐catenin, β‐ and γ‐catenin were mal increases in hepatic bile acids, serum bilirubin, and
dually deleted in both hepatocytes and cholangiocytes using reduction in bile flow due especially with age [72, 112]. Dual
Albumin‐Cre. This led to a dramatic phenotype consisting of loss of catenins in the liver, led to a notable E‐cadherin loss and
intrahepatic cholestasis, cholemia, biliary fibrosis, failure to in addition also destabilized occludin, another TJ protein, lead-
thrive, and mortality. While 80% of double knockouts died in ing to its loss as well [113]. Thus, dual loss of catenins, led to
592 THE LIVER:  ROLE IN LIVER REGENERATION AFTER PARTIAL HEPATECTOMY AND TOXICANT‐INDUCED INJURY

loss of claudin‐2 and occludin, disrupting TJ and leading to dis- (PH). The earliest study done in a rat model of PH showed tran-
ruption of the blood bile barrier and a phenotype reminiscent sient stabilization of β‐catenin protein due to post‐translational
of PFIC. While dual loss of catenins has not been reported thus modification followed by its nuclear translocation within min-
far in PFIC cases, there remains a subset of these cases whose utes after surgery [122]. Interestingly, the overall increase was
pathogenesis remains unknown [114]. Further, alterations in transient due to activation of β‐catenin degradation, β‐catenin
catenins may be a more global phenomenon contributing to persisted in the nuclei of the hepatocytes until around 24 hours.
cholestatic injury and cholemia in various other hepatic patholo- Knockdown of β‐catenin at the time of PH in rats using phospho
gies and more in depth study remains to be done. morpholino antisense oligonucleotides led to decreased hepato-
cyte proliferation and recovery of liver mass [123].
In mice, β‐catenin nuclear translocation has been observed as
early as 3–6 hours after PH [70]. Liver‐specific β‐catenin knock-
ROLE IN POSTNATAL LIVER GROWTH out mice when subjected to PH showed notable delay in peak
hepatocyte proliferation, which occurred at 72 hours instead of
The liver continues to grow during neonatal stages. In fact, in 40 hours [101, 124]. This occurred chiefly due to decreased cyc-
mice an early postnatal hepatic growth spurt occurs during the lin‐D1 along with other cyclins like A and E, thereby reducing
first month after birth. Wnt/β‐catenin signaling was identified to the number of hepatocytes in S‐phase in the knockout by at least
be active during these stages and correlated well with ongoing 50% at 40 hours after PH. While redundancy among signaling
hepatocyte proliferation through regulation of cyclin‐D1 expres- pathways ensuring successful LR after PH has been appreci-
sion [115]. Several additional models have been employed that ated, what drives it in the absence of β‐catenin remains unknown
demonstrate a positive role of β‐catenin in liver growth. A trans- [125]. To address what regulated β‐catenin activation during
genic mouse overexpressing a stable mutant of β‐catenin gener- LR, we utilized liver‐specific knockouts of Wnt co‐receptors
ated under the transcriptional control of calbindin‐D9K LRP5/6. Dual loss of these redundant receptors from hepato-
(CaBP9K) promoter and liver‐specific enhancer of the aldolase cytes led to a delay in LR after PH very similar to β‐catenin
B gene displayed three to four times larger livers due to increased knockouts [103]. This was associated with lack of β‐catenin
cell proliferation [116]. Interestingly they did not detect any activation and decreased cyclin‐D1 expression in hepatocytes
changes in any of the conventional target genes of the pathway and hence a decrease in the number of hepatocytes in S‐phase at
such as c‐myc and cyclin‐D1. Subsequent analysis of transgenic 40 hours after PH. LR improved by 72–96 hours in these ani-
livers and subtractive hybridization led to identification genes mals. Thus, β‐catenin is under the control of Wnt signaling dur-
involved in glutamine metabolism as targets of the Wnt/β‐ ing LR. To determine which cells in the liver may be the source
catenin pathway. However, no signs of neoplastic transforma- of Wnts during LR, we conditionally knocked out Wntless from
tion were reported in these animals, although APC‐conditional hepatocytes and cholangiocytes, macrophages, and endothelial
null mice that exhibits β‐catenin stabilization leads to develop- cells. Only loss of Wnt secretion capability through deletion of
ment of robust HCCs [117]. Other transgenic mice have shown Wntless from endothelial cells phenocopied β‐catenin knock-
similar hepatic growth advantage [118, 119] and also enabled outs and LRP5/6‐double knockouts [105]. Deletion from hepat-
identification of new targets such as the epidermal growth factor ocytes and cholangiocytes had no effect on peak hepatocyte
receptor [119]. Transgenic mice expressing Ser45‐mutant β‐ proliferation whereas deletion from macrophages had a modest
catenin under an albumin promoter also showed an initial decrease suggesting contribution from these cells as secondary
increase in liver size but then adapted via enhanced association source of Wnts during LR [103, 105]. Intriguingly, when
of β‐catenin to E‐cadherin at the membrane [120]. It was only endothelial cells were isolated from regenerating livers at 12
after stimulation of liver growth via hepatectomy or chemical hours after PH, the same two Wnts (Wnt2 and Wnt9b) that are
carcinogenesis that a notable growth advantage of active mutant essential for pericentral β‐catenin activation at baseline, were
of β‐catenin was visible as compared to controls. many‐fold upregulated in the endothelial cells as well [105].
Two independent groups including ours have reported condi- The most upstream effectors that are responsible for basal Wnt2
tional hepatic loss of β‐catenin affecting liver size [100, 101]. and Wnt9b expression in central vein endothelial cells versus
The decreased liver weight : body weight ratio was chiefly due induced Wnt2 and Wnt9b expression in sinusoidal endothelial
to reduced postnatal hepatocyte proliferation that led to cells after PH remain unknown. We have recently speculated
decreased hepatic mass which was sustained throughout the ani- that increased shear stress on sinusoidal endothelial cells imme-
mal’s lifespan. The major mechanism of reduced proliferation diately after PH may be upstream of increased expression and
appears to be lower cyclin‐D1 expression, a known target of release of Wnt2 and Wnt9b. Indeed, when primary or immortal-
β‐catenin in various tissues such as liver and colon [101, 121]. ized sinusoidal endothelial cells are subjected to orbital shear
stress, there was increased expression of several Wnts including
Wnt2 and Wnt9b [105].
Eventually, prolonged hepatocyte proliferation was observed
ROLE IN LIVER REGENERATION AFTER after PH when Wntless was deleted from hepatocytes. This led
PARTIAL HEPATECTOMY AND to discovery of Wnt5a as a termination signal for the Wnt/
TOXICANT‐INDUCED INJURY β‐catenin pathway which is secreted by hepatocytes when
β‐catenin activation is no longer needed [126].
The Wnt/β‐catenin pathway has been comprehensively studied Thus, the Wnt2/9b‐LRP5/6‐β‐catenin axis is the key regula-
in liver regeneration (LR) that ensues after partial hepatectomy tor of both the metabolic zonation to control expression of
β-CATENIN SIGNALING
46:  593

genes in zone three as well as in regulating cyclin‐D1 expres- clinical utility in various scenarios [129]. This would be relevant
sion and driving LR after PH (Figure  46.2). Additionally, in transplantation settings such as to induce regeneration in liv-
Wnt5a appears to terminate β‐catenin activation following ing donors, small‐for‐size syndrome, or generally after liver
achievement of appropriate hepatocyte proliferation and recov- transplantation. It may also be useful in promoting LR after
ery of hepatic mass. toxicant‐induced liver injury.
β‐catenin’s signaling in LR has also been shown in zebrafish.
Transgenic zebrafish lines that express dominant‐negative TCF
in the hepatocytes showed a blunted regenerative response to
hepatectomy due to decreased cell proliferation [78]. Similarly,
ROLE OF WNT/β‐CATENIN SIGNALING
zebrafish expressing APC mutant or overexpressing Wnt8a IN HEPATIC BILE ACID HOMEOSTASIS
showed enhanced LR due to increased β‐catenin [78]. AND INTRAHEPATIC CHOLESTASIS
In patients, we have shown β‐catenin activation to positively
correlate with hepatocyte proliferation albeit in patients with Hepatocytes are responsible for the conversion of cholesterol
acetaminophen overdose [127]. First, β‐catenin was shown to into bile acids. Bile acids, which are detergents and toxic to
be relevant in toxicant‐induced liver injury and repair. Sublethal hepatocytes, are secreted and eliminated into bile canaliculi for
dose of acetaminophen causes centrilobular necrosis that is fol- eventual transport to the gut lumen to assist in the digestion of
lowed by spontaneous hepatocyte proliferation in adjacent dietary lipids and cholesterol. Two of the key enzymes responsi-
zone that begins the repair process. Following administration ble for the bile acid biosynthesis, Cyp7a1 and Cyp27, are mainly
of a single intra‐peritoneal dose of 500 mg kg−1 of acetami- expressed in the pericentral hepatocytes, suggesting they may
nophen to CD1 male mice, ALT levels increased from 3–12 be regulated by Wnt/β‐catenin signaling [130].
hours and returned to normal by 24 hours [127]. PCNA‐posi- Intrahepatic cholestasis encompasses defects in bile flow in
tive hepatocytes were visible from 3–12 hours as well and the liver within the intrahepatic bile ducts and biliary canaliculi.
decreased by 24 hours at which time liver appeared almost nor- The injury is likely due to the detergent action of the retained bile
mal. β‐catenin stabilization and activation was evident at 1–6 acids within the liver that result in local injury to hepatocytes,
hours after acetaminophen injury and cyclin‐D1 increased from ensuing ductular reaction and associated inflammation and fibro-
at 3–12 hours. Liver‐specific β‐catenin knockouts lack cyp2e1 sis. It has also been reported that HCCs with CTNNB1 mutation
and cyp1a2, and hence are protected from acetaminophen over- and β‐catenin activation often display intratumoral cholestasis
dose since they can’t metabolize acetaminophen to generate the [131]. Liver‐specific β‐catenin knockout mice showed increased
reactive metabolite [100, 101]. When expression of these two susceptibility to steatohepatitis with the methionine and choline
P450’s was induced in the β‐catenin knockout mice, it enabled deficient diet [132]. These mice also showed increased basal
partial metabolism of acetaminophen and in turn led to a mild ­levels of hepatic bile acids. This is at least in part thought to be
injury. When hepatocyte proliferation was compared between due to a defect in biliary canaliculi as they showed dilatation,
controls and β‐catenin knockouts at equitoxic doses, there was tortuosity, and loss of canalicular microvilli, likely due to absence
a deficit in hepatocyte proliferation in the knockouts support- of β‐catenin targets claudin‐2 and regucalcin [112].
ing the role of β‐catenin in toxicant‐induced LR as well [127]. Initially it was thought that these increased bile acids may be
In a retrospective study in patients, β‐catenin nuclear/cytoplas- responsible for decreased expression of Cyp7a1 and Cyp27,
mic localization correlated with high PCNA and spontaneous which are involved in bile acid synthesis from cholesterol, in the
LR, while predominantly membranous β‐catenin without any β‐catenin knockouts as feedback mechanism [112, 132].
nuclear/cytoplasmic redistribution correlated with decreased However, chromatin immunoprecipitation has shown these two
proliferation index and need for liver transplantation [127]. genes to also be direct targets of the Wnt pathway and thus play-
Thus β‐catenin is also relevant in regulating hepatocyte prolif- ing a role in bile acid metabolism [133].
eration in humans. An intriguing finding was made when liver‐specific β‐catenin
To address if β‐catenin activation could promote LR, trans- knockout mice were subjected to bile duct ligation [72]. There
genic mice expressing a stable S45‐mutant‐β‐catenin were sub- was a notable decrease in hepatic injury observed which was a
jected to PH. These animals showed a more pronounced result of failure of elevation of bile acid levels in the knockouts
regenerative response with a shift to the left in regeneration after the surgery although these levels were notably upregulated
kinetics [120]. Likewise, when naked Wnt‐1 DNA was injected in control mice after the same procedure. It was shown that there
weekly followed by PH, there was increased β‐catenin activa- was overall decrease in hepatic bile acid biosynthesis due to
tion and premature hepatocyte proliferation [120]. More decreased expression of Cyp7a1 and Cyp27 which was in part
recently we have identified an important role of triiodothyro- due to excessive FXR activation. In fact, we showed that
nine (T3) or small molecule T3 agonist called GC‐1 (Sobetirome) β‐catenin functioned as both an inhibitor of nuclear transloca-
in inducing β‐catenin activation and in turn cyclin‐D1 expres- tion and a nuclear corepressor through formation of a physical
sion. Both these agents showed a positive effect on hepatocyte complex with FXR. Loss of β‐catenin promoted FXR nuclear
proliferation and promoting LR after PH [65, 66]. Interestingly, localization and FXR/RXRα association, resulting in SHP pro-
these agents did not promote pro‐tumorigenic activity of onco- moter activation in response bile acid or FXR agonists.
genic β‐catenin in cancer models, but only promoted activation Thus β‐catenin is an important regulator of bile acid metabo-
of hepatocyte proliferation in normal hepatocytes through lism and in specific cases, therapeutic inhibition of β‐catenin
β‐catenin activation [128]. Thus use of these agents is a highly especially with an FXR agonist may be beneficial specially to
relevant means to stimulate β‐catenin and could have potential decrease overall intrahepatic bile acid burden and related injury.
594 THE LIVER:  ROLE OF WNT/β-CATENIN SIGNALING IN HEPATOBLASTOMAS

ROLE IN HEPATIC FIBROSIS negative regulator of Wnt/β‐catenin signaling [147]. The authors
showed that mesoderm‐specific transcript homologue expres-
Chronic injury to the liver results in hepatic fibrosis, a wound sion in HSCs alleviated carbon tetrachloride‐induced collagen
healing response, which can lead to cirrhosis if the insult contin- deposition in liver tissue by decreasing the expression of
ues. Due to lack of effective treatment, other than alleviating the β‐catenin, α‐smooth muscle actin, and Smad3 both in vivo and
injury, hepatic fibrosis and cirrhosis remain a major unmet clini- in vitro.
cal need. Hepatic fibrosis is mostly a function of hepatic stellate Wnt‐β‐catenin signaling might lead to myofibroblastic acti-
cells (HSCs) that are the source of extracellular matrix deposi- vation of HSCs via negative regulation of adipogenesis [148].
tion within injured livers [134]. Expression of Wnt11 or β‐catenin with the S33Y mutation in
Wnt signaling is beginning to be assessed in hepatic stellate preadipocytes prevented their differentiation to adipocytes at
cell biology. The first studies that showed increased expression least in part through inhibition of expression of CCAAT
of the WNT pathway components in hepatic fibrosis were using enhancer binding protein‐α (CEBPα) and PPARγ. Conversely,
genomic analysis from primary biliary cholangitis livers [135, blockade of β‐catenin signaling either through dominant‐nega-
136]. These studies identified increased expression of Wnt13, tive TCF4 or axin, facilitated adipogenic differentiation of the
Wnt5a, β‐catenin, and others, although a cause and effect rela- preadipocytes. Quiescent HSCs contain lipid droplets and their
tionship was not addressed. A study directly investigating gene activation to myofibroblasts involves loss of adipogenic genes
expression differences in quiescent versus activated HSCs iden- including PPARG and CEBPA, so this event is analogous to dif-
tified, among others, an upregulated expression of Wnt5a and ferentiation of adipocytes to preadipocytes. Gain‐of‐function
Fz2 [137]. Subsequent studies have further identified aberrant mutations in adipogenic transcription factors such as PPARG
Wnt5a expression in fibrotic livers, and its suppression led to and sterol regulatory element binding protein‐1c can reverse
reduced HSC activation [138, 139]. Wnt5a can have opposing culture‐induced myofibroblast into a quiescent HSC phenotype.
roles on β‐catenin signaling based on receptors expressed in a Yet another study showed HSC activation to myofibroblasts by
cell. Whether the role of Wnt5a in promoting fibrosis is due to the Wnt pathway to be mediated by miR‐132 and miR‐212
inhibition, activation or independent of β‐catenin needs further which led to epigenetic repression of methyl‐CpG binding pro-
clarification. Unfortunately, there is evidence for all three in the tein 2 (MeCP2) leading to loss of PPARγ [149]. More recently,
literature at the present time. As mentioned before, one study in HSCs expression of stearoyl‐CoA desaturase expression was
showed no change in nuclear translocation of β‐catenin follow- shown to be regulated by the Wnt‐β‐catenin signaling. This
ing Wnt5a increase that was evident during stellate cell activa- enzyme generates monounsaturated fatty acids, which were
tion concomitant with upregulation of Fzd2 [137]. shown to provide a feed forward loop to augment Wnt signaling
Another study showed that activation of Wnt/β‐catenin sign- through the stabilization of Lrp5 and Lrp6 mRNAs, which
aling in primary rat HSC by an inhibitor of GSK3β decreased encode for Wnt co‐receptors. This was shown to eventually con-
synthesis of α‐smooth muscle actin and Wnt5a, and induced the tribute to liver fibrosis through HSC activation [150]. Taken
expression of glial fibrillary acidic protein [140]. This study fur- together these observations indicate that activation of Wnt sign-
ther showed that activation of Wnt signaling lowered DNA syn- aling via β‐catenin could be involved in HSC activation by
thesis and prevented HSCs from entering the cell cycle to inhibiting the adipogenic programs; inhibition of this signaling
eventually demonstrate the role of Wnt signaling to β‐catenin in pathway could contribute to adipogenic gene profile of a quies-
maintaining their quiescence. However, a growing body of lit- cent HSC. Inhibiting Wnt signaling to β‐catenin might therefore
erature supports the activation of β‐catenin by WNT signaling block hepatic fibrosis.
during the process of HSC activation and fibrosis [141–143]. In
fact, different molecules have been employed to block Wnt
signaling, directly or indirectly, to demonstrate an overall anti‐
fibrotic effect both in vitro and in vivo. ROLE OF WNT/β‐CATENIN SIGNALING
Activating pregnane X receptor by rifampicin, among other IN HEPATOBLASTOMAS
things, also inhibited Wnt signaling and reduced HSC prolifera-
tion and transdifferentiation to active myofibroblasts [144]. Hepatoblastoma (HB) is the commonest malignant hepatic
Another study shows that necdin, a melanoma antigen family tumor of childhood. These tumors are frequently sporadic; how-
protein that promotes neuronal and myogenic differentiation ever, the incidence is highest in patients of familial adenoma-
while inhibiting adipogenesis, which expressed in HSCs. tous polyposis coli [151]. This led to the identification of APC
Necdin is induced during HSC activation and its silencing mutations in HB in familial cases [152]. Increased frequency of
reversed them to quiescence through PPARγ and suppression of diverse APC mutations (57%) were reported in sporadic form of
Wnt/β‐catenin signaling [145]. Another study showed that the disease as well [153]. N‐terminal mutations (missense and
SEPT4, a subunit of the septin cytoskeleton specifically deletions) affecting exon 3 of CTNNB1 have also been found in
expressed in quiescent HSCs, is downregulated through trans- up to 80–90% of all HBs and are associated with nuclear and
differentiation to activated myofibroblasts. Loss of SEPT4 in cytoplasmic β‐catenin localization [154]. Mutations in AXIN1
HSCs coincided with decreased expression of Wnt inhibitor have been identified in less than 10% of these tumors [155]. A
DKK2, increased Wnt signaling via β‐catenin, and increased recent comprehensive study in 85 HB patients showed 65 cases
fibrosis [146]. Another study supported that concept that the with missense mutations and interstitial deletions in CTNNB1
canonical Wnt pathway promotes fibrogenesis, based on analy- [156]. They also identified loss‐of‐function mutations in APC
sis of mesoderm‐specific transcript homologue, a strong and AXIN1 genes. Thus 82% of HB in this cohort exhibited
β-CATENIN SIGNALING
46:  595

Wnt/β‐catenin activation. Thus, there is compelling data that β‐ of a portal triads and interlobular bile ducts within HCAs. While
catenin activation is an obligatory event in the etiopathogenesis initially considered as a well‐defined homogeneous entity, it is
of HB, although the exact mechanism of how it leads to HB is now known that several molecular classes of this tumor‐type
unclear [157]. exist which dictates tumor origin, behavior, and eventually
Activation of Wnt signaling during normal liver development determines prognosis and helps determine treatment options
has already been discussed in this chapter and plays role in both (reviewed in [165]). Around 10–15% display Wnt/β‐catenin
hepatoblast proliferation and hepatocyte differentiation and activation secondary to mutations in the CTNNB1 gene [166].
maturation (also reviewed in [75, 158, 159]). While β‐catenin β‐catenin‐mutated HCAs are exclusive from the HNF1A muta-
activation in liver development is spatio‐temporally regulated tion group but may occur in combination with gp130 or GNAS
and ligand‐dependent, in HB its activation is unrestricted, sus- mutations. Around half of the β‐catenin‐active HCAs are inflam-
tained, and non‐ligand‐dependent due to exon‐3 mutations. matory. GS is upregulated in β‐catenin‐mutated HCA. Since
How this contributes to tumors remains to be addressed although detection of nuclear β‐catenin by immunostaining is often chal-
various target genes of Wnt signaling such as c‐Myc, cyclin‐D1, lenging due to technical issues as well as heterogeneity in stain-
GS, EGFR/Axin‐2, and others have been reported in various ing patterns, it may not be conclusive. GS staining can aid in the
histological subtypes of HB [160]. Very similar to β‐catenin in diagnosis of β‐catenin‐mutated HCA with greater sensitivity
different stages of normal liver development, a more nuclear [167]. β‐catenin‐mutated HCA can occur in males. CTNNB1‐
β‐catenin is evident in embryonal HB coinciding with lack of mutated HCA histologically exhibit cholestasis and cellular
GS, and more membranous, cytoplasmic and nuclear with GS dysplasia. Most importantly, these subsets of HCA have greater
expression in fetal HB [161]. Intriguingly, when an N‐terminal propensity for malignant transformation [166]. Another subset
deletion mutant of β‐catenin is overexpressed in the liver using of HCA has activation of the JAK/STAT pathway and show
an adenoviral approach or by generation of a transgenic mouse polymorphic inflammatory infiltrates. Mutations in IL6ST
overexpressing β‐catenin under liver‐specific promoter, the (coding for gp130), STAT3, and GNAS, can activate JAK/STAT
mice never display HB [116, 118]. Also, transgenic mice signaling. A subset of this tumor‐type also exhibits CTNNB1
expressing point‐mutant form of β‐catenin do not show HB mutation, irrespective of the molecular driver, and poses an
[120]. This suggests that oncogenic β‐catenin is insufficient in increased risk of malignant transformation as well.
inducing HB. How β‐catenin activation leads to HB was unclear
until recently.
A recent study demonstrates β‐catenin mutations frequently
co‐occur with activation of Yes associated protein‐1 (Yap), a ROLE OF WNT/β‐CATENIN SIGNALING
component of the Hippo signaling pathway. In fact, 80% of all IN HEPATOCELLULAR CANCER
HB cases, irrespective of histological subtype, showed nuclear
β‐catenin due to mutations and nuclear Yap due to unknown Wnt/β‐catenin activation has been implicated in a major sub-
mechanism [162]. Co‐expression of deletion mutant of set of HCC cases. Abnormal localization of cadherins and
β‐catenin and constitutively active Yap by sleeping beauty trans- catenins in liver cancer was first shown by immunohisto-
poson/transposase and hydrodynamic tail vein injection (SB‐ chemistry [168]. A more comprehensive study identified
HTVI), led to development of HB in mice. The role of anomalous β‐catenin expression as well as mutations in the
Yap‐TEAD and β‐catenin‐TCF appears to be key in regulating CTNNB1 in around 25% of all HCC cases and up to 50% of
target genes that may be leading to HB development and further all hepatic tumors in transgenic lines such as c‐myc or H‐ras
studies will be critical in addressing the mechanism. C‐Myc was [169]. Several subsequent human studies corroborated these
shown to be dispensable for HB development in this model observations and currently around 8–44% of HCCs show
although it played role in optimum HB development by regulat- mutations in β‐catenin gene, mostly in exon‐3 although recent
ing tumor metabolism [163]. Likewise, lipocalin‐2, identified as studies have also identified a region in exon 7/8 of CTNNB1
one of the genes that was many‐fold upregulated in Yap‐β‐ to be a hot spot [170, 171] (Table  46.2). In fact, CTNNB1
catenin HB in mice and also contained TEAD and TCF binding mutations have a trunk role in HCC‐like mutations affecting
sites in its promoter, was also shown to be dispensable for HB TERT and p53 [172].
development [164]. However, serum lipocalin‐2 was identified Mutations have also been reported in other components of the
as a sensitive marker of HB tumor burden in the mouse model degradation complex of β‐catenin including AXIN1 in around
and may be of relevance clinically as well. Future studies would 3–16% [155, 173, 174] and AXIN2 in around 3% of all HCC
be invaluable in understanding the role of these pathways in cases  [155]. Additional mechanisms have also been described
tumor biology. and include overexpression of FRZ7 [175, 176], Wnt3 upregula-
tion [177], inactivation of GSK3β [178], methylation of soluble
frizzled related protein1 (sFRP1) [179], epigenetic inactivation
of several sFRPs [180], and TGFβ‐dependent activation of
ROLE OF WNT/β‐CATENIN SIGNALING β‐catenin [181]. In HCC patients, exome sequencing has also
IN HEPATOCELLULAR ADENOMA revealed cooperating mutations of CTNNB1 with mutations in
ARID2, NFE2L2, TERT, APOB, and MLL2. Likewise, certain
Hepatocellular adenomas (HCAs) are characterized by mono- mutations were also always mutually exclusive from CTNNB1
clonal proliferation of well‐differentiated hepatocytes that are and demonstrated lack of cooperation in development of HCC.
usually arranged in sheets and cords. There is a classical absence These included TP53 and AXIN1.
596 THE LIVER:  ROLE OF WNT/β-CATENIN SIGNALING IN HEPATOBLASTOMAS

Table 46.2  List of studies showing spectra of mutations in Ctnnb1 into account geographical factors, etiology, co‐morbidities, and
gene kind of mutations may be essential to clarify this discrepancy.
Cases with Decreased fibrosis was reported in a few studies and remains
mutations in exon‐3 an intriguing hallmark of β‐catenin mutated tumors [182, 188].
Study of CTNNB1 (%) Additional information
Are β‐catenin mutations a risk factor for development of HCC
[171] 36/194 (18.5%) Additional 15/194 cases had mutations independent of cirrhosis or are they merely decreasing the
(TCGA) and 90/249 in exon 7/8 of CTNNB1
(36%) (French Additional 3/249 cases had mutations
threshold of neoplastic transformation? It is also intriguing to
cohort) in exon 7/8 of CTNNB1 note that in a recent study, a small but significant subset of HCV
[170] 90/249 (36%) Overall, 52% HCC cases showed patients developed HCC without evidence of advanced fibrosis
activation of Wnt/β‐catenin including [189]. Along the same lines, it is relevant to note that the small
mutations in AXIN1 and exon 7/8 of
CTNNB1 subset of hepatic adenomas that progress to HCC in patients
[225] 41/125 (33%) Additional 15.2% showed mutations in often exhibit β‐catenin gene mutations [166]. This neoplastic
AXIN1 transformation of adenomas occurs in a healthy liver without
[182] 9/32 (28%) Tyrosine‐654 phosphorylated β‐catenin
was observed in fibrolamellar variety
any evidence of fibrosis and further supports the role of
of HCC β‐catenin in fibrosis‐independent HCC. To address the relation-
[174] 20/45 (44%) Additional seven patients had AXIN1 ship of advanced fibrosis, β‐catenin mutations and HCC, we
mutations used an experimental approach in which ser‐45‐mutant
[226] 15/45 (33%) No GSK3β mutations
[220] 16/38 (42%) Multiple mutations in two patients β‐catenin transgenic mice and control mice were fed thioaceta-
[155] 14/73 (19%) One insertion between S33 and G34 mide diet for a prolonged period [190]. No difference in HCC in
[227] 5/62 (8%) Aflatoxin study the two groups of mice suggests that β‐catenin mutations do not
[228] 7/60 (12%) 62% of tumors had cytoplasmic
enhance evolution of cirrhosis to HCC supporting β‐catenin
β‐catenin staining
[229] 57/434 (13%) 34 had mutations at GSK3β sites. 17 gene mutations and cirrhosis to be independent contributors to
showed mutations at codons 32 and 34 tumorigenesis that do not cooperate.
[230] 9/22 (41%) Multiple mutations in one patient The Wnt/β‐catenin pathway in HCC in experimental
[186] 12/35 (34%) Multiple mutations in two patients
[231] 21/119 (18%) Multiple mutations in a patient ­models  deserves mention. The first question that has been
[32] 9/38 (24%) Aberrant accumulation of β‐catenin in answered is if β‐catenin mutation, activation, or overexpression
nucleus, cytoplasm, and membrane by itself is sufficient for initiation of HCC. None of the trans-
was seen in 39% cases genic mice overexpressing either wild‐type or stable‐mutants
[169] 8/31 (26%) Two patients had mutations at D32
of β‐catenin thus far have exhibited spontaneous HCC [116,
Deletions usually involved one of the key sites: S33, S37, S45, or T41. 118–120]. However, several studies now suggest that β‐catenin
­collaborates with other signaling pathways to contribute to
It is presently unclear if the extent of Wnt activation due to hepatocarcinogenesis. β‐catenin was shown to cooperate with
these diverse mechanisms is comparable. At the same time, it is activated Ha‐ras in HCC [190]. Mice heterozygous for Lkb1
unknown if downstream signaling in the form of target gene deletion showed an accelerated progression to HCC when
expression is also varied in these disparate mechanisms of mated with adenovirus‐inducible β‐catenin mutant mice [192].
β‐catenin activation. A study showed significant correlation Similarly, when chemical carcinogen diethylnitrosamine
between CTNNB1 mutations and overexpression of target genes (DEN), which normally induces HCC through Ha‐ras activa-
GS, G‐protein‐coupled receptor (GPR)49, and glutamate trans- tion [193], was injected in serine‐45‐mutant β‐catenin trans-
porter (GLT)‐1 (P = 0.0001), but not for other target genes like genic mouse, these animals developed HCC earlier and
ornithine aminotransferase, LECT2, c‐myc, or cyclin D1 [174]. more  profoundly [120]. These findings strongly suggest that
This study showed GS to be a good immunohistochemical β‐catenin mutation is one of the hits that may be critical to the
marker of β‐catenin activation in HCC, which was also con- development of HCC, however, additional aberrations are nec-
firmed independently [182]. However, no increase in expression essary for tumorigenesis. This is also in agreement with clinical
of GS, GPR49, or GLT‐1 was evident in loss of Axin‐1 function studies where CTNNB1 mutations significantly coexisted with
due to AXIN1 mutations. Thus, it is likely that functional equiv- mutations in ARID2, NFE2L2, TERT, APOB, and MLL2 [170,
alence of various modes of β‐catenin activation is distinct and 194]. Since concomitant Met overexpression/activation and
may have differing consequences in tumor phenotype. Indeed mutations in CTNNB1 were found in around 11% of all HCC
β‐catenin‐active HCC due to mutations in CTNNB1, AXIN1, or cases, these two proto‐oncogenes were co‐expressed using SB‐
additional modes of β‐catenin activation, have all been shown to HTVI. These mice developed HCC with gene expression pro-
have distinct phenotype in the transcriptomic classification of files that displayed high correlation with the gene profiles of a
HCC [174, 181, 183]. Furthermore, recent studies have even subset of human HCC patients with both CTNNB1 mutations
shown disparity in extent of Wnt signaling based on differences and Met activation signatures [171]. To address if Ras activa-
in mutations within exon‐3 of CTNNB1 [184]. tion downstream of Met could be contributing to Met‐β‐catenin
The effect of β‐catenin mutations on prognosis remains HCC, G12D‐KRAS and mutant‐β‐catenin were expressed using
debatable. β‐catenin mutations have been associated with better SB‐HTVI, which also yielded HCC with around 90% molecu-
prognosis and a more differentiated tumor type in some studies lar similarity to Met‐β‐catenin HCC [195]. It will be useful to
[183, 185]. Others have noted high nuclear and cytoplasmic generate relevant models using SB‐HTVI that represent subsets
β‐catenin in more proliferating and poorly differentiated HCC of human HCC which can then be assessed for biological as
[181, 186, 187]. Eventually, more careful meta‐analysis taking well as therapeutic studies.
β-CATENIN SIGNALING
46:  597

Several proof of principle preclinical studies have demon- that showed uniform positivity for GS, also showed positivity
strated an important benefit of therapeutic inhibition of β‐ for p‐mTOR‐S2448 in mice and in patients. Based on these
catenin for treatment of HCC [196]. Various cox2 inhibitors observations, Met‐β‐catenin HCC model showed a dramatic
such as rofecoxib have shown efficacy in decreasing β‐catenin response to mTOR inhibitor rapamycin. Additionally, with the
levels along with shrinkage of tumors [197]. R‐Etodolac, an reports showing an efficacy of immune checkpoint inhibitors in
enantiomer of a cox2 inhibitor that lacks an inhibitory effect on small subsets of HCC, a recent study showed exclusion of
cox2 has also shown anti‐β‐catenin effect [198]. Another group immune cells from HCC that are mutated for β‐catenin gene,
of agents including Exisulind and analogues that are inhibitors making them unlikely candidates for these agents [208]. Thus
of cyclic GMP phosphodiesterases (PDE) have been shown to assessing β‐catenin mutation ­status to stratify cases for receiv-
activate protein kinase G (PKG) that in turn decrease β‐catenin ing specific agents or excluding them from receiving specific
levels via a novel GSK3β‐independent processing mechanism drugs could form the basis of ­personalized medicine in HCC.
[199]. ICG‐001, a small molecule known to inhibit β‐catenin’s Since β‐catenin is also present at the cell surface in associa-
interaction with CREB‐binding protein (CBP) was shown to tion with E‐cadherin, a concern remains if β‐catenin suppres-
affect β‐catenin‐TCF‐dependent target gene expression [200]. sion would lead to destabilization of the AJ and may inadvertently
Its new generation analogue PRI‐724 is in clinical trial for vari- promote tumor cell migration and metastasis. However, multi-
ous malignancies that exhibit β‐catenin activation [201]. Using ple studies have now shown that β‐catenin knockdown is com-
computational chemoinformatics to identify another small mol- pensated by increased γ‐catenin which maintains AJ through
ecule with structural similarity to ICG‐001, we identified a com- association with E‐cadherin and actin cytoskeleton [109, 110].
pound we labeled as PMED‐1 [202]. This inhibitor showed in Intriguingly, we did not identify any compensation by γ‐catenin
vitro and in vivo efficacy against β‐catenin in HCC cells and in in the Wnt signaling upon β‐catenin loss. We also failed to
zebrafish, respectively. Several of these inhibitors may have detect nuclear γ‐catenin in β‐catenin knockout mice even after
potential therapeutic utility in the correct subset of HCC PH and via TopFlash (β‐catenin‐TCF) reporter assay in vitro.
patients. While GS as a biomarker of β‐catenin mutations is Thus, the redundancy in catenins at the AJ but not in the Wnt
helpful in identifying that subset, biopsies may not be feasible in signaling is reassuring and makes β‐catenin a viable therapeutic
majority of HCC patients due to underlying liver disease and target in HCC. It will be critical to further address the mecha-
cirrhosis. Thus, there is a need for secreted biomarkers to detect nisms whereby γ‐catenin is stabilized at AJ following β‐catenin
β‐catenin mutations to select eligible patients for anti‐β‐catenin suppression [110].
therapies since β‐catenin is not a global therapeutic target in
HCC [203]. Due to this unmet need, we attempted and identi-
fied Lect2 in cell culture and in mice to be a faithful biomarker
in identifying activating β‐catenin mutations [204]. However, WNT/β‐CATENIN SIGNALING IN BILE
serum Lect2 did not correspond to β‐catenin mutations in HCC DUCT AND GALL BLADDER TUMORS
patients although levels greater than 50 ng ml−1 in patients had a
positive predictive value of 97% in detecting HCC [204]. The most common tumor that arises in the biliary tree is the
Using SB‐HTVI to co‐express mutant β‐catenin and mutant cholangiocarcinoma that can either originate from the intrahe-
Kras, we showed HCC development in these mice that resem- patic portion, intrahepatic cholangiocarcinoma (ICC), or the
bled by gene expression the Met‐β‐catenin HCC, which in turn hilum (hilar cholangiocarcinoma) (reviewed in [206]). In chol-
showed around 70% resemblance by gene expression studies to angiocarcinoma, reduced expression of β‐catenin and E‐cad-
around 11% of all human HCC [171, 195]. We used these mod- herin at the membrane is observed as compared to the
els to examine the effect of Met inhibition on HCC development surrounding non‐cancerous ducts [208]. Nuclear localization of
and showed a lack of any notable response using two different β‐catenin is seen in a subset of tumors based on histology and
modalities [128, 205]. On the other hand, use of a lipid nanopar- location of the tumor (reviewed in [210]). For most ICCs, aber-
ticle to deliver siRNA to β‐catenin showed a complete response rant nuclear localization is observed in around 15% of cases and
due to β‐catenin suppression which in turn affected cell prolif- a decrease in membranous localization is related to poorer his-
eration and survival [195]. Previous in vivo studies had also tological differentiation [211]. No mutations in CTNNB1 were
shown β‐catenin suppression albeit using locked nucleic acid identified although mutations in any other components of the
antisense in a chemical carcinogenesis induced HCC model, to Wnt pathway were not assessed in this study. Our personal
also lead to complete response [206]. We believe β‐catenin to be experience in 62 ICCs showed only 2/62 tumors with nuclear
a viable therapeutic target in HCC, once patients are carefully β‐catenin [162]. Another study detected exon 3 mutations in
selected and verified especially for β‐catenin gene mutations. 7.5% of biliary tract cancer and in 57% of gallbladder adenomas
Despite no FDA approved anti‐β‐catenin agents currently, [213]. Higher frequency of mutations was seen in ampullary and
a  recent study demonstrates the role of mTOR inhibition in gallbladder carcinomas than the bile duct cancers. A higher cor-
β‐catenin‐mutated liver tumors [207]. This study showed locali- relation of CTNBN1 mutations and papillary adenocarcinoma
zation of p‐mTOR‐S2448, marker of mTORC1 activation, in the was also observed. Intraductal papillary neoplasms also showed
same cells as GS in a normal adult liver. This localization of GS anomalous nuclear localization of β‐catenin in around 25% of
and p‐mTOR‐S2448 in zone three hepatocytes was absent in patients without any exon‐3 CTNNB1 mutation [214]. Thus,
knockouts of hepatocyte β‐catenin or Wnt co‐receptors LRP5/6. while the Wnt/β‐catenin pathway may be active in a subset of
Forced re‐expression of β‐catenin or GS led to re‐expression of biliary tract neoplasms, more studies are needed to comprehend
p‐mTOR‐S2448. More importantly, β‐catenin‐mutated HCC the mechanism of its observed deregulation.
598 THE LIVER:  REFERENCES

REFERENCES 26. Pinson, K.I., Brennan, J., Monkley, S., Avery, B.J., and Skarnes, W.C. An
LDL‐receptor‐related protein mediates Wnt signalling in mice. Nature,
2000;407(6803):535–8.
  1. Moon, R.T., Bowerman, B., Boutros, M., and Perrimon, N. The promise and
27. Tamai, K., Semenov, M., Kato, Y. et  al. LDL‐receptor‐related proteins in
perils of Wnt signaling through beta‐catenin. Science, 2002;296(5573):1644–6.
Wnt signal transduction. Nature, 2000;407(6803):530–5.
  2. Peifer, M. and Polakis, P. Wnt signaling in oncogenesis and embryogenesis –
28. Tamai, K., Zeng, X., Liu, C. et al. A mechanism for Wnt coreceptor activa-
a look outside the nucleus. Science, 2000;287(5458):1606–9.
tion. Mol Cell, 2004;13(1):149–56.
  3. Arias, A.M. Epithelial mesenchymal interactions in cancer and development.
29. MacDonald, B.T., Tamai, K., and He, X. Wnt/beta‐catenin signaling: compo-
Cell, 2001;105(4):425–31.
nents, mechanisms, and diseases. Dev Cell, 2009;17(1):9–26.
  4. Perrimon, N. and Mahowald, A.P. Multiple functions of segment polarity
30. Carmon, K.S., Gong, X., Lin, Q., Thomas, A., and Liu, Q. R‐spondins func-
genes in Drosophila. Dev Biol, 1987;119(2):587–600.
tion as ligands of the orphan receptors LGR4 and LGR5 to regulate Wnt/
  5. Babu, P. Early developmental subdivisions of the wing disk in Drosophila.
beta‐catenin signaling. Proc Natl Acad Sci USA, 2011;108(28):11452–7.
Mol Gen Genet, 1977;151(3):289–94.
31. Cadigan, K.M. and Nusse, R. Wnt signaling: a common theme in animal
  6. Sharma, R.P. and Chopra, V.L. Effect of the Wingless (wg1) mutation on
development. Genes Dev, 1997;11(24):3286–305.
wing and haltere development in Drosophila melanogaster. Dev Biol,
32. Mikels, A.J. and Nusse, R. Purified Wnt5a protein activates or inhibits beta‐
1976;48(2):461–5.
catenin‐TCF signaling depending on receptor context. PLoS Biol,
 7. Patel, N.H., Schafer, B., Goodman, C.S., and Holmgren, R. The role of
2006;4(4):e115.
­segment polarity genes during Drosophila neurogenesis. Genes Dev, 1989;
33. Oishi, I., Suzuki, H., Onishi, N. et al. The receptor tyrosine kinase Ror2 is
3(6):890–904.
involved in non‐canonical Wnt5a/JNK signalling pathway. Genes Cells,
  8. Peifer, M., Rauskolb, C., Williams, M., Riggleman, B., and Wieschaus, E.
2003;8(7):645–54.
The segment polarity gene armadillo interacts with the wingless signaling
34. Pez, F., Lopez, A., Kim, M., Wands, J.R., Caron de Fromentel, C.,and Merle,
pathway in both embryonic and adult pattern formation. Development,
P. Wnt signaling and hepatocarcinogenesis: molecular targets for the devel-
1991;111(4):1029–43.
opment of innovative anticancer drugs. J Hepatol, 2013;59(5):1107–17.
 9. Riggleman, B., Schedl, P., and Wieschaus, E. Spatial expression of the
35. Nishita, M., Enomoto, M., Yamagata, K., and Minami, Y. Cell/tissue‐tropic
Drosophila segment polarity gene armadillo is posttranscriptionally regu-
functions of Wnt5a signaling in normal and cancer cells. Trends Cell Biol,
lated by wingless. Cell, 1990;63(3):549–60.
2010;20(6):346–54.
10. Peifer, M., Orsulic, S., Pai, L.M., and Loureiro, J. A model system for
36. Acebron, S.P., Karaulanov, E., Berger, B.S., Huang, Y.L., and Niehrs, C.
cell  adhesion and signal transduction in Drosophila. Dev Suppl, 1993:
Mitotic wnt signaling promotes protein stabilization and regulates cell size.
163–76.
Mol Cell, 2014;54(4):663–74.
11. Clevers, H. and Nusse, R. Wnt/beta‐catenin signaling and disease. Cell,
37. Koch, S., Acebron, S.P., Herbst, J., Hatiboglu, G., and Niehrs, C. Post‐tran-
2012;149(6):1192–205.
scriptional Wnt signaling governs epididymal sperm maturation. Cell,
12. Nusse, R. and Clevers, H. Wnt/beta‐catenin signaling, disease, and emerging
2015;163(5):1225–36.
therapeutic modalities. Cell, 2017;169(6):985–99.
38. Vinyoles, M., Del Valle‐Perez, B., Curto, J. et  al. Multivesicular GSK3
13. Haegel, H., Larue, L., Ohsugi, M., Fedorov, L., Herrenknecht, K., and
sequestration upon Wnt signaling is controlled by p120‐catenin/cadherin
Kemler, R. Lack of beta‐catenin affects mouse development at gastrulation.
interaction with LRP5/6. Mol Cell, 2014;53(3):444–57.
Development, 1995;121(11):3529–37.
39. Acebron, S.P. and Niehrs, C. Beta‐catenin‐independent roles of Wnt/LRP6
14. Hari, L., Brault, V., Kleber, M. et al. Lineage‐specific requirements of beta‐
signaling. Trends Cell Biol, 2016;26(12):956–67.
catenin in neural crest development. J Cell Biol, 2002;159(5):867–80.
40. Inoki, K., Ouyang, H., Zhu, T. et al. TSC2 integrates Wnt and energy signals
15. Brault, V., Moore, R., Kutsch, S. et al. Inactivation of the beta‐catenin gene
via a coordinated phosphorylation by AMPK and GSK3 to regulate cell
by Wnt1‐Cre‐mediated deletion results in dramatic brain malformation and
growth. Cell, 2006;126(5):955–68.
failure of craniofacial development. Development, 2001;128(8):1253–64.
41. Knudsen, K.A., Soler, A.P., Johnson, K.R., and Wheelock, M.J. Interaction
16. Lickert, H., Kutsch, S., Kanzler, B., Tamai, Y., Taketo, M.M., and Kemler, R.
of alpha‐actinin with the cadherin/catenin cell‐cell adhesion complex via
Formation of multiple hearts in mice following deletion of beta‐catenin in
alpha‐catenin. J Cell Biol, 1995;130(1):67–77.
the embryonic endoderm. Dev Cell, 2002;3(2):171–81.
42. Huber, A.H. and Weis, W.I. The structure of the beta‐catenin/E‐cadherin
17. Mucenski, M.L., Wert, S.E., Nation, J.M. et al. beta‐Catenin is required for
complex and the molecular basis of diverse ligand recognition by beta‐
specification of proximal/distal cell fate during lung morphogenesis. J Biol
catenin. Cell, 2001;105(3):391–402.
Chem, 2003;278(41):40231–8.
43. Jou, T.S., Stewart, D.B., Stappert, J., Nelson, W.J., and Marrs, J.A. Genetic
18. Huelsken, J., Vogel, R., Erdmann, B., Cotsarelis, G., and Birchmeier, W.
and biochemical dissection of protein linkages in the cadherin‐catenin com-
beta‐Catenin controls hair follicle morphogenesis and stem cell differentia-
plex. Proc Natl Acad Sci USA, 1995;92(11):5067–71.
tion in the skin. Cell, 2001;105(4):533–45.
44. Lilien, J., Balsamo, J., Arregui, C., and Xu, G. Turn‐off, drop‐out: functional
19. Kispert, A., Vainio, S.,and McMahon, A.P. Wnt‐4 is a mesenchymal signal
state switching of cadherins. Dev Dyn, 2002;224(1):18–29.
for epithelial transformation of metanephric mesenchyme in the developing
45. Ozawa, M. and Kemler, R. Altered cell adhesion activity by pervanadate due
kidney. Development, 1998;125(21):4225–34.
to the dissociation of alpha‐catenin from the E‐cadherin.catenin complex. J
20. Behrens, J., Jerchow, B.A., Wurtele, M. et al. Functional interaction of an
Biol Chem, 1998;273(11):6166–70.
axin homolog, conductin, with beta‐catenin, APC, and GSK3beta. Science,
46. Roura, S., Miravet, S., Piedra, J., Garcia de Herreros, A., and Dunach, M.
1998;280(5363):596–9.
Regulation of E‐cadherin/catenin association by tyrosine phosphorylation. J
21. Amit, S., Hatzubai, A., Birman, Y. et al. Axin‐mediated CKI phosphorylation
Biol Chem, 1999;274(51):36734–40.
of beta‐catenin at Ser 45: a molecular switch for the Wnt pathway. Genes
47. Hu, P., O’Keefe, E.J., and Rubenstein, D.S. Tyrosine phosphorylation of human
Dev, 2002;16(9):1066–76.
keratinocyte beta‐catenin and plakoglobin reversibly regulates their binding to
22. Aberle, H., Bauer, A., Stappert, J., Kispert, A., and Kemler, R. Beta‐catenin
E‐cadherin and alpha‐catenin. J Invest Dermatol, 2001;117(5):1059–67.
is a target for the ubiquitin‐proteasome pathway. EMBO, J, 1997;
48. Piedra, J., Martinez, D., Castano, J., Miravet, S., Dunach, M., and de
16(13):3797–804.
Herreros, A.G. Regulation of beta‐catenin structure and activity by tyrosine
23. Barrott, J.J., Cash, G.M., Smith, A.P., Barrow, J.R., and Murtaugh, L.C.
phosphorylation. J Biol Chem, 2001;276(23):20436–43.
Deletion of mouse Porcn blocks Wnt ligand secretion and reveals an ectoder-
49. Rosato, R., Veltmaat, J.M., Groffen, J., and Heisterkamp, N. Involvement of
mal etiology of human focal dermal hypoplasia/Goltz syndrome. Proc Natl
the tyrosine kinase fer in cell adhesion. Mol Cell Biol, 1998;18(10):5762–70.
Acad Sci USA, 2011;108(31):12752–7.
50. Behrens, J., Vakaet, L., Friis, R. et al. Loss of epithelial differentiation and
24. Bartscherer, K., Pelte, N., Ingelfinger, D., and Boutros, M. Secretion of Wnt
gain of invasiveness correlates with tyrosine phosphorylation of the E‐cad-
ligands requires Evi, a conserved transmembrane protein. Cell,
herin/beta‐catenin complex in cells transformed with a temperature‐sensitive
2006;125(3):523–33.
v‐SRC gene. J Cell Biol, 1993;120(3):757–66.
25. Bhanot, P., Brink, M., Samos, C.H. et al. A new member of the frizzled fam-
51. Shibamoto, S., Hayakawa, M., Takeuchi, K. et al. Tyrosine phosphorylation
ily from Drosophila functions as a Wingless receptor. Nature, 1996;
of beta‐catenin and plakoglobin enhanced by hepatocyte growth factor and
382(6588):225–30.
β-CATENIN SIGNALING
46:  599

epidermal growth factor in human carcinoma cells. Cell Adhes Commun, tures: role in proliferation, apoptosis, and lineage specification.
1994;1(4):295–305. Gastroenterology, 2003;124(1):202–16.
52. Hoschuetzky, H., Aberle, H., and Kemler, R. Beta‐catenin mediates the inter- 74. Micsenyi, A., Tan, X., Sneddon, T., Luo, J.H., Michalopoulos, G.K., and
action of the cadherin‐catenin complex with epidermal growth factor recep- Monga, S.P. Beta‐catenin is temporally regulated during normal liver devel-
tor. J Cell Biol, 1994;127(5):1375–80. opment. Gastroenterology, 2004;126(4):1134–46.
53. Kanai, Y., Ochiai, A., Shibata, T. et al. c‐erbB‐2 gene product directly associ- 75. Nejak‐Bowen, K. and Monga, S.P. Wnt/beta‐catenin signaling in hepatic
ates with beta‐catenin and plakoglobin. Biochem Biophys Res Commun, organogenesis. Organogenesis, 2008;4(2):92–9.
1995;208(3):1067–72. 76. McLin, V.A., Rankin, S.A., and Zorn, A.M. Repression of Wnt/beta‐catenin
54. Takahashi, K., Suzuki, K.,and Tsukatani, Y. Induction of tyrosine phospho- signaling in the anterior endoderm is essential for liver and pancreas devel-
rylation and association of beta‐catenin with EGF receptor upon tryptic opment. Development, 2007;134(12):2207–17.
digestion of quiescent cells at confluence. Oncogene, 1997;15(1):71–8. 77. Li, Y., Rankin, S.A., Sinner, D., Kenny, A.P., Krieg, P.A., and Zorn, A.M. Sfrp5
55. Birchmeier, C., Birchmeier, W., Gherardi, E., and Vande Woude, G.F. Met, coordinates foregut specification and morphogenesis by antagonizing both canon-
metastasis, motility and more. Nat Rev Mol Cell Biol, 2003;4(12):915–25. ical and noncanonical Wnt11 signaling. Genes Dev, 2008;22(21):3050–63.
56. Monga, S.P., Mars, W.M., Pediaditakis, P. et  al. Hepatocyte growth factor 78. Goessling, W., North, T.E., Lord, A.M. et al. APC mutant zebrafish uncover
induces Wnt‐independent nuclear translocation of beta‐catenin after Met‐ a changing temporal requirement for wnt signaling in liver development. Dev
beta‐catenin dissociation in hepatocytes. Cancer Res, 2002;62(7):2064–71. Biol, 2008;320(1):161–74.
57. Zeng, G., Apte, U., Micsenyi, A., Bell, A., and Monga, S.P. Tyrosine residues 79. Ober, E.A., Verkade, H., Field, H.A., and Stainier, D.Y. Mesodermal Wnt2b sig-
654 and 670 in beta‐catenin are crucial in regulation of Met‐beta‐catenin nalling positively regulates liver specification. Nature, 2006;442(7103):688–91.
interactions. Exp Cell Res, 2006;312(18):3620–30. 80. Poulain, M. and Ober, E.A. Interplay between Wnt2 and Wnt2bb controls
58. Hiscox, S. and Jiang, W.G. Association of the HGF/SF receptor, c‐met, with multiple steps of early foregut‐derived organ development. Development,
the cell‐surface adhesion molecule, E‐cadherin, and catenins in human 2011;138(16):3557–68.
tumor cells. Biochem Biophys Res Commun, 1999;261(2):406–11. 81. Shin, D., Lee, Y., Poss, K.D., and Stainier, D.Y.R. Restriction of hepatic
59. Papkoff, J. and Aikawa, M. WNT‐1 and HGF regulate GSK3 beta activity competence by Fgf signaling. Development, 2011;138(7):1339–48.
and beta‐catenin signaling in mammary epithelial cells. Biochem Biophys 82. So, J., Martin, B.L., Kimelman, D., and Shin, D. Wnt/beta‐catenin signaling
Res Commun, 1998;247(3):851–8. cell‐autonomously converts non‐hepatic endodermal cells to a liver fate. Biol
60. Danilkovitch‐Miagkova, A., Miagkov, A., Skeel, A., Nakaigawa, N., Zbar, Open, 2013;2(1):30–6.
B., and Leonard, E.J. Oncogenic mutants of RON and MET receptor tyrosine 83. Tan, X., Yuan, Y., Zeng, G. et al. Beta‐catenin deletion in hepatoblasts dis-
kinases cause activation of the beta‐catenin pathway. Mol Cell Biol, rupts hepatic morphogenesis and survival during mouse development.
2001;21(17):5857–68. Hepatology, 2008;47(5):1667–79.
61. Shiota, G., Umeki, K., Okano, J., and Kawasaki, H. Hepatocyte growth fac- 84. Suksaweang, S., Lin, C.M., Jiang, T.X., Hughes, M.W., Widelitz, R.B., and
tor and acute phase proteins in patients with chronic liver diseases. J Med, Chuong, C.M. Morphogenesis of chicken liver: identification of localized
1995;26(5–6):295–308. growth zones and the role of beta‐catenin/Wnt in size regulation. Dev Biol,
62. Yamagami, H., Moriyama, M., Tanaka, N., and Arakawa, Y. Detection of 2004;266(1):109–22.
serum and intrahepatic human hepatocyte growth factor in patients with type 85. Hussain, S.Z., Sneddon, T., Tan, X., Micsenyi, A., Michalopoulos, G.K., and
C liver diseases. Intervirology, 2001;44(1):36–42. Monga, S.P. Wnt impacts growth and differentiation in ex vivo liver develop-
63. Bonvini, P., An, W.G., Rosolen, A. et  al. Geldanamycin abrogates ErbB2 ment. Exp Cell Res, 2004;292(1):157–69.
association with proteasome‐resistant beta‐catenin in melanoma cells, 86. Decaens, T., Godard, C., de Reynies, A. et al. Stabilization of beta‐catenin
increases beta‐catenin‐E‐cadherin association, and decreases beta‐catenin‐ affects mouse embryonic liver growth and hepatoblast fate. Hepatology,
sensitive transcription. Cancer Res, 2001;61(4):1671–7. 2008;47(1):247–58.
64. Taurin, S., Sandbo, N., Qin, Y., Browning, D., and Dulin, N.O. 87. Cordi, S., Godard, C., Saandi, T. et al. Role of beta‐catenin in development
Phosphorylation of beta‐catenin by cyclic AMP‐dependent protein kinase. J of bile ducts. Differentiation, 2016;91(1–3):42–9.
Biol Chem, 2006;281(15):9971–6. 88. So, J., Khaliq, M., Evason, K. et  al. Wnt/beta‐catenin signaling controls
65. Alvarado, T.F., Puliga, E., Preziosi, M. et al. Thyroid hormone receptor beta intrahepatic biliary network formation in zebrafish by regulating notch activ-
agonist induces beta‐catenin‐dependent hepatocyte proliferation in mice: ity. Hepatology, 2018;67(6):2352–66.
implications in hepatic regeneration. Gene Expr, 2016;17(1):19–34. 89. Kiyohashi, K., Kakinuma, S., Kamiya, A. et al. Wnt5a signaling mediates
66. Fanti, M., Singh, S., Ledda‐Columbano, G.M., Columbano, A., and Monga, biliary differentiation of fetal hepatic stem/progenitor cells in mice.
S.P. Tri‐iodothyronine induces hepatocyte proliferation by protein kinase A‐ Hepatology, 2013;57(6):2502–13.
dependent beta‐catenin activation in rodents. Hepatology, 90. Saneyoshi, T., Kume, S., Amasaki, Y., and Mikoshiba, K. The Wnt/calcium
2014;59(6):2309–20. pathway activates NF‐AT and promotes ventral cell fate in Xenopus embryos.
67. Spirli, C., Locatelli, L., Morell, C.M. et  al. Protein kinase A‐dependent Nature, 2002;417(6886):295–9.
pSer(675) ‐beta‐catenin, a novel signaling defect in a mouse model of con- 91. Matsumoto, K., Miki, R., Nakayama, M., Tatsumi, N., and Yokouchi, Y.
genital hepatic fibrosis. Hepatology, 2013;58(5):1713–23. Wnt9a secreted from the walls of hepatic sinusoids is essential for morpho-
68. Deng, J., Miller, S.A., Wang, H.Y. et  al. beta‐catenin interacts with and genesis, proliferation, and glycogen accumulation of chick hepatic epithe-
inhibits NF‐kappa B in human colon and breast cancer. Cancer Cell, lium. Dev Biol, 2008;319(2):234–47.
2002;2(4):323–34. 92. Apte, U., Zeng, G., Muller, P. et al. Activation of Wnt/beta‐catenin pathway
69. Nejak‐Bowen, K., Kikuchi, A., and Monga, S.P. Beta‐catenin‐NF‐kappaB during hepatocyte growth factor‐induced hepatomegaly in mice. Hepatology,
interactions in murine hepatocytes: a complex to die for. Hepatology, 2006;44(4):992–1002.
2013;57(2):763–74. 93. Schmidt, C., Bladt, F., Goedecke, S. et al. Scatter factor/hepatocyte growth
70. Nejak‐Bowen, K., Moghe, A., Cornuet, P., Preziosi, M., Nagarajan, S., and factor is essential for liver development. Nature, 1995;373(6516):699–702.
Monga, S.P. Role and regulation of p65/beta‐catenin association during liver 94. Uehara, Y., Minowa, O., Mori, C. et  al. Placental defect and embryonic
injury and regeneration: a “complex” relationship. Gene Expr, lethality in mice lacking hepatocyte growth factor/scatter factor. Nature,
2017;17(3):219–35. 1995;373(6516):702–5.
71. Du, Q., Zhang, X., Cardinal, J. et al. Wnt/beta‐catenin signaling regulates 95. Berg, T., Rountree, C.B., Lee, L. et al. Fibroblast growth factor 10 is critical
cytokine‐induced human inducible nitric oxide synthase expression by for liver growth during embryogenesis and controls hepatoblast survival via
inhibiting nuclear factor‐kappaB activation in cancer cells. Cancer Res, beta‐catenin activation. Hepatology, 2007;46(4):1187–97.
2009;69(9):3764–71. 96. Sekhon, S.S., Tan, X., Micsenyi, A., Bowen, W.C., and Monga, S.P.
72. Thompson, M.D., Moghe, A., Cornuet, P. et al. beta‐Catenin regulation of Fibroblast growth factor enriches the embryonic liver cultures for hepatic
farnesoid X receptor signaling and bile acid metabolism during murine chol- progenitors. Am J Pathol, 2004;164(6):2229–40.
estasis. Hepatology, 2018;67(3):955–71. 97. Gebhardt, R. and Mecke, D. Heterogeneous distribution of glutamine syn-
73. Monga, S.P., Monga, H.K., Tan, X., Mule, K., Pediaditakis, P., and thetase among rat liver parenchymal cells in situ and in primary culture.
Michalopoulos, G.K. Beta‐catenin antisense studies in embryonic liver cul- EMBO J, 1983;2(4):567–70.
600 THE LIVER:  REFERENCES

  98. Benhamouche, S., Decaens, T., Godard, C. et al. Apc tumor suppressor gene 122. Monga, S.P., Pediaditakis, P., Mule, K., Stolz, D.B., and Michalopoulos,
is the “zonation‐keeper” of mouse liver. Dev Cell, 2006;10(6):759–70. G.K. Changes in WNT/beta‐catenin pathway during regulated growth in rat
  99. Cadoret, A., Ovejero, C., Terris, B. et al. New targets of beta‐catenin signal- liver regeneration. Hepatology, 2001;33(5):1098–109.
ing in the liver are involved in the glutamine metabolism. Oncogene, 123. Sodhi, D., Micsenyi, A., Bowen, W.C., Monga, D.K., Talavera, J.C.,
2002;21(54):8293–301. and  Monga, S.P. Morpholino oligonucleotide‐triggered beta‐catenin
100. Sekine, S., Lan, B.Y., Bedolli, M., Feng, S., and Hebrok, M. Liver‐specific knockdown compromises normal liver regeneration. J Hepatol, 2005;
­
loss of beta‐catenin blocks glutamine synthesis pathway activity and 43(1):132–41.
cytochrome p450 expression in mice. Hepatology, 2006;43(4):817–25. 124. Sekine, S., Gutierrez, P.J., Lan, B.Y., Feng, S., and Hebrok, M. Liver‐­
101. Tan, X., Behari, J., Cieply, B., Michalopoulos, G.K., and Monga, S.P. specific loss of beta‐catenin results in delayed hepatocyte proliferation after
Conditional deletion of beta‐catenin reveals its role in liver growth and partial hepatectomy. Hepatology, 2007;45(2):361–8.
regeneration. Gastroenterology, 2006;131(5):1561–72. 125. Michalopoulos, G.K. Principles of liver regeneration and growth homeosta-
102. Wang, B., Zhao, L., Fish, M., Logan, C.Y., and Nusse, R. Self‐renewing sis. Compr Physiol, 2013;3(1):485–513.
diploid Axin2(+) cells fuel homeostatic renewal of the liver. Nature, 126. Yang, J., Cusimano, A., Monga, J.K. et al. WNT5A inhibits hepatocyte pro-
2015;524(7564):180–5. liferation and concludes beta‐catenin signaling in liver regeneration. Am J
103. Yang, J., Mowry, L.E., Nejak‐Bowen, K.N. et al. Beta‐catenin signaling in Pathol, 2015;185(8):2194–205.
murine liver zonation and regeneration: a Wnt‐Wnt situation! Hepatology, 127. Apte, U., Singh, S., Zeng, G. et al. Beta‐catenin activation promotes liver
2014;60(3):964–76. regeneration after acetaminophen‐induced injury. Am J Pathol, 2009;
104. Leibing, T., Geraud, C., Augustin, I. et al. Angiocrine Wnt signaling con- 175(3):1056–65.
trols liver growth and metabolic maturation in mice. Hepatology, 128. Puliga, E., Min, Q., Tao, J. et al. Thyroid hormone receptor‐beta agonist
2018;68(2):707–22. GC‐1 inhibits Met‐beta‐catenin‐driven hepatocellular cancer. Am J Pathol,
105. Preziosi, M., Okabe, H., Poddar, M., Singh, S., and Monga, S.P. Endothelial 2017;187(11):2473–85.
Wnts regulate β‐catenin signaling in murine liver zonation and regenera- 129. Gebhardt, R. Speeding up hepatocyte proliferation: how triiodothyronine
tion: a sequel to the Wnt–Wnt situation. Hepatol Commun, 2018;2(7). and beta‐catenin join forces. Hepatology, 2014;59(6):2074–6.
106. Planas‐Paz, L., Orsini, V., Boulter, L. et  al. The RSPO‐LGR4/5‐ZNRF3/ 130. Gebhardt, R. Metabolic zonation of the liver: regulation and implications
RNF43 module controls liver zonation and size. Nat Cell Biol, for liver function. Pharmacol Ther, 1992;53(3):275–354.
2016;18(5):467–79. 131. Audard, V., Grimber, G., Elie, C. et al. Cholestasis is a marker for hepato-
107. Nejak‐Bowen, K.N., Zeng, G., Tan, X., Cieply, B., and Monga, S.P. Beta‐ cellular carcinomas displaying beta‐catenin mutations. J Pathol, 2007;
catenin regulates vitamin C biosynthesis and cell survival in murine liver. J 212(3):345–52.
Biol Chem, 2009;284(41):28115–27. 132. Behari, J., Yeh, T.H., Krauland, L. et al. Liver‐specific beta‐catenin knock-
108. Sambrotta, M., Strautnieks, S., Papouli, E. et al. Mutations in TJP2 cause out mice exhibit defective bile acid and cholesterol homeostasis and
progressive cholestatic liver disease. Nat Genet, 2014;46(4):326–8. increased susceptibility to diet‐induced steatohepatitis. Am J Pathol,
109. Wickline, E.D., Awuah, P.K., Behari, J., Ross, M., Stolz, D.B., and Monga, 2010;176(2):744–53.
S.P. Hepatocyte gamma‐catenin compensates for conditionally deleted 133. Gougelet, A., Torre, C., Veber, P. et al. T‐cell factor 4 and beta‐catenin chro-
beta‐catenin at adherens junctions. J Hepatol, 2011;55(6):1256–62. matin occupancies pattern zonal liver metabolism in mice. Hepatology,
110. Wickline, E.D., Du, Y., Stolz, D.B., Kahn, M., and Monga, S.P. Gamma‐catenin 2014;59(6):2344–57.
at adherens junctions: mechanism and biologic implications in hepatocellular 134. Tsuchida, T. and Friedman, S.L. Mechanisms of hepatic stellate cell activa-
cancer after beta‐catenin knockdown. Neoplasia, 2013;15(4):421–34. tion. Nat Rev Gastroenterol Hepatol, 2017;14(7):397–411.
111. Mankertz, J., Hillenbrand, B., Tavalali, S., Huber, O., Fromm, M., and 135. Shackel, N.A., McGuinness, P.H., Abbott, C.A., Gorrell, M.D., and
Schulzke, J.D. Functional crosstalk between Wnt signaling and Cdx‐related McCaughan, G.W. Identification of novel molecules and pathogenic path-
transcriptional activation in the regulation of the claudin‐2 promoter activ- ways in primary biliary cirrhosis: cDNA array analysis of intrahepatic
ity. Biochem Biophys Res Commun, 2004;314(4):1001–7. ­differential gene expression. Gut, 2001;49(4):565–76.
112. Yeh, T.H., Krauland, L., Singh, V. et al. Liver‐specific beta‐catenin knock- 136. Tanaka, A., Leung, P.S., Kenny, T.P. et al. Genomic analysis of differen-
out mice have bile canalicular abnormalities, bile secretory defect, and tially expressed genes in liver and biliary epithelial cells of patients with
intrahepatic cholestasis. Hepatology, 2010;52(4):1410–9. primary biliary cirrhosis. J Autoimmun, 2001;17(1):89–98.
113. Pradhan‐Sundd, T., Zhou, L., Vats, R. et  al. Dual catenin loss in murine 137. Jiang, F., Parsons, C.J., and Stefanovic, B. Gene expression profile of qui-
liver causes tight junctional deregulation and progressive intrahepatic chol- escent and activated rat hepatic stellate cells implicates Wnt signaling path-
estasis. Hepatology, 2018;67(6):2320–37. way in activation. J Hepatol, 2006;45(3):401–9.
114. Vitale, G., Gitto, S., Raimondi, F. et al. Cryptogenic cholestasis in young 138. Rashid, S.T., Humphries, J.D., Byron, A. et al. Proteomic analysis of extra-
and adults: ATP8B1, ABCB11, ABCB4, and TJP2 gene variants analysis by cellular matrix from the hepatic stellate cell line LX‐2 identifies CYR61
high‐throughput sequencing. J Gastroenterol, 2018;53(8):945–58. and Wnt‐5a as novel constituents of fibrotic liver. J Proteome Res,
115. Apte, U., Zeng, G., Thompson, M.D. et al. Beta‐catenin is critical for early 2012;11(8):4052–64.
postnatal liver growth. Am J Physiol Gastrointest Liver Physiol, 139. Xiong, W.J., Hu, L.J., Jian, Y.C. et al. Wnt5a participates in hepatic stellate
2007;292(6):G1578–85. cell activation observed by gene expression profile and functional assays.
116. Cadoret, A., Ovejero, C., Saadi‐Kheddouci, S. et al. Hepatomegaly in trans- World J Gastroenterol, 2012;18(15):1745–52.
genic mice expressing an oncogenic form of beta‐catenin. Cancer Res, 140. Kordes, C., Sawitza, I., and Haussinger, D. Canonical Wnt signaling main-
2001;61(8):3245–9. tains the quiescent stage of hepatic stellate cells. Biochem Biophys Res
117. Colnot, S., Decaens, T., Niwa‐Kawakita, M. et al. Liver‐targeted disruption Commun, 2008;367(1):116–23.
of Apc in mice activates beta‐catenin signaling and leads to hepatocellular 141. Ge, W.S., Wang, Y.J., Wu, J.X., Fan, J.G., Chen, Y.W., and Zhu, L. beta‐
carcinomas. Proc Natl Acad Sci USA, 2004;101(49):17216–21. catenin is overexpressed in hepatic fibrosis and blockage of Wnt/­
118. Harada, N., Miyoshi, H., Murai, N. et  al. Lack of tumorigenesis in the beta‐catenin signaling inhibits hepatic stellate cell activation. Mol Med Rep,
mouse liver after adenovirus‐mediated expression of a dominant stable 2014;9(6):2145–51.
mutant of beta‐catenin. Cancer Res, 2002;62(7):1971–7. 142. Cheng, J.H., She, H., Han, Y.P. et al. Wnt antagonism inhibits hepatic stel-
119. Tan, X., Apte, U., Micsenyi, A. et al. Epidermal growth factor receptor: a late cell activation and liver fibrosis. Am J Physiol Gastrointest Liver
novel target of the Wnt/beta‐catenin pathway in liver. Gastroenterology, Physiol, 2008;294(1):G39–49.
2005;129(1):285–302. 143. Myung, S.J., Yoon, J.H., Gwak, G.Y. et al. Wnt signaling enhances the acti-
120. Nejak‐Bowen, K.N., Thompson, M.D., Singh, S. et  al. Accelerated liver vation and survival of human hepatic stellate cells. FEBS Lett,
regeneration and hepatocarcinogenesis in mice overexpressing serine‐45 2007;581(16):2954–8.
mutant beta‐catenin. Hepatology, 2010;51(5):1603–13. 144. Haughton, E.L., Tucker, S.J., Marek, C.J. et al. Pregnane X receptor activa-
121. Tetsu, O. and McCormick, F. Beta‐catenin regulates expression of cyclin tors inhibit human hepatic stellate cell transdifferentiation in vitro.
D1 in colon carcinoma cells. Nature, 1999;398(6726):422–6. Gastroenterology, 2006;131(1):194–209.
β-CATENIN SIGNALING
46:  601

145. Zhu, N.L., Wang, J., and Tsukamoto, H. The Necdin‐Wnt pathway causes 169. de La Coste, A., Romagnolo, B., Billuart, P. et al. Somatic mutations of the
epigenetic peroxisome proliferator‐activated receptor gamma repression in beta‐catenin gene are frequent in mouse and human hepatocellular carcino-
hepatic stellate cells. J Biol Chem, 2010;285(40):30463–71. mas. Proc Natl Acad Sci USA, 1998;95(15):8847–51.
146. Yanagida, A., Iwaisako, K., Hatano, E. et al. Downregulation of the Wnt 170. Schulze, K., Imbeaud, S., Letouze, E. et al. Exome sequencing of hepato-
antagonist Dkk2 links the loss of Sept4 and myofibroblastic transformation cellular carcinomas identifies new mutational signatures and potential
of hepatic stellate cells. Biochim Biophys Acta, 2011;1812(11):1403–11. therapeutic targets. Nat Genet, 2015;47(5):505–11.
147. Li, W., Zhu, C., Li, Y., Wu, Q., and Gao, R. Mest attenuates CCl4‐induced 171. Tao, J., Xu, E., Zhao, Y. et al. Modeling a human hepatocellular carcinoma
liver fibrosis in rats by inhibiting the Wnt/beta‐catenin signaling pathway. subset in mice through coexpression of met and point‐mutant beta‐catenin.
Gut Liver, 2014;8(3):282–91. Hepatology, 2016;64(5):1587–605.
148. Ross, S.E., Hemati, N., Longo, K.A. et al. Inhibition of adipogenesis by 172. Torrecilla, S., Sia, D., and Harrington, A.N. Trunk mutational events pre-
Wnt signaling. Science, 2000;289(5481):950–3. sent minimal intra‐ and inter‐tumoral heterogeneity in hepatocellular carci-
149. Kweon, S.M., Chi, F., Higashiyama, R., Lai, K., and Tsukamoto, H. Wnt noma. J Hepatol, 2017;67(6):1222–31.
pathway stabilizes MeCP2 protein to repress PPAR‐gamma in activation of 173. Satoh, S., Daigo, Y., Furukawa, Y. et al. AXIN1 mutations in hepatocellular
hepatic stellate cells. PLoS One, 2016;11(5):e0156111. carcinomas, and growth suppression in cancer cells by virus‐mediated
150. Lai, K.K.Y., Kweon, S.M., Chi, F. et al. Stearoyl‐CoA desaturase promotes transfer of AXIN1. Nat Genet, 2000;24(3):245–50.
liver fibrosis and tumor development in mice via a Wnt positive‐signaling 174. Zucman‐Rossi, J., Benhamouche, S., Godard, C. et al. Differential effects
loop by stabilization of low‐density lipoprotein‐receptor‐related proteins 5 of inactivated Axin1 and activated beta‐catenin mutations in human hepato-
and 6. Gastroenterology, 2017;152(6):1477–91. cellular carcinomas. Oncogene, 2007;26(5):774–80.
151. Hughes, L.J. and Michels, V.V. Risk of hepatoblastoma in familial adeno- 175. Merle, P., de la Monte, S., Kim, M. et al. Functional consequences of friz-
matous polyposis. Am J Med Genet, 1992;43(6):1023–5. zled‐7 receptor overexpression in human hepatocellular carcinoma.
152. Kurahashi, H., Takami, K., Oue, T. et al. Biallelic inactivation of the APC Gastroenterology, 2004;127(4):1110–22.
gene in hepatoblastoma. Cancer Res, 1995;55(21):5007–11. 176. Nambotin, S.B., Lefrancois, L., Sainsily, X. et al. Pharmacological inhibi-
153. Oda, H., Imai, Y., Nakatsuru, Y., Hata, J., and Ishikawa, T. Somatic muta- tion of Frizzled‐7 displays anti‐tumor properties in hepatocellular carci-
tions of the APC gene in sporadic hepatoblastomas. Cancer Res, noma. J Hepatol, 2010.
1996;56(14):3320–3. 177. Kim, M., Lee, H.C., Tsedensodnom, O. et al. Functional interaction between
154. Koch, A., Denkhaus, D., Albrecht, S., Leuschner, I., von Schweinitz, D., and Wnt3 and Frizzled‐7 leads to activation of the Wnt/beta‐catenin signaling
Pietsch, T. Childhood hepatoblastomas frequently carry a mutated degrada- pathway in hepatocellular carcinoma cells. J Hepatol, 2008;48(5):780–91.
tion targeting box of the beta‐catenin gene. Cancer Res, 1999;59(2):269–73. 178. Ban, K.C., Singh, H., Krishnan, R., and Seow, H.F. GSK‐3beta phospho-
155. Taniguchi, K., Roberts, L.R., Aderca, I.N. et  al. Mutational spectrum of rylation and alteration of beta‐catenin in hepatocellular carcinoma. Cancer
beta‐catenin, AXIN1, and AXIN2 in hepatocellular carcinomas and hepato- Lett, 2003;199(2):201–8.
blastomas. Oncogene, 2002;21(31):4863–71. 179. Shih, Y.L., Shyu, R.Y., Hsieh, C.B. et  al. Promoter methylation of the
156. Cairo, S., Armengol, C., De Reynies, A. et al. Hepatic stem‐like phenotype secreted frizzled‐related protein 1 gene SFRP1 is frequent in hepatocellular
and interplay of Wnt/beta‐catenin and Myc signaling in aggressive child- carcinoma. Cancer, 2006;107(3):579–90.
hood liver cancer. Cancer Cell, 2008;14(6):471–84. 180. Takagi, H., Sasaki, S., Suzuki, H. et al. Frequent epigenetic inactivation of
157. Bell, D., Ranganathan, S., Tao, J., and Monga, S.P. Novel advances in SFRP genes in hepatocellular carcinoma. J Gastroenterol, 2008;43(5):
understanding of molecular pathogenesis of hepatoblastoma: a Wnt/ 378–89.
beta‐catenin perspective. Gene Expr, 2017;17(2):141–54. 181. Hoshida, Y., Nijman, S.M., Kobayashi, M. et al. Integrative transcriptome
158. Lade, A.G. and Monga, S.P. Beta‐catenin signaling in hepatic development analysis reveals common molecular subclasses of human hepatocellular
and progenitors: which way does the WNT blow? Dev Dyn, 2011;240(3): carcinoma. Cancer Res, 2009;69(18):7385–92.
486–500. 182. Cieply, B., Zeng, G., Proverbs‐Singh, T., Geller, D.A., and Monga, S.P.
159. Shin, D. and Monga, S.P. Cellular and molecular basis of liver develop- Unique phenotype of hepatocellular cancers with exon‐3 mutations in beta‐
ment. Compr Physiol, 2013;3(2):799–815. catenin gene. Hepatology, 2009;49(3):821–31.
160. Lopez‐Terrada, D., Gunaratne, P.H., Adesina, A.M. et al. Histologic sub- 183. Boyault, S., Rickman, D.S., de Reynies, A. et al. Transcriptome classifica-
types of hepatoblastoma are characterized by differential canonical Wnt tion of HCC is related to gene alterations and to new therapeutic targets.
and Notch pathway activation in DLK+ precursors. Hum Pathol, Hepatology, 2007;45(1):42–52.
2009;40(6):783–94. 184. Rebouissou, S., Franconi, A., Calderaro, J. et al. Genotype‐phenotype cor-
161. Armengol, C., Cairo, S., Fabre, M., and Buendia, M.A. Wnt signaling and relation of CTNNB1 mutations reveals different ss‐catenin activity associ-
hepatocarcinogenesis: the hepatoblastoma model. Int J Biochem Cell Biol, ated with liver tumor progression. Hepatology, 2016;64(6):2047–61.
2011;43(2):265–70. 185. Yuan, R.H., Jeng, Y.M., Hu, R.H. et al. Role of p53 and beta‐catenin muta-
162. Tao, J., Calvisi, D.F., Ranganathan, S. et al. Activation of beta‐catenin and tions in conjunction with CK19 expression on early tumor recurrence and
Yap1 in human hepatoblastoma and induction of hepatocarcinogenesis in prognosis of hepatocellular carcinoma. J Gastrointest Surg, 2011;15(2):
mice. Gastroenterology, 2014;147(3):690–701. 321–9.
163. Wang, H., Lu, J., Edmunds, L.R. et al. Coordinated activities of multiple 186. Nhieu, J.T., Renard, C.A., Wei, Y., Cherqui, D., Zafrani, E.S., and Buendia,
Myc‐dependent and Myc‐independent biosynthetic pathways in hepato- M.A. Nuclear accumulation of mutated beta‐catenin in hepatocellular car-
blastoma. J Biol Chem, 2016;291(51):26241–51. cinoma is associated with increased cell proliferation. Am J Pathol,
164. Molina, L., Bell, D., Tao, J. et al. Hepatocyte‐derived lipocalin 2 is a poten- 1999;155(3):703–10.
tial serum biomarker reflecting tumor burden in hepatoblastoma. Am J 187. Suzuki, T., Yano, H., Nakashima, Y., Nakashima, O., and Kojiro, M. Beta‐
Pathol, 2018;188(8):1895–909. catenin expression in hepatocellular carcinoma: a possible participation of
165. Nault, J.C., Bioulac‐Sage, P., and Zucman‐Rossi, J. Hepatocellular benign beta‐catenin in the dedifferentiation process. J Gastroenterol Hepatol,
tumors‐from molecular classification to personalized clinical care. 2002;17(9):994–1000.
Gastroenterology, 2013;144(5):888–902. 188. Dal Bello, B., Rosa, L., Campanini, N. et al. Glutamine synthetase immu-
166. Zucman‐Rossi, J., Jeannot, E., Nhieu, J.T. et al. Genotype‐phenotype cor- nostaining correlates with pathologic features of hepatocellular carcinoma
relation in hepatocellular adenoma: new classification and relationship with and better survival after radiofrequency thermal ablation. Clin Cancer
HCC. Hepatology, 2006;43(3):515–24. Res,Cancer Res, 2010;16(7):2157–66.
167. Bioulac‐Sage, P., Rebouissou, S., Thomas, C. et  al. Hepatocellular ade- 189. Lok, A.S., Seeff, L.B., Morgan, T.R. et al. Incidence of hepatocellular car-
noma subtype classification using molecular markers and immunohisto- cinoma and associated risk factors in hepatitis C‐related advanced liver
chemistry. Hepatology, 2007;46(3):740–8. disease. Gastroenterology, 2009;136(1):138–48.
168. Ihara, A., Koizumi, H., Hashizume, R., and Uchikoshi, T. Expression of 190. Lee, J.M., Yang, J., Newell, P. et al. beta‐Catenin signaling in hepatocellu-
epithelial cadherin and alpha‐ and beta‐catenins in nontumoral livers and lar cancer: Implications in inflammation, fibrosis, and proliferation. Cancer
hepatocellular carcinomas. Hepatology, 1996;23(6):1441–7. Lett, 2014;343(1):90–7.
602 THE LIVER:  REFERENCES

191. Harada, N., Oshima, H., Katoh, M., Tamai, Y., Oshima, M., and Taketo, 213. Rashid, A., Gao, Y.T., Bhakta, S. et  al. Beta‐catenin mutations in biliary
M.M. Hepatocarcinogenesis in mice with beta‐catenin and Ha‐ras gene tract cancers: a population‐based study in China. Cancer Res, 2001;
mutations. Cancer Res, 2004;64(1):48–54. 61(8):3406–9.
192. Miyoshi, H., Deguchi, A., Nakau, M. et  al. Hepatocellular carcinoma 214. Abraham, S.C., Lee, J.H., Hruban, R.H., Argani, P., Furth, E.E., and Wu,
development induced by conditional beta‐catenin activation in Lkb1+/‐ T.T. Molecular and immunohistochemical analysis of intraductal papillary
mice. Cancer Sci, 2009;100(11):2046–53. neoplasms of the biliary tract. Hum Pathol, 2003;34(9):902–10.
193. Bauer‐Hofmann, R., Klimek, F., Buchmann, A., Muller, O., Bannasch, P., 215. Lustig, B., Jerchow, B., Sachs, M. et  al. Negative feedback loop of Wnt
and Schwarz, M. Role of mutations at codon 61 of the c‐Ha‐ras gene during signaling through upregulation of conductin/axin2 in colorectal and liver
diethylnitrosamine‐induced hepatocarcinogenesis in C3H/He mice. Mol tumors. Mol Cell Biol, 2002;22(4):1184–93.
Carcinog, 1992;6(1):60–7. 216. Takayasu, H., Horie, H., Hiyama, E. et al. Frequent deletions and mutations
194. Nault, J.C., Mallet, M., Pilati, C. et  al. High frequency of telomerase of the beta‐catenin gene are associated with overexpression of cyclin D1
reverse‐transcriptase promoter somatic mutations in hepatocellular carci- and fibronectin and poorly differentiated histology in childhood hepato-
noma and preneoplastic lesions. Nat Commun, 2013;4:2218. blastoma. Clin Cancer Res, 2001;7(4):901–8.
195. Tao, J., Zhang, R., Singh, S. et al. Targeting beta‐catenin in hepatocellular 217. Torre, C., Benhamouche, S., Mitchell, C. et al. The transforming growth
cancers induced by coexpression of mutant beta‐catenin and K‐Ras in mice. factor‐alpha and cyclin D1 genes are direct targets of beta‐catenin signaling
Hepatology, 2017;65(5):1581–99. in hepatocyte proliferation. J Hepatol, 2011;55(1):86–95.
196. Zeng, G., Apte, U., Cieply, B., Singh, S., and Monga, S.P. siRNA‐mediated 218. Araki, Y., Okamura, S., Hussain, S.P. et al. Regulation of cyclooxygenase‐2
beta‐catenin knockdown in human hepatoma cells results in decreased expression by the Wnt and ras pathways. Cancer Res, 2003;63(3):728–34.
growth and survival. Neoplasia, 2007;9(11):951–9. 219. Loeppen, S., Koehle, C., Buchmann, A., and Schwarz, M. A beta‐catenin‐
197. Yao, M., Kargman, S., Lam, E.C. et al. Inhibition of cyclooxygenase‐2 by dependent pathway regulates expression of cytochrome P450 isoforms in
rofecoxib attenuates the growth and metastatic potential of colorectal carci- mouse liver tumors. Carcinogenesis, 2005;26(1):239–48.
noma in mice. Cancer Res, 2003;63(3):586–92. 220. Yamamoto, Y., Sakamoto, M., Fujii, G. et al. Overexpression of orphan G‐
198. Behari, J., Zeng, G., Otruba, W. et al. R‐Etodolac decreases beta‐catenin protein‐coupled receptor, Gpr49, in human hepatocellular carcinomas with
levels along with survival and proliferation of hepatoma cells. J Hepatol, beta‐catenin mutations. Hepatology, 2003;37(3):528–33.
2007;46(5):849–57. 221. Ovejero, C., Cavard, C., Perianin, A. et al. Identification of the leukocyte
199. Li, H., Pamukcu, R., and Thompson, W.J. Beta‐catenin signaling: therapeu- cell‐derived chemotaxin 2 as a direct target gene of beta‐catenin in the liver.
tic strategies in oncology. Cancer Biol Ther, 2002;1(6):621–5. Hepatology, 2004;40(1):167–76.
200. Emami, K.H., Nguyen, C., Ma, H. et al. A small molecule inhibitor of beta‐ 222. Cavard, C., Terris, B., Grimber, G. et al. Overexpression of regenerating
catenin/CREB‐binding protein transcription [corrected]. Proc Natl Acad islet‐derived 1 alpha and 3 alpha genes in human primary liver tumors with
Sci USA, 2004;101(34):12682–7. beta‐catenin mutations. Oncogene, 2006;25(4):599–608.
201. Lenz, H.J. and Kahn, M. Safely targeting cancer stem cells via selective 223. Renard, C.A., Labalette, C., Armengol, C. et al. Tbx3 is a downstream tar-
catenin coactivator antagonism. Cancer Sci, 2014. get of the Wnt/beta‐catenin pathway and a critical mediator of beta‐catenin
202. Delgado, E.R., Yang, J., So, J. et al. Identification and characterization of a survival functions in liver cancer. Cancer Res, 2007;67(3):901–10.
novel small‐molecule inhibitor of beta‐catenin signaling. Am J Pathol, 224. Delgado, E., Bahal, R., Yang, J., Lee, J.M., Ly, D.H., and Monga, S.P. Beta‐
2014;184(7):2111–22. catenin knockdown in liver tumor cells by a cell permeable gamma guanidine‐
203. Nejak‐Bowen, K.N. and Monga, S.P. Beta‐catenin signaling, liver regener- based peptide nucleic acid. Curr Cancer Drug Targets, 2013;13(8):
ation and hepatocellular cancer: sorting the good from the bad. Semin 867–78.
Cancer Biol, 2011;21(1):44–58. 225. Guichard, C., Amaddeo, G., Imbeaud, S. et  al. Integrated analysis of
204. Okabe, H., Delgado, E., Lee, J.M. et  al. Role of leukocyte cell‐derived somatic mutations and focal copy‐number changes identifies key genes and
chemotaxin 2 as a biomarker in hepatocellular carcinoma. PLoS One, pathways in hepatocellular carcinoma. Nat Genet, 2012;44(6):694–8.
2014;9(6):e98817. 226. Cui, J., Zhou, X., Liu, Y., Tang, Z., and Romeih, M. Alterations of
205. Zhan, N., Michael, A.A., Wu, K. et al. The effect of selective c‐MET inhibi- beta‐catenin and Tcf‐4 instead of GSK‐3beta contribute to activation
tor on hepatocellular carcinoma in the MET‐active, beta‐catenin‐mutated of  Wnt pathway in hepatocellular carcinoma. Chin Med J (Engl),
mouse model. Gene Expr, 2018;18(2):135–47. 2003;116(12):1885–92.
206. Delgado, E., Okabe, H., Preziosi, M. et al. Complete response of Ctnnb1‐ 227. Devereux, T.R., Stern, M.C., Flake, G.P. et  al. CTNNB1 mutations and
mutated tumours to beta‐catenin suppression by locked nucleic acid anti- beta‐catenin protein accumulation in human hepatocellular carcinomas
sense in a mouse hepatocarcinogenesis model. J Hepatol, 2014. associated with high exposure to aflatoxin B1. Mol Carcinog, 2001;31(2):
207. Adebayo Michael, A.O., Ko, S., and Tao, J. Inhibiting glutamine‐dependent 68–73.
mTORC1 activation ameliorates liver cancers driven by β‐catenin muta- 228. Wong, C.M., Fan, S.T., and Ng, I.O. Beta‐catenin mutation and overexpres-
tions. Cell Metab, 2019;29(5):1135–50. sion in hepatocellular carcinoma: clinicopathologic and prognostic signifi-
208. Ruiz de Galarreta, M., Bresnahan, E., and Molina‐Sanchez, P. β‐catenin cance. Cancer, 2001;92(1):136–45.
activation promotes immune escape and resistance to anti‐PD‐1 therapy in 229. Hsu, H.C., Jeng, Y.M., Mao, T.L., Chu, J.S., Lai, P.L., and Peng, S.Y. Beta‐
hepatocellular carcinoma. Cancer Discov, 2019. doi: 10.1158/2159‐8290. catenin mutations are associated with a subset of low‐stage hepatocellular
CD‐19‐0074 carcinoma negative for hepatitis B virus and with favorable prognosis. Am
209. Nakanuma, Y., Harada, K., Ishikawa, A., Zen, Y., and Sasaki, M. Anatomic J Pathol, 2000;157(3):763–70.
and molecular pathology of intrahepatic cholangiocarcinoma. J Hepatobiliary 230. Huang, H., Fujii, H., Sankila, A. et al. Beta‐catenin mutations are frequent
Pancreat Surg, 2003;10(4):265–81. in human hepatocellular carcinomas associated with hepatitis C virus infec-
210. Ashida, K., Terada, T., Kitamura, Y., and Kaibara, N. Expression of E‐cad- tion. Am J Pathol, 1999;155(6):1795–801.
herin, alpha‐catenin, beta‐catenin, and CD44 (standard and variant iso- 231. Legoix, P., Bluteau, O., Bayer, J. et al. Beta‐catenin mutations in hepatocel-
forms) in human cholangiocarcinoma: an immunohistochemical study. lular carcinoma correlate with a low rate of loss of heterozygosity.
Hepatology, 1998;27(4):974–82. Oncogene, 1999;18(27):4044–6.
211. Rashid, A. Cellular and molecular biology of biliary tract cancers. Surg 232. Kondo, Y., Kanai, Y., Sakamoto, M. et al. Beta‐catenin accumulation and
Oncol Clin N Am, 2002;11(4):995–1009. mutation of exon 3 of the beta‐catenin gene in hepatocellular carcinoma.
212. Sugimachi, K., Taguchi, K., Aishima, S. et al. Altered expression of beta‐ Jpn J Cancer Res, 1999;90(12):1301–9.
catenin without genetic mutation in intrahepatic cholangiocarcinoma. Mod
Pathol, 2001;14(9):900–5.
Polyploidy in Liver Function,
47 Mitochondrial Metabolism,
and Cancer
Evan R. Delgado, Elizabeth C. Stahl, Nairita Roy, Patrick D. Wilkinson,
and Andrew W. Duncan
Department of Pathology, McGowan Institute for Regenerative Medicine, Pittsburgh Liver Research Center,
University of Pittsburgh, Pittsburgh, PA, USA

HEPATIC CHROMOSOMAL VARIATIONS Polyploidy levels change profoundly during liver growth and
AND LIVER FUNCTION maturation. In the early stages of liver development, the liver is
primarily populated with small, uniform diploid (2n) hepato-
The development of polyploid cells in tissues and organs is a cytes [9, 10]. As the hepatocytes proliferate, the diploids fail to
highly regulated process. In the liver, failure to complete the complete cytokinesis, and in turn generate a daughter cell with
final stage of mitosis, known as cytokinesis, drives the retention two diploid nuclei (2n × 2n), referred to as a binucleate tetra-
of nuclear content to a single cell. The heterogeneity of poly- ploid (Figure  47.1a). Next, tetraploid cells with two diploid
ploid cells in the liver is further nuanced by the occurrence of nuclei replicate their DNA and generate two mononucleate
aneuploidy, where single chromosomes may be gained or lost, tetraploid (4n) daughter cells by successfully completing cytoki-
and by reductive mitosis, all driving the “ploidy conveyor.” nesis. Mononucleate tetraploid cells can produce mono‐ and
binucleate octaploid (8n and 4n × 4n) hepatocytes through sub-
sequent cell divisions with complete and incomplete cytokine-
sis. The process continues, producing hexadecaploid (16n) cells
Polyploidy in somatic tissues and the liver and hepatocytes with even greater ploidy states [2]. While fail-
The majority of mammalian cells have a diploid nuclear content ure to complete cytokinesis is the most common mechanism by
and contain two sets of chromosomes. However, there is a phe- which hepatocytes increase ploidy states, heterotypic cell fusion
nomenon called polyploidy, which describes cells with addi- between macrophages and hepatocytes has been shown to occur,
tional sets of chromosomes. Polyploid cells are an innate albeit infrequently (at approximately 1/100 000 hepatocytes)
component of many mammalian species, including rodents and [11]. In general, as the liver matures the hepatocyte population
humans [1, 2]. The formation and location of these polyploid becomes increasingly heterogeneous in size, function, and
cells is tissue‐dependent. For example, myofibrils and osteo- nuclear content, including number of nuclei per cell and chro-
clasts become multinucleate, polyploid cells via cell fusion mosome sets per nucleus [2, 8, 12].
[3, 4]. Others, including megakaryocytes and trophoblast giant The rate and percentage of liver polyploidization differ
cells, become polyploid through endoreduplication. Here, pro- between species [2]. At birth, the majority of human hepato-
liferating cells progress through the cell cycle but fail to com- cytes in the liver are diploid, with polyploids representing less
plete nuclear division during mitosis, resulting in a polyploid than 10% [13]. Notably, tetraploid hepatocytes appear within
cell with a single nucleus [5, 6]. On the other hand, failure of the first few weeks of development, followed by octaploid
cytokinesis (the final step of cell division where the parent cell’s hepatocytes at two to three months after birth [14]. By adult-
cytoplasm divides to produce two daughter cells) primarily hood, the percentage of polyploidy is nearly doubled to 20%,
drives the generation of polyploid cardiomyocytes and hepato- and after the age of 60, approximately 50% of the human liver is
cytes [2, 7, 8]. comprised of polyploid hepatocytes [13, 15]. Alternatively, the

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
604 THE LIVER:  HEPATIC CHROMOSOMAL VARIATIONS AND LIVER FUNCTION

Figure 47.1  Polyploidy and aneuploidy in the liver. (a) Hepatocytes are mononucleate or binucleate, and ploidy is determined by the number of
nuclei per cell and DNA content of each nucleus. Cell division associated with failed cytokinesis drives hepatic polyploidization. The cartoon shows
how a single diploid hepatocyte can give rise to mononucleate and binucleate tetraploid and octaploid hepatocytes. (b) In humans, diploid, tetra-
ploid, and octaploid hepatocytes contain 46, 92, and 184 chromosomes (Chr), respectively. Euploid karyotypes contain balanced sets of chromo-
somes, whereas aneuploid karyotypes contain one or more random chromosome gains and/or losses. (c) Polyploid hepatocytes can undergo
chromosome segregation errors (e.g. lagging chromosomes) and multipolar cell divisions to generate aneuploid daughter cells with reduced ploidy.
The flowchart outlines the various types of daughter cells that can be generated by a dividing tetraploid hepatocyte undergoing bipolar (left) or
multipolar (middle and right) nuclear segregation. (d) The ploidy conveyor model describes hepatic ploidy changes that occur during postnatal
development and proliferation. Cell divisions involving cytokinesis failure occur frequently and lead to increased ploidy. In contrast, cell divisions
associated with multipolar nuclear segregation are infrequent and lead to ploidy reversal and formation of diploid or near‐diploid daughters.
47:  Polyploidy in Liver Function, Mitochondrial Metabolism, and Cancer 605

livers of mice are already 50% polyploid at weaning, whereas THE IMPACT OF POLYPLOIDY
the livers of rats begin to polyploidize upon weaning, triggered ON GENE EXPRESSION AND LIVER
by changes in insulin AKT‐dependent signaling that result in
cytokinesis failure [16]. By eight to twelve weeks of age both
REGENERATION
mouse and rat livers become 80% polyploid, a faster rate than
The liver is a vitally important organ situated to metabolize hun-
human polyploidization, making rodents a useful model to
dreds of molecules for energy regulation and drug toxicity, as
study polyploidization in the span of a few weeks [17]. The sig-
well as to produce important blood anticoagulation proteins,
nals regulating the steady accumulation of polyploid hepato-
immune complexes, and digestive bile acids. These processes
cytes in developing and aging humans remain unknown but may
require the orchestration of complex gene networks and neces-
also be related to the insulin‐AKT pathway. Interestingly, the
sitate the maintenance of liver function for survival. As such, the
frequency of cells within each ploidy state remains relatively
liver has a robust capacity to adapt to injury and undergo com-
constant among individuals of the same age, suggesting the
plete regeneration. The impact of polyploidy on gene expression
existence of a ploidy sensor to maintain homeostasis in the liver.
and liver regeneration is just beginning to be understood.

Liver aneuploidy and the ploidy conveyor


Polyploidy in liver gene expression
While hepatocytes are heterogeneous in regard to the number of and function
chromosome sets, they also exhibit diversity at the individual
chromosome level. G‐banding, fluorescence in situ hybridiza- While the mechanisms that produce polyploid hepatocytes have
tion, and single cell DNA sequencing studies have indicated been well characterized, the relationship between ploidy and
4–50% of hepatocytes in healthy livers of mice and humans are gene expression remains unclear. Experiments have shown that
aneuploid, having gained or lost individual chromosomes polyploidy is associated with specific gene expression patterns
(Figure 47.1b) [13, 17, 18]. Recently, it was shown that the liv- in yeast [24]. In regards to mammals, studies have demonstrated
ers of mice deficient in microRNA‐122 had reduced polyploid that nuclear content affects expression levels of genes regulating
levels and were enriched with diploid hepatocytes. Moreover, megakaryocyte differentiation [25] and that giant trophoblast
the livers had abnormally low aneuploidy levels [19]. cells could exhibit biallelic X‐chromosome gene expression due
Additionally, studies have shown that polyploidy and aneu- to incomplete X‐inactivation [26]. Additionally, a large‐scale
ploidy levels follow similar trends, as both increase with age genome comparative study illustrated that cardiac stress genes
and are the highest in adult murine and human livers [2, 20]. The were expressed at different levels between diploid and polyploid
degree of aneuploidy plateaus in mice from four to fifteen cardiomyocytes [27].
months of age [17]. Similarly, in humans, hepatic aneuploidy It has been postulated that hepatic ploidy state could affect
levels remain stable between the ages of 10 and 60 years [13]. gene expression. One hypothesis is that gene expression levels
During mitosis, polyploid hepatocytes can form multipolar increase proportionally with ploidy. In this case, tetraploids and
spindles and drive “reductive mitosis,” or a reduction in nuclear octaploids would have 2× and 4× greater gene expression levels,
content, a process called “ploidy reversal” (Figure  47.1c,d). respectively, compared to diploids. Thus, tetraploid and octa-
One example of ploidy reversal is illustrated with a proliferat- ploid cells would also potentially contain 2× and 4× the total
ing tetraploid hepatocyte that forms a tripolar spindle during RNA and protein content as diploid cells [2]. Another hypothe-
mitosis (Figure 47.1c, middle panel). The chromosomes of the sis is that diploid and polyploid hepatocytes exhibit differential
dividing cell are pulled toward three poles, resulting in two gene expression. In this scenario, gene expression levels would
daughter cells with approximately diploid nuclei and one vary on a gene to gene basis between ploidy populations, with
daughter cell with a ~tetraploid nucleus. Additionally, poly- polyploids exhibiting greater expression of specific genes ver-
ploid cells can perform ploidy reversal by undergoing double sus diploids and vice versa. A 2007 study by Lu et al. compared
mitosis (Figure  47.1c, right panel). In this case, proliferating gene expression between quiescent diploid, tetraploid, and octa-
tetraploids can generate up to four ~diploid daughter cells and ploid hepatocyte populations isolated by flow cytometry.
octaploids can generate up to eight ~diploid daughter cells [17, Surprisingly, the study found gene expression was mostly
21, 22]. Chromosome segregation errors often occur during equivalent between ploidy populations, and the magnitude of
multipolar cell division [17, 21]. For instance, microtubules difference was small for those genes with different expression
associated with the outer poles occasionally attach to the same levels [28]. This suggested that few genes are differentially
kinetochore during multipolar spindle construction. If left expressed, at least in quiescent hepatocytes. Further studies are
unchecked, this error prevents chromosomes from migrating needed to determine if differential gene expression occurs in
to discrete poles during anaphase. Consequently, these “lag- disease conditions or in response to specific stimuli because
ging” chromosomes are frequently excluded from the daughter variations at the molecular level could ultimately endow certain
cell nuclei and form what is considered a micronucleus. ploidy subsets with unique functional capabilities [2, 29].
Cytogenetic analyses of hepatocytes have shown that liver
aneuploidy is widespread, with chromosome gains and losses
Liver ploidy supports hepatic adaptation
occurring in an unbiased fashion [13, 23]. Polyploidization,
ploidy reversal and aneuploidy, which are collectively called It has been hypothesized that the diverse population of polyploid
the “ploidy conveyor”, prominently drive hepatocyte heteroge- and aneuploid hepatocytes endows the liver with the ability to
neity (Figure 47.1c,d) [2]. adapt to a variety of environmental stresses [23]. Interestingly,
606 THE LIVER:  THE IMPACT OF POLYPLOIDY ON GENE EXPRESSION AND LIVER REGENERATION

Figure 47.2  Proof of concept studies demonstrated that disease‐resistant aneuploid hepatocytes promote adaptation to chronic liver injury.
Healthy mouse livers contain randomly aneuploid hepatocytes, which are illustrated by multicolored cells. In response to chronic injury
(induced by tyrosinemia in Hgd+/− Fah−/− mice), liver function was severely impaired. Hepatocytes lacking chromosome 16 (blue) were resist-
ant to the injury and capable of proliferation. Clonal expansion by these injury‐resistant hepatocytes spontaneously repopulated the liver and
restored liver function.

studies have shown that budding yeast strains with specific chro- proliferative capacities of diploids and polyploids are equivalent.
mosome numbers are able to survive toxic insults and deleterious One study demonstrated that livers from rats that had undergone
mutations [30, 31]. In the context of the liver, it has been demon- PH had increased polyploidy levels and changes associated with
strated that aneuploid hepatocytes could protect against chronic cell senescence and aging [37]. Another study showed that mice
liver disease. Notably, a study illustrated that mice suffering with drug‐induced necrosis and cirrhosis developed regenerating
from tyrosinemia due to loss of fumarylacetoacetate hydrolase nodules enriched with diploid hepatocytes [38]. The idea that
(FAH) an enzyme in the tyrosine catabolic pathway, developed diploids are more proliferative than polyploid hepatocytes is also
resistance to the disease partially through hepatic aneuploidy supported by observations that diploid hepatocytes are enriched
[23] (Figure  47.2). Mice suffering from tyrosinemia can be in patients and rodents with hepatocellular carcinomas (HCC)
treated and kept healthy by blocking the tyrosine catabolic path- [38–41]. In contrast, it was shown that transplanted octaploid
way upstream of FAH using the drug 2‐(2‐nitro‐4‐trifluoro‐ and diploid hepatocytes proliferate equivalently in the Fah−/−
methylbenzoyl)‐1,3‐cyclo‐hexanedione (NTBC) or deleting the liver repopulation model. However, a limitation of this study was
enzyme homogentisic acid dioxygenase (HGD) [32, 33]. In the that the ploidy states of the transplanted cells changed as the
aforementioned study, NTBC was withdrawn from Fah−/− Hgd+/− octaploid cells underwent ploidy reversal and the diploids poly-
mice and, contrary to the expectation that all mice would suc- ploidized during the course of repopulation (Figure 47.1c) [17].
cumb to liver failure, some mice developed numerous small, Taken together, these experiments illustrate that further studies
healthy nodules in their livers. Karyotyping and array compara- are needed to determine the precise proliferative capacities of
tive genomic hybridization analyses revealed that many of the diploid and polyploid hepatocyte populations, and the mecha-
healthy nodules were comprised of aneuploid hepatocytes lack- nisms regulating proliferation of each ploidy population.
ing a copy of chromosome 16, which contains the wild‐type
copy of Hgd. Thus, the healthy liver nodules were populated Polyploidy in aging and impaired
with Hgd‐deficient hepatocytes. It was hypothesized that chronic
injury was toxic to the majority of hepatocytes except those that
liver regeneration
had previously lost the chromosome with the Hgd wild‐type Aging is known to lead to the accumulation of senescent cells in
copy. The disease‐resistant hepatocytes (monosomic for chro- multiple tissues and organ systems, including the liver [42].
mosome 16), proliferated and repopulated the liver and restored Senescence is the irreversible exit of the cell cycle, driven by the
normal liver function [23]. shortening of telomeres (i.e. the protective caps of chromo-
somes) through exhaustive replication or other stressors. Many
hypothesize that senescence is a protective mechanism against
Polyploidy and liver regeneration
tumorigenesis that might arise from genomic instability, but
The regenerative capacity of the liver has been well documented, senescence also has the effect of limiting tissue regeneration
but the role of polyploidy in this process remains unclear. It is and promoting the secretion of inflammatory mediators that can
hypothesized that diploid and polyploid populations could play be damaging to surrounding tissue [43]. Aged octaploid cells
unique roles in the process. Originally, it was believed that poly- were found to express a higher proportion of senescence mark-
ploid hepatocytes were mature, terminally differentiated cells ers compared to tetraploid and diploid cells, including p16, p21,
with little proliferative capacity. This was suggested from studies and p53, which also serve important tumor suppressor roles
demonstrating that hepatocytes in livers of mice and rats become [44]. Methods to deplete senescent cells in aged organs and
increasingly polyploid with age, and that more than 99% of ­tissues, such as the INK‐ATTAC mouse, were not capable of
hepatocytes in adult livers were quiescent [34, 35]. However, depleting senescent cells in the liver but reduced the spontane-
polyploids have been observed to proliferate in response to par- ous occurrence of liver tumors by depleting senescent cells at
tial hepatectomy (PH), the surgical removal of up to two‐thirds peripheral sites, thus demonstrating the role of the aged
of the liver mass [36]. While these experiments confirmed that ­microenvironment in the etiology of liver cancers and other age‐
polyploid hepatocytes can proliferate, it is unclear if the related diseases [42, 43].
47:  Polyploidy in Liver Function, Mitochondrial Metabolism, and Cancer 607

Unlike senescent diploid cells, which must be cleared by the including disrupted mitochondrial membranes, dysregulation of
immune system, senescent polyploid cells have the unique oxidative phosphorylation (OXPHOS), production of reactive
capacity to undergo ploidy reversal [44]. Aged hepatocytes that oxygen species (superoxides, nitric oxides, and lipid peroxides),
undergo ploidy reversal also downregulate expression of senes- glutathione depletion, release of cytochrome c from the mito-
cence markers, acting as a type of cell rejuvenation [44]. Ploidy chondrial matrix, defects in mitochondrial fission and fusion,
reversal might partially explain the capacity of the aged liver to abnormal mitophagy, and defective retrograde or anterograde
regenerate, although regeneration is substantially reduced in signaling between the mitochondria and nucleus [50].
elderly individuals. After PH in mice, hepatocytes in aged livers Insulin resistance and hyperlipidemia are the major underly-
have delayed cell cycle entry and the livers are significantly ing metabolic states of NAFLD [52]. Insulin signaling fortifies
deficient in the number of proliferating hepatocytes [45–47]. the integrity of the electron transport chain in the mitochondria
The age‐related decline in hepatocyte proliferation has been by suppressing FOXO1 and maintaining the NAD+ : NADH
attributed to transcription factor C/EBPα complexing with Brm, ratio [53]. This NAD+ : NADH ratio then determines the activity
a chromatin remodeling protein detected only in aged livers of PGC1α, which is the major regulator of mitochondrial bio-
[48], which inhibits E2f‐regulated gene expression [49]. genesis [54]. Desdouets and colleagues recently showed that
Notably, when the circulatory systems of young and aged mice advanced fatty liver disease, known as non‐alcoholic steatohep-
are connected through heterochronic parabiosis, the number of atitis, is associated with enhanced polyploidy [55]. Livers from
hepatocytes undergoing proliferation increased in the aged liv- mice fed with diets that induce NAFLD, such as the methionine‐
ers. While the ploidy of the rejuvenated hepatocytes was not choline‐deficient diet or high‐fat diet, were also dramatically
documented, a reduction in the C/EBPα–Brm complexes was enriched with polyploid hepatocytes. Furthermore, in these fatty
observed [48]. liver models, oxidative stress was found to be the primary driver
When aged diploid or octaploid hepatocytes are implanted of increased polyploidy. Thus, there appears to be an intricate
into young Fah−/− mice, the octaploid hepatocytes undergo nexus between mitochondrial dysfunction, correlating with
ploidy reversal, giving rise to lower hepatic ploidy states that no metabolic liver disease and polyploidization.
longer express markers of senescence [17, 44]. This study not
only implicated the ability of a young systemic environment to The nexus of hepatic polyploidy
rejuvenate hepatocyte proliferative capacity, but also suggested
and mitochondrial metabolism
that hepatocytes with differential ploidy states have equivalent
proliferation kinetics during long‐term repopulation. Future Liver regeneration and diseases associated with altered mito-
studies will need to decipher the role of cell intrinsic factors, chondrial metabolism, including NAFLD and hemochromato-
such as ploidy and gene stability, mitochondrial function, and sis, are associated with alterations in ploidy [37, 55–57].
autophagy, versus extrinsic factors in the local or systemic envi- Moreover, many genes that regulate mitochondrial metabolism
ronment, in the process of hepatocyte function and regeneration are reported to alter hepatic polyploidy such as Mir122, mem-
after injury and during aging. bers of the E2F family (E2F1, E2F7, and E2F8), Birc5, Ercc1,
Myc, p53, Rb, and Skp2 [2, 19, 58, 59]. Table 47.1 summarizes
the dual role of key molecular regulators involved in mitochon-
drial metabolism and hepatic polyploidy. Together, these obser-
THE MITOCHONDRIA‐PLOIDY NEXUS vations strongly suggest that mitochondrial metabolism and
IN LIVER DISEASE hepatic polyploidy are closely linked and could play a role in
progression of liver diseases.
Liver regeneration and a host of liver diseases are associated
with hepatic ploidy alterations and mitochondrial metabolic Mitochondria–ploidy nexus in liver
dysfunction [50]. These observations suggest that the regulation
of energy expenditure or the breakdown of mitochondrial func-
regeneration
tion is linked to the development of differential ploidy states. Hepatic polyploidy has been reported to transiently increase
This section describes the metabolic states and molecular regu- when the liver regenerates following PH, during which concur-
lators that connect mitochondrial metabolism and polyploidy in rent changes occur in mitochondrial metabolism [37, 56]. First,
the liver. there is a strong positive correlation between the regenerative
capacity of the liver and the efficiency of OXPHOS, as demon-
Mitochondrial metabolism and strated by a subsequent increase in ATP levels and respiratory
control ratio, an indicator of OXPHOS efficiency [84]. Second,
metabolic disease expression of proteins involved in carbohydrate metabolism,
Mitochondria serve an important role in managing energy lipid metabolism, and OXPHOS increases 24 hours post PH
expenditure of the liver, through metabolism of lipids and other [85]. Third, there is heterogeneity among the type of mitochon-
molecules for ATP generation, which is essential for many cell drial populations in hepatocytes. These mitochondrial subtypes
functions. Dysregulated mitochondrial metabolism has been differ in iron uptake capacity, iron metabolism, and complex IV
reported in liver diseases such as drug‐induced liver injury, alco- activity, ranging from 2–42 days post PH [86]. In addition, oxi-
holic liver disease, non‐alcoholic fatty liver disease (NAFLD), dative stress is associated with liver regeneration, and this could
viral hepatitis, liver cancer, and hemochromatosis [50, 51]. be caused, at least in part, by regeneration‐associated OXPHOS
Mitochondrial dysfunction can manifest in several ways, [56]. Collectively, these reports highlight the heavy demand for
Table 47.1  The nexus of hepatic polyploidy and mitochondrial metabolism
Gene Model Hepatic ploidy status Relevance to mitochondrial metabolism
E2f7, E2f8 Deletion via Reduced polyploidy The E2F family is known to regulate cell proliferation genes where E2F7 and E2F8 cooperate to repress E2F1 expression and function [62].
knockout [60, 61] 1. E2F1 regulates metabolism in organs such as liver, muscle, pancreas, and adipose tissue that determine metabolic homeostasis. Target
E2f1 Deletion via Increased polyploidy genes of E2F1 are responsible for lipid synthesis (FAS), oxidative metabolism (TOP1MT, EVOVL2, NANOG), and glycolysis (PFKB,
knockout [61] Sirt6, PDK) [63]
2. E2F7 and E2F8‐mediated regulation of E2F1 also occurs in metabolic scenarios such as oxidative stress induced DNA damage response
[64]
Mir122 Knockout Reduced polyploidy Mir122 deficiency has been reported in diseases with dysregulated mitochondrial metabolism such as HCC [65] and NAFLD.
[11] 1. Mice lacking miR‐122 have a severe defect in lipid metabolism [66]
2. miR‐122 regulates the expression of PGC1α, which is a master regulator of mitochondrial biogenesis [67]
3. Other targets of miR‐122, such as LCMT1, PPP1CC, ATF4, MEKK3, and MAPKAP2, are believed to regulate the expression of PGC1α
[67]
4. Another direct target of miR‐122, pyruvate kinase (PKM2) [68], is a key regulator of glycolysis and, thus, can influence mitochondrial
metabolism
Myc Overexpression Increased polyploidy Myc provides the clearest example of programmed expansion of mitochondrial content linked to the cell cycle [69]
via transgenic [58] 1. Myc regulates the expression of TFAM [70], a major determinant of mitochondrial DNA replication and maintenance, as well as
mouse PGC‐1α and PGC‐1β, which are prime regulators of mitochondrial mass and energy metabolism
Deletion via Reduced polyploidy 2. Myc targets include genes involved in bi‐genomic and mitochondrial transcription, mitochondrial translation, protein import, and ETC
knockout [58] complex assembly [71]
3. Myc also indirectly regulates mitochondrial gene expression via repression of microRNAs controlling nuclear genes encoding
mitochondrial proteins (e.g. miR‐23a/b and glutaminase) [72, 73]
Trp53 Deletion via Increased polyploidy 1. p53 transcriptionally represses glucose transporters GLUT1 and GLUT4 along with the insulin receptor (IR) to inhibit cellular glucose
knockout [58] uptake into the cell [74]
2. Through transcriptional activation of TP53‐induced glycolysis and apoptosis regulator (TIGAR), p53 decreases the rate of glycolysis
and redirects glycolytic intermediates into the pentose phosphate pathway (PPP) [74]
3. Glycolysis is also dampened by negative regulation of phosphor‐glycerate‐mutase (PGM) by p53. Transcriptional activation of
hexokinase II (HK II) by mutant p53 stimulates glycolysis [74]
4. p53 promotes OXPHOS through transcriptional activation of synthesis of cytochrome c oxidase 2 (SCO2), a regulator of complex IV,
and apoptosis‐inducing factor (AIF), which acts directly on complex I. By regulating transcription and stability of ribonucleotide
reductase subunit (p53R2), p53 maintains mitochondrial homeostasis and mitochondrial genome integrity [74]
5. p53 is able to transcriptionally regulate and interact with the nuclear encoded mitochondrial transcription factor A (TFAM) and plays a
role in mitochondrial DNA (mtDNA) transcription and in regulating mtDNA content [74]
6. Genotoxic stress signals trigger cytoplasmic p53 to undergo MDM2‐dependent mono‐ubiquitination that induces translocation of p53 to
the mitochondria, which can trigger mitochondrial dependent apoptosis [74]
7. Mitochondrial p53 has been demonstrated to block the antioxidant function of manganese superoxide dismutase (MnSOD), thus creating
a state of oxidative stress [74]
Rb1 Deletion via Increased polyploidy 1. Proteomic effects of Rb1 ablation reveal a striking decrease in multiple mitochondrial proteins [75]
knockout [58] 2. Rb1‐deficient cells have decreased mitochondrial mass and reduced OXPHOS activity [75]
Skp2 Deletion via Increased polyploidy One of the prime targets of Skp2 ubiquitination‐mediated degradation is mitochondrial sirtuin 3 (SIRT3) [76], a histone NAD+ deactylase
knockout [58] 1. SIRT3 controls the flow of mitochondrial oxidative pathways and, consequently, the rate of production of reactive oxygen species [77]
2. SIRT3 controls energy demand during stress conditions (fasting and exercise) and has the ability to quench reactive oxygen species [77]
3. SIRT3 targets enzymes involved in energy metabolism processes, including the respiratory chain, tricarboxylic acid cycle, fatty acid
β‐oxidation and ketogenesis [77]
Birc5 Deletion via Increased polyploidy Proteomic analysis of Birc5‐deleted livers reveal overexpression of mTOR [78], which is a key player in the mitochondrial metabolism
(Survivin) knockout [58] [79]. This axis of BIRC5 and mTOR can possibly act as a node between ploidy and mitochondrial metabolism.
1. mTOR responds to a number of metabolic stimuli such as insulin, insulin growth factor, nutrients (amino acids), energy status, and
oxygen levels and, thus, regulates proliferation [80]
2. mTOR also inhibits mitochondrial autophagy [81].
Independent of mTOR, BIRC5 regulates mitochondrial fission and complex I activity in neuroblastoma cell lines [82], suggesting that it
might be involved in determining mitochondrial metabolism in the liver as well
Ercc1 Deletion via Increased polyploidy Ercc1 deficiency leads to increased oxidative damage as well as decreased levels of genes that are responsible for oxidative metabolism in
knockout [58] the liver [83].

c47.indd 608 03-11-2019 19:52:11


47:  Polyploidy in Liver Function, Mitochondrial Metabolism, and Cancer 609

energy during liver regeneration and underscore an essential role diseased, HCC tissue became predominately diploid [41, 95,
for mitochondrial metabolism in this process [87]. Additional 96]. It is unclear if this shift from polyploid to diploid cells after
studies are required to elucidate the mechanisms that connect malignant HCC transformation is a cause or effect in pathogen-
liver regeneration with mitochondrial metabolism and ploidy. esis, but a potential explanation for the enrichment of diploid
hepatocytes in HCCs is that polyploid hepatocytes protect from
oncogenic transformation. This concept may seem counterin-
tuitive, however, because polyploidy is traditionally associated
PLOIDY AND LIVER CANCER with increased disease severity in other cancers [92].
Recent studies support the idea that polyploidy plays a protec-
Polyploidy has been considered a hallmark of cancer for nearly tive role in the liver. Using a tunable system for altering hepatic
a century [88–90]. In the liver, however, some have suggested ploidy in vivo, Zhu and colleagues showed that Anln deficient
that polyploidy serves as a protective mechanism by providing mice with highly polyploid livers were protected from HCC in
multiple copies of tumor suppressor genes in the event that one the DEN tumor initiation model, whereas E2f8 deficient mice
is damaged [91]. Many forms of cancer exhibit polyploidy at with predominately diploid livers were predisposed to HCC for-
both early and late stages of disease (e.g. renal cell carcinoma, mation [97]. Most importantly, these studies revealed that the
myeloid leukemia, glioblastoma, pre‐esophageal cancer, and polyploid state protects against transformation by providing
lung cancer [92]), but in the context of HCC a reduction of extra tumor suppressor alleles. With two homologous chromo-
ploidy has been observed. Although HCC is associated with somes per cell, diploid hepatocytes have two alleles of each
chromosomal variations [93], the relationship between poly- tumor suppressor. If one tumor suppressor allele is mutated and
ploidy, HCC development and progression, and HCC drug inactivated, the cell is protected by only one functional allele.
resistance is poorly understood. However, the cell is highly susceptible to tumorigenesis if the
second allele is subsequently inactivated. In contrast, polyploid
hepatocytes have four or more homologous chromosomes and a
Polyploidy and liver cancer
corresponding number of tumor suppressor alleles. When a sin-
HCC is the most common form of liver cancer and is associated gle tumor suppressor is inactivated in a polyploid hepatocyte, the
with high mortality and morbidity worldwide [94]. Studies con- cell is protected or “buffered” by three or more functional alleles.
ducted in the 1990s and early 2000s of patients with HCC Thus, even if a second mutation event inactivates a remaining
revealed HCC tumors contained more diploid cells than poly- allele, the cell still retains a minimum of two functional alleles
ploids [94] (Figure 47.3a). Consistent with these observations, and can maintain tumor suppressor activity. Taken together, stud-
work conducted on rats exposed to diethylnitrosamine (DEN) or ies in patients and rodent models strongly support the notion that
2‐acetyl‐aminofluorene indicated that the cells within the diploid hepatocytes are more susceptible to oncogenic

Figure 47.3  Putative role of hepatic ploidy populations in liver cancer. Diploid hepatocytes are enriched in human HCC and are speculated to
have enhanced proliferative potential. Polyploid hepatocytes are reduced in HCC and may have reduced proliferative potential. Recent studies
indicate that polyploid hepatocytes protect against oncogenesis induced by tumor suppressor loss/inactivation; wild‐type copies of the tumor sup-
pressor found on additional chromosomes serve as “back‐ups” that are capable of restricting oncogenic proliferation.
610 THE LIVER:  CONCLUSION

transformation, while polyploid hepatocytes provide a protective Previous work from Michalopoulos and colleagues found
mechanism against HCC formation (Figure 47.3a). 461 copy number variations (CNVs) in 98 human HCC sam-
ples, demonstrating that chromosomal variations are common in
liver cancer [93]. The CNVs identified in this study contained
Polyploidy and drug resistance in cancer discrete genomic amplifications and deletions but not gains or
Increased polyploidy in cancers is associated with poor prognosis losses of entire chromosomes. While aneuploidy in patient HCC
primarily driven by an association with spontaneous drug resist- tumors is not well studied, hepatoblastoma tumors have been
ance [92]. In colon cancer cell lines, for instance, polyploidy is extensively characterized and shown to contain near‐diploid or
attributed to increased resistance to DNA damage caused by near‐tetraploid cells that have gained or lost entire chromo-
camptothecin, cisplatin, oxaliplatin, gamma irradiation, and ultra- somes [102]. Additionally, underlying pathologies of hepatocel-
violet irradiation [98, 99]. Observations supporting a link between lular carcinogenesis, such as chronic hepatitis B virus infection,
liver ploidy and drug sensitivity have also been made in in vivo aflatoxin exposure, and oxidative damage, are associated with
studies. For example, the livers of Mir122 liver‐specific knockout chromosomal variations [103, 104].
mice are predominantly diploid, and when challenged with aceta- It is largely accepted that increased aneuploidy, or chromo-
minophen, Mir122 knockouts are more susceptible to liver dam- some instability (CIN), associates with worse disease prognosis
age compared to wild‐type mice with polyploid hepatocytes [100]. [105]. Unexpectedly, patients with lung, breast, ovarian, and
However, the data are difficult to interpret as Mir122 knockout gastric cancers with the highest degree of CIN were reported to
mice express elevated levels of cytochrome P450 family 2 subfam- have better outcomes than patients with moderate CIN [105].
ily E member 1 (CYP2E1) and cytochrome P450 family 1 sub- One potential explanation is that high‐degree CIN leads to
family A member 2 (CYP1A2), which metabolize acetaminophen mitotic catastrophe and cell death, which effectively kills tumors
to its toxic byproduct N‐acetyl‐p‐benzoquinone imine (NAPQI) with the most aneuploidy. It is difficult to determine if aneu-
[100]. These findings indicate that the relationship between ploidy ploidy promotes HCC cancer cell survival because experimen-
and drug resistance in the liver needs to be further investigated. tally‐induced aneuploidy and chromosomal rearrangements
Chemoresistance is a major concern as it may lead to inappro- have produced ambiguous and conflicting results [105]. To
priate treatment regimens, poor disease prognosis, and decreased study aneuploidy’s role in HCC development and progression in
quality of life for patients with liver cancer. In the clinic, the vitro and in vivo, researchers have used genotoxic agents or spe-
multi‐tyrosine kinase inhibitor sorafenib and the topoisomerase II cifically tailored genetic models [105]. Recent work suggests
inhibitor doxorubicin are used to treat patients with unresectable that the relationship between aneuploidy and HCC can be stud-
HCC [101]. Unfortunately, chemotherapeutic intervention is min- ied by modulating miR‐122 levels in vitro and in vivo because
imally effective and frequently leads to tumors adapting and Mir122 deficiency results in an increase in mononucleate hepat-
becoming refractory to the agent(s) administered [101]. The ocytes with a concomitant reduction in aneuploidy [19]. To
mechanisms by which HCC becomes chemoresistant are poorly effectively investigate if aneuploidy contributes to HCC devel-
understood, but in other cancers, polyploidy has been associated opment, the ability to manipulate ploidy in cancer models is
with drug resistance [92]. The shift in human HCC from poly- critical. Ultimately, the development of innovative strategies
ploid to diploid cell predominance within the diseased tissue indi- that can evade chemoresistance and eliminate tumors will
cates that chemoresistance in liver cancer may not be derived depend upon determining the relationship between aneuploidy,
from polyploidy but rather by other mechanisms. Overall, further polyploidy, and cancer cell survival.
studies are necessary to determine how polyploidy affects disease
severity or drug resistance, which may lead to the development of
novel, innovative, and effective therapeutics.
CONCLUSION
Aneuploidy and liver cancer In adults, the majority of hepatocytes in the liver are polyploid.
Aneuploidy and chromosomal mis‐segregations have been This suggests that polyploid cells are critical to the function and
linked to cancer and neoplastic transformation for almost one homeostasis of the liver. The accumulation of polyploid hepato-
hundred years [88]. Poor disease prognosis and spontaneous cytes with age suggests they may serve as a reservoir for prolifera-
drug resistance are strongly linked to aneuploidy and chromo- tion and adaptation in the case of damage and injury. Recent
somal instability as these events can contribute to cancer cell studies have shown that polyploid hepatocytes act as a critical
survival (Figure 47.3b) [92]. reserve for cells of lower ploidy states, which can be generated by
Although polyploidy and aneuploidy involve chromosomal ploidy reversal. Also, polyploid hepatocytes appear to be more
alterations, it is critical to understand the difference between the resistant to toxic liver injury and cancer, and they may play a dis-
two when considering disease. Polyploidy is a whole number crete role in liver metabolism and regeneration. The molecular
increase in complete chromosome sets in a euploid cell. Depending mechanisms affecting differential activity of diploid and polyploid
on the context, aneuploidy can include whole chromosome gains hepatocytes are poorly understood, and a key question that remains
and losses or structural changes in individual chromosomes. For is whether gene expression in hepatocytes is dictated by ploidy
example, these changes can include complete chromosomal rear- state. A better understanding of hepatic polyploidy will ultimately
rangements or even chromosome segments that are gained or lost. benefit patients in the future by driving the improvement and
In the healthy adult liver of humans and mice, whole chromosome development of innovative strategies that prevent disease, stimu-
gains and losses are a normal feature [13, 17, 18, 22]. late liver regeneration, and advance artificial organ construction.
47:  Polyploidy in Liver Function, Mitochondrial Metabolism, and Cancer 611

ACKNOWLEDGMENTS 22. Faggioli, F., Vezzoni, P., and Montagna, C. Single‐cell analysis of ploidy and
centrosomes underscores the peculiarity of normal hepatocytes. PloS One,
This work was supported by grants to AWD from the NIH (R01 2011;6(10):e26080.
23. Duncan, A.W., Hanlon Newell, A.E., Bi, W. et al. Aneuploidy as a mechanism
DK103645), the Commonwealth of Pennsylvania, and the for stress‐induced liver adaptation. J Clin Invest, 2012;122(9):3307–15.
University of Pittsburgh Physicians Academic Foundation and 24. Galitski, T., Saldanha, A.J., Styles, C.A., Lander, E.S., and Fink, G.R. Ploidy
to ECS from the NIH (F31 DK112633) and the NIBIB training regulation of gene expression. Science, 1999;285(5425):251–4.
grant (T32 EB001026) entitled “Cellular Approaches to Tissue 25. Raslova, H., Kauffmann, A., Sekkai, D. et  al. Interrelation between poly-
ploidization and megakaryocyte differentiation: a gene profiling approach.
Engineering and Regeneration.” Thanks to Frances Alencastro
Blood, 2007;109(8):3225–34.
for helpful discussions and critical reading of the manuscript. 26. Corbel, C., Diabangouaya, P., Gendrel, A.V., Chow, J.C., and Heard, E.
The authors of this chapter apologize to the authors whose Unusual chromatin status and organization of the inactive X chromosome in
works were not cited because of space restrictions. murine trophoblast giant cells. Development, 2013;140(4):861–72.
27. Anatskaya, O.V. and Vinogradov, A.E. Genome multiplication as adaptation
to tissue survival: evidence from gene expression in mammalian heart and
liver. Genomics, 2007;89(1):70–80.
28. Lu, P., Prost, S., Caldwell, H., Tugwood, J.D., Betton, G.R., and Harrison,
REFERENCES D.J. Microarray analysis of gene expression of mouse hepatocytes of differ-
ent ploidy. Mamm Genom, 2007;18(9):617–26.
1. Orr‐Weaver, T.L. When bigger is better: the role of polyploidy in organogen- 29. Rajvanshi, P., Liu, D., Ott, M., Gagandeep, S., Schilsky, M.L., and Gupta, S.
esis. Trends Genet, 2015;31(6):307–15. Fractionation of rat hepatocyte subpopulations with varying metabolic
2. Duncan, A.W. Aneuploidy, polyploidy and ploidy reversal in the liver. Semin potential, proliferative capacity, and retroviral gene transfer efficiency. Exp
Cell Dev Biol, 2013;24(4):347–56. Cell Res, 1998;244(2):405–19.
3. Abmayr, S.M. and Pavlath, G.K. Myoblast fusion: lessons from flies and 30. Pavelka, N., Rancati, G., Zhu, J. et al. Aneuploidy confers quantitative pro-
mice. Development, 2012;139(4):641–56. teome changes and phenotypic variation in budding yeast. Nature,
4. Helming, L. and Gordon, S. Molecular mediators of macrophage fusion. 2010;468(7321):321–5.
Trends Cell Biol, 2009;19(10):514–22. 31. Rancati, G., Pavelka, N., Fleharty, B. et al. Aneuploidy underlies rapid adap-
5. Edgar, B.A. and Orr‐Weaver, T.L. Endoreplication cell cycles: more for less. tive evolution of yeast cells deprived of a conserved cytokinesis motor. Cell,
Cell, 2001;105(3):297–306. 2008;135(5):879–93.
6. Lee, H.O., Davidson, J.M., and Duronio, R.J. Endoreplication: polyploidy 32. Grompe, M., Lindstedt, S., al‐Dhalimy, M. et al. Pharmacological correction
with purpose. Genes Dev, 2009;23(21):2461–77. of neonatal lethal hepatic dysfunction in a murine model of hereditary tyrosi-
7. Engel, F.B., Schebesta, M., and Keating, M.T. Anillin localization defect in naemia type I. Nat Genet, 1995;10(4):453–60.
cardiomyocyte binucleation. J Mol Cell Cardiol, 2006;41(4):601–12. 33. Manning, K., Al‐Dhalimy, M., Finegold, M., and Grompe, M. In vivo sup-
8. Margall‐Ducos, G., Celton‐Morizur, S., Couton, D., Bregerie, O., and pressor mutations correct a murine model of hereditary tyrosinemia type I.
Desdouets, C. Liver tetraploidization is controlled by a new process of Proc Natl Acad Sci USA, 1999;96(21):11928–33.
incomplete cytokinesis. J Cell Sci, 2007;120(20):3633–9. 34. Gupta, S. Hepatic polyploidy and liver growth control. Semin Cancer Biol,
9. Bohm, N. and Noltemeyer, N. Development of binuclearity and DNA‐poly- 2000;10(3):161–71.
ploidization in the growing mouse liver. Histochemistry, 1981;72(1):55–61. 35. Fausto, N. and Campbell, J.S. The role of hepatocytes and oval cells in liver
10. Celton‐Morizur, S., Merlen, G., Couton, D., and Desdouets, C. Polyploidy regeneration and repopulation. Mech Dev, 2003;120(1):117–30.
and liver proliferation: central role of insulin signaling. Cell Cycle, 36. Miyaoka, Y., Ebato, K., Kato, H., Arakawa, S., Shimizu, S., and Miyajima,
2010;9(3):460–6. A. Hypertrophy and unconventional cell division of hepatocytes underlie
11. Wang, X., Montini, E., Al‐Dhalimy, M., Lagasse, E., Finegold, M., and liver regeneration. Curr Biol, 2012;22(13):1166–75.
Grompe, M. Kinetics of liver repopulation after bone marrow transplanta- 37. Sigal, S.H., Rajvanshi, P., Gorla, G.R. et  al. Partial hepatectomy‐induced
tion. Am J Pathol, 2002;161(2):565–74. polyploidy attenuates hepatocyte replication and activates cell aging events.
12. Wilson, J.W. and Leduc, E.H. The occurrence and formation of binucleate Am J Physiol, 1999;276(5 Pt 1):G1260–72.
and multinucleate cells and polyploid nuclei in the mouse liver. Am J Anat, 38. Gandillet, A., Alexandre, E., Royer, C., Cinqualbre, J., Jaeck, D., and
1948;82(3):353–91. Richert, L. Hepatocyte ploidy in regenerating livers after partial hepatec-
13. Duncan, A.W., Hanlon Newell, A.E., Smith, L. et al. Frequent aneuploidy tomy, drug‐induced necrosis, and cirrhosis. Eur Surg Res, 2003;35(3):
among normal human hepatocytes. Gastroenterology, 2012;142(1):25–8. 148–60.
14. Wang, M.J., Chen, F., Lau, J.T.Y., and Hu, Y.P. Hepatocyte polyploidization 39. Anti, M., Marra, G., Rapaccini, G.L. et  al. DNA ploidy pattern in human
and its association with pathophysiological processes. Cell Death Dis, chronic liver diseases and hepatic nodular lesions. Flow cytometric analysis
2017;8(5):e2805. on echo‐guided needle liver biopsy. Cancer, 1994;73(2):281–8.
15. Watanabe, T. and Tanaka, Y. Age‐related alterations in the size of human 40. Rua, S., Comino, A., Fruttero, A. et  al. Flow cytometric DNA analysis of
hepatocytes. A study of mononuclear and binucleate cells. Virchows Arch B cirrhotic liver cells in patients with hepatocellular carcinoma can provide a
Cell Pathol Incl Mol Pathol, 1982;39(1):9–20. new prognostic factor. Cancer, 1996;78(6):1195–202.
16. Celton‐Morizur, S., Merlen, G., Couton, D., Margall‐Ducos, G., and 41. Schwarze, P.E., Pettersen, E.O., Shoaib, M.C., and Seglen, P.O. Emergence
Desdouets, C. The insulin/Akt pathway controls a specific cell division pro- of a population of small, diploid hepatocytes during hepatocarcinogenesis.
gram that leads to generation of binucleated tetraploid liver cells in rodents. Carcinogenesis, 1984;5(10):1267–75.
J Clin Invest, 2009;119(7):1880–7. 42. Baker, D.J., Childs, B.G., Durik, M. et al. Naturally occurring p16(Ink4a)‐
17. Duncan, A.W., Taylor, M.H., Hickey, R.D. et al. The ploidy conveyor of mature positive cells shorten healthy lifespan. Nature, 2016;530(7589):184–9.
hepatocytes as a source of genetic variation. Nature, 2010;467(7316):707–10. 43. Sturmlechner, I., Durik, M., Sieben, C.J., Baker, D.J., and van Deursen, J.M.
18. Knouse, K.A., Wu, J., Whittaker, C.A., and Amon, A. Single cell sequencing Cellular senescence in renal ageing and disease. Nat Rev Nephrol,
reveals low levels of aneuploidy across mammalian tissues. Proc Natl Acad 2017;13(2):77–89.
Sci USA, 2014;111(37):13409–14. 44. Wang, M.J., Chen, F., Li, J.X. et al. Reversal of hepatocyte senescence after
19. Hsu, S.H., Delgado, E.R., Otero, P.A. et al. MicroRNA‐122 regulates poly- continuous in vivo cell proliferation. Hepatology, 2014;60(1):349–61.
ploidization in the murine liver. Hepatology, 2016;64(2):599–615. 45. Bucher, N.L., Swaffield, M.N., and Ditroia, J.F. The influence of age upon
20. Kudryavtsev, B.N., Kudryavtseva, M.V., Sakuta, G.A., and Stein, G.I. the incorporation of thymidine‐2‐C14 into the DNA of regenerating rat liver.
Human hepatocyte polyploidization kinetics in the course of life cycle. Cancer Res, 1964;24:509–12.
Virchows Arch B Cell Pathol Incl Mol Pathol, 1993;64(6):387–93. 46. Fry, M., Silber, J., Loeb, L.A., and Martin, G.M. Delayed and reduced cell
21. Duncan, A.W., Hickey, R.D., Paulk, N.K. et al. Ploidy reductions in murine replication and diminishing levels of DNA polymerase‐alpha in regenerating
fusion‐derived hepatocytes. PLoS Genet, 2009;5(2):e1000385. liver of aging mice. J Cell Physiol, 1984;118(3):225–32.
612 THE LIVER:  REFERENCES

47. Timchenko, N.A., Wilde, M., Kosai, K.I. et al. Regenerating livers of old rats 75. Nicolay, B.N., Danielian, P.S., Kottakis, F. et al. Proteomic analysis of pRb
contain high levels of C/EBPalpha that correlate with altered expression of loss highlights a signature of decreased mitochondrial oxidative phospho-
cell cycle associated proteins. Nucleic Acids Res, 1998;26(13):3293–9. rylation. Genes Dev, 2015;29(17):1875–89.
48. Conboy, I.M., Conboy, M.J., Wagers, A.J., Girma, E.R., Weissman, I.L., and 76. Iwahara, T., Bonasio, R., Narendra, V., and Reinberg, D. SIRT3 functions in
Rando, T.A. Rejuvenation of aged progenitor cells by exposure to a young the nucleus in the control of stress‐related gene expression. Mol Cell Biol,
systemic environment. Nature, 2005;433(7027):760–4. 2012;32(24):5022–34.
49. Iakova, P., Awad, S.S., and Timchenko, N.A. Aging reduces proliferative 77. Hirschey, M.D., Shimazu, T., Huang, J.Y., Schwer, B., and Verdin, E.
capacities of liver by switching pathways of C/EBPalpha growth arrest. Cell, SIRT3 regulates mitochondrial protein acetylation and intermediary metab-
2003;113(4):495–506. olism. Cold Spring Harb Symp Quant Biol, 2011;76:267–77.
50. Grattagliano, I., Russmann, S., Diogo, C. et al. Mitochondria in chronic liver 78. Bracht, T., Hagemann, S., Loscha, M. et al. Proteome analysis of a hepato-
disease. Curr Drug Targets, 2011;12(6):879–93. cyte‐specific BIRC5 (survivin)‐knockout mouse model during liver regen-
51. Du, K., Ramachandran, A., and Jaeschke, H. Oxidative stress during aceta- eration. J Proteome Res, 2014;13(6):2771–82.
minophen hepatotoxicity: sources, pathophysiological role and therapeutic 79. Morita, M., Gravel, S.P., Hulea, L. et al. mTOR coordinates protein synthesis,
potential. Redox Biol, 2016;10:148–56. mitochondrial activity and proliferation. Cell Cycle, 2015;14(4):473–80.
52. Lonardo, A., Ballestri, S., Marchesini, G., Angulo, P., and Loria, P. 80. Laplante, M. and Sabatini, D.M. mTOR signaling in growth control and
Nonalcoholic fatty liver disease: a precursor of the metabolic syndrome. Dig disease. Cell, 2012;149(2):274–93.
Liver Dis, 2015;47(3):181–90. 81. Kim, J., Kundu, M., Viollet, B., and Guan, K.L. AMPK and mTOR regulate
53. Cheng, Z., Tseng, Y., and White, M.F. Insulin signaling meets mitochondria autophagy through direct phosphorylation of Ulk1. Nat Cell Biol,
in metabolism. Trends Endocrinol Metab, 2010;21(10):589–98. 2011;13(2):132–41.
54. Scarpulla, R.C. Metabolic control of mitochondrial biogenesis through the 82. Hagenbuchner, J., Kuznetsov, A.V., Obexer, P., and Ausserlechner, M.J.
PGC‐1 family regulatory network. Biochim Biophys Acta, 2011;1813(7): BIRC5/Survivin enhances aerobic glycolysis and drug resistance by altered
1269–78. regulation of the mitochondrial fusion/fission machinery. Oncogene,
55. Gentric, G., Maillet, V., Paradis, V. et al. Oxidative stress promotes patho- 2013;32(40):4748–57.
logic polyploidization in nonalcoholic fatty liver disease. J Clin Invest, 83. Gregg, S.Q., Gutierrez, V., Robinson, A.R. et al. A mouse model of acceler-
2015;125(3):981–92. ated liver aging caused by a defect in DNA repair. Hepatology,
56. Gorla, G.R., Malhi, H., and Gupta, S. Polyploidy associated with oxidative 2012;55(2):609–21.
injury attenuates proliferative potential of cells. J Cell Sci, 2001;114(16): 84. Guerrieri, F., Muolo, L., Cocco, T. et  al. Correlation between rat liver
2943–51. regeneration and mitochondrial energy metabolism. Biochim Biophys Acta,
57. Troadec, M.B., Courselaud, B., Detivaud, L. et al. Iron overload promotes 1995;1272(2):95–100.
Cyclin D1 expression and alters cell cycle in mouse hepatocytes. J Hepatol, 85. Sun, Q., Miao, M., Jia, X. et al. Subproteomic analysis of the mitochondrial
2006;44(2):391–9. proteins in rats 24 h after partial hepatectomy. J Cell Biochem,
58. Duncan, A.W. Changes in hepatocyte ploidy during liver regeneration, in 2008;105(1):176–84.
Liver Regeneration: Basic Mechanisms, Relevant Models and Clinical 86. Gear, A.R. Observations on iron uptake, iron metabolism, cytochrome c
Applications, (ed. U. Apte) Elsevier, 2015. content, cytochrome a content and cytochrome c‐oxidase activity in regen-
59. Pandit, S.K., Westendorp, B., Nantasanti, S. et al. E2F8 is essential for poly- erating rat liver. Biochem J, 1965;97(2):532–9.
ploidization in mammalian cells. Nat Cell Biol, 2012;14(11):1181–91. 87. Huang, J. and Rudnick, D.A. Elucidating the metabolic regulation of liver
60. Kent, L.N., Rakijas, J.B., Pandit, S.K. et al. E2f8 mediates tumor suppression regeneration. Am J Pathol, 2014;184(2):309–21.
in postnatal liver development. J Clin Invest, 2016;126(8):2955–69. 88. Boveri, T. Zur Frage der Entstehung maligner Tumoren (The Origin of
61. Pandit, S.K., Westendorp, B., and de Bruin, A. Physiological significance of Malignant Tumors), Jena, Gustav Fischer, 1914.
polyploidization in mammalian cells. Trends Cell Biol, 89. Wolff, J. The Theory of Cancer, Science History Publishers, Sagamore
2013;23(11):556–66. Beach, 1907.
62. Moon, N.S. and Dyson, N. E2F7 and E2F8 keep the E2F family in balance. 90. Duensing, A. and Duensing, S. Centrosomes, polyploidy and cancer, in
Dev Cell, 2008;14(1):1–3. Polyploidization and Cancer, (ed. R.Y.C. Poon), Springer, New York, 2010,
63. Denechaud, P.D., Fajas, L., and Giralt, A. E2F1, a novel regulator of metabo- pp. 93–103.
lism. Front Endocrinol, 2017;8:311. 91. Nowell, P.C. The clonal evolution of tumor cell populations. Science,
64. Zalmas, L.P., Zhao, X., Graham, A.L. et al. DNA‐damage response control 1976;194(4260):23–8.
of E2F7 and E2F8. EMBO Rep, 2008;9(3):252–9. 92. Merlo, L.M., Wang, L.S., Pepper, J.W., Rabinovitch, P.S., and Maley, C.C.
65. Kutay, H., Bai, S., Datta, J. et al. Downregulation of miR‐122 in the rodent Polyploidy, aneuploidy and the evolution of cancer. Adv Exp Med Biol,
and human hepatocellular carcinomas. J Cell Biochem, 2006;99(3):671–8. 2010;676:1–13.
66. Hsu, S.H., Wang, B., Kota, J. et al. Essential metabolic, anti‐inflammatory, 93. Nalesnik, M.A., Tseng, G., Ding, Y. et al. Gene deletions and amplifications
and anti‐tumorigenic functions of miR‐122 in liver. J Clin Invest, in human hepatocellular carcinomas: correlation with hepatocyte growth
2012;122(8):2871–83. regulation. Am J Pathol, 2012;180(4):1495–508.
67. Burchard, J., Zhang, C., Liu, A.M. et  al. microRNA‐122 as a regulator of 94. Nagasue, N., Kohno, H., Chang, Y.C. et al. DNA ploidy pattern in synchro-
mitochondrial metabolic gene network in hepatocellular carcinoma. Mol Syst nous and metachronous hepatocellular carcinomas. J Hepatol, 1992;
Biol, 2010;6:402. 16(1–2):208–14.
68. Liu, A.M., Xu, Z., Shek, F.H. et al. miR‐122 targets pyruvate kinase M2 and 95. Saeter, G., Schwarze, P.E., Nesland, J.M., Juul, N., Pettersen, E.O., and
affects metabolism of hepatocellular carcinoma. PloS One, 2014;9(1):e86872. Seglen, P.O. The polyploidizing growth pattern of normal rat liver is
69. Morrish, F. and Hockenbery, D. MYC and mitochondrial biogenesis. Cold replaced by divisional, diploid growth in hepatocellular nodules and carci-
Spring Harb Perspect Med, 2014;4(5). nomas. Carcinogenesis, 1988;9(6):939–45.
70. Ahuja, P., Zhao, P., Angelis, E. et al. Myc controls transcriptional regulation 96. Schwarze, P.E., Saeter, G., Armstrong, D. et al. Diploid growth pattern of
of cardiac metabolism and mitochondrial biogenesis in response to patho- hepatocellular tumours induced by various carcinogenic treatments.
logical stress in mice. J Clin Invest, 2010;120(5):1494–505. Carcinogenesis, 1991;12(2):325–7.
71. Li, F., Wang, Y., Zeller, K.I. et al. Myc stimulates nuclearly encoded mito- 97. Zhang, S., Zhou, K., Luo, X. et al. The polyploid state plays a tumor‐sup-
chondrial genes and mitochondrial biogenesis. Mol Cell Biol, 2005; pressive role in the liver. Dev Cell, 2018;44(4):447–59 e5.
25(14):6225–34. 98. Castedo, M., Coquelle, A., Vitale, I. et al. Selective resistance of tetraploid
72. Dang, C.V., Le, A., and Gao, P. MYC‐induced cancer cell energy metabolism cancer cells against DNA damage‐induced apoptosis. Ann N Y Acad Sci,
and therapeutic opportunities. Clin Cancer Res, 2009;15(21):6479–83. 2006;1090:35–49.
73. Psathas, J.N. and Thomas‐Tikhonenko, A. MYC and the art of microRNA 99. Castedo, M., Coquelle, A., Vivet, S. et al. Apoptosis regulation in tetraploid
maintenance. Cold Spring Harb Perspect Med, 2014;4(8). cancer cells. EMBO J, 2006;25(11):2584–95.
74. Itahana, Y. and Itahana, K. Emerging roles of p53 family members in glucose 100. Chowdhary, V., Teng, K.Y., Thakral, S. et al. miRNA‐122 protects mice and
metabolism. Int J Mol Sci, 2018;19(3). human hepatocytes from acetaminophen toxicity by regulating cytochrome
47:  Polyploidy in Liver Function, Mitochondrial Metabolism, and Cancer 613

P450 family 1 subfamily A member 2 and family 2 subfamily E member 1 Guinea‐Conakry, West Africa. J Gastroenterol Hepatol, 2002;17
expression. Am J Pathol, 2017;187(12):2758–74. Suppl:S441–8.
101. Nguyen, K., Jack, K., and Sun, W. Hepatocellular carcinoma: past and 104. Slagle, B.L., Zhou, Y.Z., and Butel, J.S. Hepatitis B virus integration event
future of molecular target therapy. Diseases, 2016;4(1):1. in human chromosome 17p near the p53 gene identifies the region of the
102. Weaver, B.A. and Cleveland, D.W. Does aneuploidy cause cancer? Curr chromosome commonly deleted in virus‐positive hepatocellular carcino-
Opin Cell Biol, 2006;18(6):658–67. mas. Cancer Res, 1991;51(1):49–54.
103. Turner, P.C., Sylla, A., Diallo, M.S., Castegnaro, J.J., Hall, A.J., and 105. Zasadil, L.M., Britigan, E.M., and Weaver, B.A. 2n or not 2n: aneuploidy,
Wild, C.P. The role of aflatoxins and hepatitis viruses in the etiopatho- polyploidy and chromosomal instability in primary and tumor cells. Semin
genesis of hepatocellular carcinoma: a basis for primary prevention in Cell Dev Biol, 2013;24(4):370–9.
PART FOUR:
PATHOBIOLOGY OF LIVER
DISEASE
Hepatic Encephalopathy
48 Roger F. Butterworth
Department of Medicine, University of Montreal, Montreal, Canada

INTRODUCTION Further subdivisions then occur as a function of the time course


of HE as depicted in Table 48.1.
Hepatic encephalopathy (HE) is a serious central nervous Episodic HE
­system (CNS) complication of liver diseases that is character- Recurrent HE characterized by bouts of HE that occur within
ized by a range of neurological and neuropsychiatric symptoms a period of six months or less
that include perturbations of psychomotor, neuro‐cognitive, and Permanent HE in which symptoms are unremitting
fine motor function. Deficits in attention, visual perception,
visuo‐spatial construction in addition to motor speed and West Haven criteria for grading
­accuracy are among the early symptoms of HE [1]. These symp- of HE severity
toms are generally considered to be potentially reversible.
The trans‐jugular intrahepatic portosystemic shunt (TIPS) In patients with cirrhosis, clinically symptomatic (overt) HE
procedure is effective for the prevention and treatment of com- (OHE) heralds the decompensated phase of the disease and grad-
plications of cirrhosis such as portal hypertension, gastrointesti- ing of the severity of OHE has traditionally made use of a battery
nal (variceal) bleeding, and refractory ascites. However, the of neuropsychiatric tests known as the West Haven criteria where
procedure may result in new or worsening episodes of encepha- grade I characterized by shortened attention span and sleep dis-
lopathy in up to 50% of patients [2]. orders is followed by changes in personality, disorientation, and
The burden of cirrhosis and HE is multidimensional and is asterixis (grade II). Grade III is typified by disorientation, confu-
increasing. Direct economic costs related to hospitalizations, sion, and semi‐stupor followed by the comatose state (grade IV).
medical procedures, and medications are substantial and to If required, assessment of the depth and duration of coma/
these must be added the economic cost to the patient resulting impaired consciousness can be made using the Glasgow coma
from lost income. Cognitive dysfunction is associated with scale, a points scale based upon motor responsiveness, verbal
worsening financial status, employment prospects, and q­ uality performance, and eye opening in response to stimuli.
of life with consequent burden to the patient’s family and car- Minimal HE (MHE), a subtype of type C HE in which clini-
egivers [3]. cally‐manifest symptoms are absent, occurs in up to 80% of
patients with cirrhosis and has a significant impact on health‐
related quality of life (HRQOL) of the patient [3]. MHE is diag-
nosed by one or more of a series of well‐established psychometric
NEW CLASSIFICATION tests. MHE is associated with increased numbers of falls and
OF HE SYNDROMES road traffic accidents [3, 4].
In recent times, largely due to the inherent subjective nature
The current practice guidelines provide a system of classifica- of the neuropsychiatric symptoms defining grade I HE by
tion according to the nature of the underlying disease. In this West Haven, efforts were made to modify the nomenclature
way, HE is subdivided into three major types: and classification of HE. In the resulting AASLD/EASL
Type A: HE associated with acute liver failure guidelines, a new term pertinent to HE classification, namely
Type B: HE associated with portosystemic bypass/shunting covert HE (CHE), has appeared defined as MHE together
Type C: HE associated with cirrhosis with what was formerly grade I HE [4]. This change was

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
618 THE LIVER:  NEUROPATHOLOGY

accompanied by a streamlining of the diagnosis of overt HE may have prognostic value for prediction of the severe neurologi-
into grades II, III, and IV as shown in Table 48.1. cal complications of acute liver failure (ALF) [7].
Grading of HE in patients with cirrhosis using West Haven
criteria has been criticized due to its inherent subjective nature
with the likelihood to lead to high inter‐examiner variability.
CLINICAL PRESENTATION One alternative technique that is gaining in popularity makes
AND DIAGNOSIS use of critical flicker frequency (CFF), a quantitative measure
based upon the correlation between cerebral processing of oscil-
HE is a diagnosis of exclusion of other metabolic disorders, latory visual stimuli impairment and increasing severity of HE.
infectious diseases, intracranial vascular events, and intracranial The CFF procedure is used extensively for the grading of neuro-
space‐occupying lesions all of which may potentially result in cognitive changes in a variety of neurological disorders that
neuropsychiatric symptomatology [5]. Covert HE is accompa- includes low‐grade HE in cirrhosis where a CFF cut‐off of 39
nied by alterations of performance in psychometric tests directed Hz differentiates low‐grade overt HE from non‐HE patients [8].
toward attention, working memory, psychomotor speed, and
visuospatial ability. Overt HE is considered to start with asterixis
or flapping tremor [6] and staging of severity in OHE by West
Haven. Identification of well‐established precipitating factors NEUROPATHOLOGY
for HE such as infection, gastrointestinal bleeding, and consti-
pation may support the diagnosis of HE in cirrhosis. For patients Glial pathology
with an evidently altered state of consciousness, the Glasgow
coma scale is widely employed. Glial cells manifest the most conspicuous and reproducible
Despite strong evidence for a role of ammonia in the pathogen- morphological and functional changes related to HE in both
esis of HE, measurements of venous ammonia add little informa- ALF and in cirrhosis. However, the nature and extent of these
tion by way of diagnostic or prognostic value in patients with changes are determined by the type of liver injury (acute versus
cirrhosis. On the other hand, arterial ammonia concentrations chronic), the extent of portal‐systemic shunting, and the number
and duration of the episodes of HE. Glial pathology in HE
involves one of two major types of cells, namely astrocytes and
Table 48.1  Updated nomenclature of hepatic encephalopathy (HE) microglia.
syndromes is dependent upon the type of HE (A, B, or C), the grade
of HE (minimal, I, II, III, IV) that is now subdivided into covert Astrocytes
(a combination of MHE and grade I overt) and overt (II, III, and IV).
Depending upon the time course, HE is then rated as episodic, Pathological evaluation of brain sections from patients with
recurrent, or permanent and may fall into one of two categories end‐stage ALF reveals cytotoxic brain edema as shown in
namely spontaneous or precipitated Figure 48.1a where the astrocyte manifests massive swelling of
Type HE grade Time course an astrocyte end foot (A) [9]. The astrocyte swelling has the
Minimal potential to result in the development of intracranial hyperten-
A I Covert Episodic sion leading ultimately to brain herniation which remains one of
B II Recurrent the major causes of mortality in ALF.
III Overt
C Persistent In contrast to ALF, the cardinal neuropathological feature
IV
of decompensated cirrhosis is a characteristic morphological

(a) (b)

Figure 48.1  Characteristic ultrastructural changes in the brain in acute and in chronic liver failure. (a) Electron micrograph from a patient with
acute liver failure due to acetaminophen overdose showing cytotoxic brain edema with marked swelling of perivascular astrocyte (A), and dilatation
of endoplasmic reticulum (arrows) and mitochondria (M). (b) Electron micrograph from a 51‐year‐old cirrhotic patient who died in hepatic coma
presenting Alzheimer type II astrocytes with a large, pale nucleus and margination of chromatin (Alz) compared with normal astrocytes with normal
chromatin pattern (N).
48:  Hepatic Encephalopathy 619

alteration of the astrocyte nucleus known as Alzheimer type 2 to the fact that liver is a key site for thiamine synthesis and stor-
astrocytosis [10]. The Alzheimer type 2 phenotype consists of age in humans and Wernicke’s encephalopathy results solely
nuclear pallor and swelling of both the cell itself and its from thiamine deficiency.
nucleus. Margination of the chromatin pattern and deposition
of glycogen is also characteristic of the Alzheimer type 2 Cerebellar degeneration
change as depicted in Figure  48.1b. Such changes are une-
Mild to severe cerebellar degeneration characterized by lesions
venly distributed in brain with particularly high densities
in cerebellar vermis and severe loss of Purkinje cells occurs in
observed in cerebral cortex, basal ganglia, and cerebellum
both alcoholic and non‐alcoholic patients with end‐stage cirrho-
[11]. It is important to note that there is also evidence of brain
sis although the severity of neuropathologic changes is more
edema in cirrhosis although, in contrast to the situation in
severe in the alcoholics [11]. There are no clear correlations
ALF, the edema in cirrhosis is low grade in nature, rarely
between the incidence and extent of cerebellar degeneration,
leading to alterations of intracranial pressure. Low‐grade
Wernicke‐type thalamic lesions or Alzheimer type 2 astrocyto-
edema in cirrhosis has been demonstrated in basal ganglia and
sis suggesting that these neuropathologic phenotypes are under-
corticospinal tract of patients with cirrhosis using magnetic
pinned by distinct mechanisms.
resonance imaging (MRI) where decreases of magnetization
transfer ratios and increases of water apparent diffusion coef-
Post‐shunt myelopathy
ficients were reported [12].
Although rare, myelopathy may occur in patients with cirrhosis
Microglia following multiple episodes of coma or following the TIPS pro-
cedure. Symptoms of spastic paresis or paralysis of the lower
Microglia are the immunomodulatory cells of the brain and
limbs are apparent and result from demyelination of both direct
results of studies in animal models of acute or chronic liver fail-
and crossed corticospinal tracts in this condition.
ure as well as in material from patients with these conditions
confirm that activation of microglia is a common occurrence in
Parkinsonism in cirrhosis
HE [13]. Activation of microglia is indicative of neuroinflamma-
tion and is observed in a wide range of neurodegenerative CNS Characterized by extrapyramidal symptoms (hypokinesia,
disorders including HE. Neuroinflammation associated with tremor, rigidity) the condition is rapidly progressive and may
microglial activation is currently considered to be a feature of occur independent of the extent of cognitive dysfunction. With a
HE in both acute and chronic liver diseases [13,14]. This issue is prevalence as high as 21% in one report of patients with cirrhosis
discussed further in a subsequent section on Neuroinflammation listed for transplantation [16], the disorder has been attributed to
in this chapter. dopaminergic neuronal deficits resulting from manganese depo-
sition in basal ganglia [17]. Extrapyramidal symptoms in these
Neuronal pathology patients may respond to liver transplantation and, in some cases
to L‐DOPA therapy. One useful procedure that is available for
Although the principal neuropathological change typical of HE confirmation of diagnosis of Parkinsonism in cirrhosis ­consists
in cirrhosis is primarily glial in nature, it is important to bear in of T1‐weighted MRI where bilateral signal hyperintensities are
mind that neuronal cell death frequently also occurs. Changes of observed in globus pallidus and substantia nigra. Further discus-
neuronal morphological and functional characteristics are not sion relating to brain manganese deposition in cirrhosis appears
uncommon [15]. These changes occur in distinct brain struc- in the Manganese section of the current chapter.
tures and have been attributed to a spectrum of diverse patho-
physiologic mechanisms. The neuropathological and clinical Global brain reserve
characteristics of these neurodegenerative disorders are summa-
rized in the following sections. A novel concept composed of a structural component (brain
reserve [BR]) together with a patient’s ability to tolerate changes
Acquired non‐Wilsonian hepatocerebral (cognitive reserve), BR is dependent upon disease etiology,
degeneration (ANHD) severity, and progression. Decreased BR examined by magnetic
resonance techniques is related to three factors namely ­structural
ANHD occurs in patients with cirrhosis following a protracted changes in white and gray matter of multiple brain structures,
clinical course often with multiple episodes of coma and evi- brain edema, and altered brain metabolism.
dence of portal‐systemic shunting. The neuropathology consists Results of a multi‐modal MRI study reveal that patients with
of spongiform degeneration in deep cerebrocortical layers as alcoholic cirrhosis have, despite abstinence, a poor BR whereas
well as in basal ganglia structures, cerebellum, and subcortical patients with non‐alcoholic cirrhosis have a greater potential for
white matter. deterioration of BR following HE or TIPS. Enhancement of the
direct effects of chronic alcohol on brain (Wernicke’s encepha-
Wernicke’s encephalopathy
lopathy, cerebellar degeneration) or indirectly via the liver
Unsuspected Wernicke‐type hemorrhagic lesions in thalamus resulting in ANHD could form part of the basis for the structural
occur in up to 30% of patients with end‐stage cirrhosis of alco- changes characteristic of BR. Worsening of BR as a c­ onsequence
holic etiology [11] where lesions may be acute or chronic (long- of alcoholic etiology of cirrhosis (despite abstinence) has the
standing). The higher incidence of Wernicke‐type lesions in potential to modulate the impact of HE on disease p­ rogression
these patients compared to noncirrhotic alcoholics likely relates and suitability for transplantation [18].
620 THE LIVER:  PATHOPHYSIOLOGY

PATHOPHYSIOLOGY exclusively on the synthesis of glutamine and the enzyme


responsible, GS, is preferentially expressed by the astrocyte.
Ammonia
1
H‐magnetic resonance spectroscopic studies of patients with
cirrhosis reveal increased concentrations of brain glutamine as a
Liver is the organ primarily tasked with the body’s removal of function of the severity of HE in these patients [22].
excess ammonia in the form of urea via the urea cycle in peri- Consequently, the accumulation of glutamine in the astrocyte
portal hepatocytes and glutamine via the enzyme glutamine syn- may be causally‐related to the pathogenesis of brain edema in
thetase (GS) localized in perivenous hepatocytes as shown in a ALF. However, studies in experimental animal models with
simplified manner in Figure 48.2. Patients with cirrhosis com- ALF resulting from toxic liver injury failed to demonstrate a
monly develop intra‐ and extra‐hepatic portal‐systemic shunts significant correlation between brain glutamine content, its syn-
and this shunting combined with the loss of metabolic capacity thesis or turnover, and severity of HE or brain edema [23] sug-
of hepatocytes results in impaired removal of ammonia and con- gesting the presence of alternative or additional mechanisms.
sequent hyperammonemia. Established mechanisms responsible for the deleterious
One key adaptive consequence of impairment of hepatic effects of ammonia on brain function include the following:
ammonia removal in cirrhosis is activation of the gene coding
for the enzyme protein GS in skeletal muscle [19]. This meta- Direct effects of the ammonium ion (NH4 +)
bolic adaptation provides an alternate pathway for ammonia on the neuronal membrane
removal in the form of muscle glutamine production as shown
schematically in Figure 48.2. Concentrations of ammonia equivalent to those reported in
Arterial blood and brain ammonia levels are increased sev- brain in acute and/or chronic liver failure are known to exert
eral‐fold in patients with cirrhosis and HE and dynamic 13NH3‐ direct deleterious effects on both inhibitory and excitatory neu-
positron emission tomography (PET) studies demonstrate rotransmission [24] by virtue of the fact that, under normal
significant increases of brain ammonia and of the cerebral meta- physiological conditions, ammonia exists primarily in its proto-
bolic rate for ammonia (CMRA, defined as the rate at which nated form (NH4+) an entity with comparable ionic radius to that
ammonia is captured by brain and incorporated into metabo- of potassium (K+) which is an ion with potent depolarizing
lites) in these patients [20]. These studies went on to demon- properties on the neuronal membrane.
strate that, contrary to earlier reports, this increase of brain
ammonia in patients with cirrhosis was primarily a consequence Effects of ammonia on brain energy metabolism
of increased arterial blood concentrations rather than to altered Ammonia, in low millimolar concentrations equivalent to those
kinetics of blood–brain transfer of ammonia. observed in HE is also a potent inhibitor of the rate‐limiting
Venous ammonia measurements are of little predictive value tricarboxylic acid (TCA) cycle enzyme α‐ketoglutarate dehy-
in terms of HE occurrence or severity in patients with cirrhosis. drogenase [25] with the potential to result in impairment of
In contrast, arterial ammonia concentrations are useful predic- glucose oxidation, brain lactate production, and impending
­
tors of high risk for the development of severe complications of brain energy failure.
brain edema in patients with ALF such as intracranial hyperten-
sion [21] and brain herniation [7] that remains a major cause of
Effects of ammonia on the brain GABA system
mortality in this condition.
The brain is devoid of an effective urea cycle. Consequently, Recent studies show that ammonia is an activator of translocator
removal of excess ammonia by the brain depends almost protein (TLP), a protein located on the mitochondrial membrane

BRAIN BRAIN

urea LIVER urea LIVER

Periportal Periportal
hepatocytes hepatocytes

Perivenous Perivenous
hepatocytes hepatocytes
NH3 GUT NH3 GUT
MUSCLE MUSCLE

GLUTAMINE KIDNEY GLUTAMINE KIDNEY

Normal Liver Failure

Figure 48.2  Schematic representation of inter‐organ trafficking of ammonia and glutamine under normal physiological conditions and in liver
failure. Gut‐derived ammonia is normally removed as urea (periportal hepatocytes) or glutamine (perivenous hepatocytes). In liver failure, ammonia
removal by both types of hepatocytes is decreased. Brain ammonia uptake increases but there is limited capacity for further increases in brain glu-
tamine synthesis. In contrast, skeletal muscle becomes the principal route for ammonia removal due to post‐translational increase in the enzyme
glutamine synthetase.
48:  Hepatic Encephalopathy 621

of glial cells (astrocytes and microglia) where it serves in the SIRS the extent of which increases as a function of the nature
transport of cholesterol. Activation by ammonia results in and severity of liver injury. SIRS results from the release into
increased uptake and conversion to a novel series of compounds the circulation of proinflammatory cytokines such as tumor
known as neurosteroids (NS) one of which, allopregnanolone, is necrosis factor alpha (TNFα), interleukin‐1beta (IL‐1β), and
a potent agonist of the modulatory site on the gamma‐amino interleukin‐6 (IL‐6). Circulating levels of TNFα are invariably
butyric acid (GABA) receptor. In this way, ammonia serves as increased in patients with cirrhosis and the magnitude of the
an activator of GABAergic transmission with the potential to increase correlates well with the grade of OHE [28].
increase chloride flux and neuroinhibition. This mechanism is Systemic inflammation resulting from lipopolysaccharide
covered in greater detail in the section on Inhibitory neurotrans- (LPS, endotoxin) has been shown to precipitate HE and increase
mission of the current chapter. blood–brain barrier (BBB) permeability in mice with ALF
resulting from toxic liver injury [29]. These findings suggest
that both cytotoxic (cell swelling) and vasogenic (BBB break-
Manganese down) mechanisms could contribute to the pathogenesis of
T1‐weighted MRI of patients with cirrhosis consistently reveals brain edema and its CNS complications in ALF.
bilateral symmetric signal hyperintensities in substructures of
the basal ganglia particularly the globus pallidus and substantia
nigra. In a landmark study of 51 patients with cirrhosis who
The concept of synergism
were listed for liver transplantation and were assessed prospec- Interest in synergistic mechanisms as an integral part of the
tively over a one‐year period, 11 patients (21.6%) exhibited pathogenesis of HE in cirrhosis began with the pioneering work
definite Parkinsonism together with typical MRI signal hyperin- of Les Zieve in the 1970s where synergistic actions of estab-
tensities. The degree of signal hyperintensity was not correlated lished liver‐derived toxins including ammonia, methanethiol,
with the etiology of cirrhosis, Child–Pugh scores, or fasting and octanoic acid when administered to laboratory animals
blood ammonia and no patients had OHE at the time of imaging. were shown to act in concert leading to HE [30]. Since that time,
Neurological symptoms included a symmetric akinetic‐rigid evidence of synergism between three of the major entities impli-
syndrome, tremor, stooped posture, and gait impairment with cated in the pathogenesis of HE in cirrhosis namely ammonia,
rapid progression over months [16]. Blood manganese concen- manganese, and proinflammatory cytokines has been reported.
trations were elevated up to sevenfold in all of nine patients in In patients with cirrhosis and documented infection, induc-
whom measurements were made and cerebrospinal fluid (CSF) tion of hyperammonemia is associated with worsening of neu-
manganese concentrations were increased in all of three cases in ropsychiatric status [31] and there is evidence indicating that
which CSF was available. Liver transplantation resulted in nor- synergism between ammonia, manganese, and proinflammatory
malization of MRI signal hyperintensities and circulating man- cytokines represents a cascade of mechanisms resulting in the
ganese levels as well as improved neurological symptoms. worsening of HE [32]. The trail of evidence consists of the
Treatment of two of these patients with L‐DOPA resulted in ­following: glutamate is the principal excitatory neurotransmitter
substantial improvements of motor function [16]. Motor impair- of mammalian brain and the termination of its action relies on
ments including the akinetic‐rigid syndrome that are character- transport into surrounding astrocytes via specific transporters.
istic of Parkinsonism in cirrhosis could contribute to the poor Both ammonia and manganese have been shown to inhibit these
performance in psychometric testing using paper and pencil transporters resulting in an excess of glutamate in the synaptic
tests such as NCT‐A and NCT‐B in which there is a significant cleft leading to hyperexcitability and activation of post‐synaptic
motor component thus casting some doubt on the sole use of glutamate receptors. This cascade results in nitration and conse-
these tests for the assessment of cognitive function and diagno- quent inactivation of key proteins via a process known as pro-
sis of MHE or covert HE in patients with cirrhosis. Alternative tein tyrosine nitration (PTN) [33]. One such tyrosine‐nitrated
tests with less dependency on unimpaired motor performance protein is the enzyme GS that is solely responsible for ammonia
are available. removal by the brain. Under conditions of PTN, brain ammonia
In a separate study, basal ganglia tissue obtained at autopsy levels become increased and the synergistic cycle is amplified.
from patients with cirrhosis who died in hepatic coma were
found to contain several‐fold increased manganese concentra-
tions [26] as well as alterations of marker proteins and metabo-
Neuroinflammation
lite patterns related to the dopamine (DA) neurotransmitter Recent studies provide convincing evidence in support of a
system that are characteristic of idiopathic (non‐cirrhosis‐ novel concept namely the role of neuroinflammation (inflam-
related) Parkinson’s disease [17]. mation of the brain per se) in the pathogenesis of HE in patients
with ALF [34]. Neuroinflammation in ALF is characterized by
activation of microglial cells together with increased production
Systemic inflammation of proinflammatory cytokines such as TNFα and the interleu-
Systemic inflammation resulting from infection and/or hepato- kins IL‐1β and IL‐6 in the brain. The grade of HE severity cor-
cellular damage is a common occurrence in acute or chronic relates well with that of proinflammatory cytokine production
liver failure and the acquisition of a systemic inflammatory and the use of proinflammatory gene deletion strategies results
response (SIRS) is a major predictor of HE in patients with cir- in slowing of HE progression in experimental ALF [35].
rhosis or ALF [27]. Cellular damage to liver tissue resulting Microglial activation has also been reported in autopsied brain
from toxins, particularly ethanol, has the potential to exacerbate tissue samples from ALF patients with grade IV HE [13].
622 THE LIVER:  PATHOPHYSIOLOGY

Microglial activation also occurs in patients with cirrhosis who Cerebral blood flow (CBF)
died in hepatic coma [14].
Activated microglia are known to express transcripts for the Although global CBF in patients with cirrhosis is generally
mitochondrial protein translocator protein (TLP) and using somewhat reduced, PET studies reveal that CBF changes in
PET and the selective TLP ligand 11C‐PK11195, activation these patients with mild HE occurred in a region‐selective
of  microglia indicative of a neuroinflammatory response is manner whereby flow to cerebral cortical regions was
observed in patients with cirrhosis and MHE. Microglial acti- decreased while flow to basal ganglia, cerebellar, and thalamic
vation appears to be a relatively early occurrence in patients structures was significantly increased. Since this pattern of
with decompensated cirrhosis. Particularly intense signals were changes in CBF parallels the regional changes of brain glucose
observed in the anterior cingulate cortex, a brain structure utilization, CBF autoregulation (defined as the capacity of
known to be associated with the control of attention in MHE CBF to match brain activity independent of changes of sys-
patients [36]. temic arterial pressure) appears to be preserved in patients
Liver‐brain proinflammatory signaling occurs in HE and a with cirrhosis [42].
growing list of possible mechanisms have been proposed [34]. In contrast to the situation in cirrhosis, loss of CBF autoreg-
At the cellular level, human cerebrovascular endothelial cells ulation is common in patients with ALF where the variability of
exposed to TNFα manifest increased transport of ammonia [37] CBF rates are influenced by a variety of factors including the
and cultured neural cells exposed to combinations of ammonia complex interplay between local energy demands, ammonia
and recombinant proinflammatory cytokines display increased neurotoxicity, systemic arterial pressure, and intracranial
expression of genes coding for proteins implicated in cellular hypertension. A five‐phase classification system based upon a
dysfunction in HE suggestive of synergism. Factors other than retrospective analysis of sequential intracranial pressure (ICP)
ammonia and manganese that are increased in brain in liver fail- and CBF measurements in ALF patients has been proposed
ure also have the potential to elicit a neuroinflammatory [43] as follows:
response. For example, millimolar concentrations of lactate are Phase 1: An initial phase of low CBF resulting from lower neu-
known to substantially increase the release of TNFα and IL‐6 ronal activity
from cultured microglial cells [38]. Phase 2: A progressive increase of CBF
Phase 3: A subsequent increase of ICP
Brain energy metabolism Phase 4: Final stage where increased ICP results in decreased CBF
Phase 5: Brain death
There is no convincing evidence from either biochemical or
spectroscopic investigations in support of the hypothesis that The classification integrates well with previously‐published
HE is primarily caused by a sufficiently severe loss in concen- data; it has a mechanistic correlation and is expected to facilitate
tration of high energy phosphates indicative of cerebral energy prognostic evaluation in patients with ALF.
failure. However, alterations of uptake and metabolism of the
brain’s primary energy substrate, glucose, have consistently Cell volume regulation and brain edema
been described in both acute and chronic liver diseases. Studies
using PET and the non‐metabolized glucose transport ligand Brain edema has been described in patients with cirrhosis and in
18
fluorodeoxyglucose in patients with cirrhosis and mild HE patients with ALF. However, the nature and pathological char-
show significant decreases in uptake in anterior cingulate cor- acteristics of the brain edema in the two conditions are quite
tex, a brain structure implicated in the monitoring of responses distinct. In ALF brain edema has been characterized consisting
to visual stimuli [39]. Not surprisingly, decreased brain glu- primarily of cytotoxic brain edema that, if uncontrolled, may
cose uptake in these patients was correlated with impaired per- progress to intracranial hypertension (ICH) and brain hernia-
formance on psychometric testing. This decrease of brain tion. Brain edema in ALF appears to have multiple causes
glucose uptake is currently considered to be the conse- including the accumulation of brain lactate [25] as described
quence of decreased brain energy requirements resulting from above as well as proinflammatory mechanisms resulting from
decreased neural activation in (rather than the cause) of HE microglial activation and release of cytokines [34].
[39]. Pathophysiologically‐relevant concentrations of ammo- Brain edema leading to ICH is uncommon in cirrhosis since
nia inhibit the TCA cycle, an essential metabolic pathway the swelling is low‐grade. However, it has the potential to result
involved in the maintenance of brain energy requirements in impairments of cell–cell signalling and to cause malfunction
[25]. Slowing of the TCA cycle results in the shuttling of pyru- of astrocytic proteins. NMR spectroscopic studies in patients
vate to lactate and cerebrospinal fluid lactate concentrations with cirrhosis have consistently shown disturbances of brain
correlate well with severity of HE in patients with cirrhosis organic osmolytes including increases of glutamate/glutamine
[40] as well as in brain microdialysates from patients with with compensatory reductions of myo‐inositol [44].
ALF where surges of increased intracranial pressure were fre-
quently found to accompany increased lactate concentrations
[41]. Brain lactate concentrations in excess of 12 mM have
Inhibitory neurotransmission
been reported at coma stages of encephalopathy in animal GABA is the principal inhibitory neurotransmitter system of
models of ALF [25]. Such concentrations are beyond mammalian brain and alterations of GABAergic transmission
the brain’s capacity to buffer lactate with the potential to result are implicated in a wide range of neurodegenerative and meta-
in acidosis. bolic brain diseases including HE.
48:  Hepatic Encephalopathy 623

GABAergic neurotransmission is mediated by the amino acid pathogenesis of HE in chronic liver diseases, investigations of
GABA via activation of a post‐synaptic GABA‐receptor com- the other modulatory site on the GRC, the NS modulatory site in
plex (GRC), a protein complex and ligand‐gated ion channel HE patient material yielded more positive results.
that is selective for chloride (Cl−) ion. Activation of the GRC by The molecular steps involved in NS synthesis in the brain and
GABA results in the entry of chloride through its pore resulting the relationship between ammonia exposure and NS synthesis is
in hyperpolarization of the post‐synaptic neuron. This results in shown in a simplified schematic form in Figure 48.3.
inhibition of central neurotransmission by diminishing the The process starts with the activation of a mitochondrial pro-
chance that an action potential occurs. tein located in glia (astrocytes and microglia) known as translo-
Activation of GABAergic transmission (also frequently cator protein (TLP) resulting in the entry of cholesterol into the
referred to as “increased GABAergic tone”) was originally pro- mitochondrion. A series of well‐established metabolic steps
posed in the 1980s based on visual evoked response patterns in then transform cholesterol into NS that are then available for
experimental animals with toxic liver injury and HE that were release into the synaptic cleft.
identical to patterns observed in normal animals treated with Two of these NS namely allopregnanolone (ALLO) and tet-
well‐established activators of the GRC [45]. These findings led rahydro‐deoxy corticosterone (THDOC), are potent positive
to intense interest in activation of the GRC as a major factor allosteric modulators of the GRC and, hence, are activators of
implicated in the pathogenesis of HE, an interest that continues GABAergic neurotransmission leading to enhanced central
to this day. Initially, studies were focused on the measurement neuroinhibition.
of integral components of the GABA system including GABA The first direct evidence for a role of NS in the pathogenesis
concentrations and metabolites, activities of enzymes responsi- of HE in cirrhosis was provided by the report of significant
ble for GABA synthesis as well as post‐synaptic GABA recep- increases of ALLO in brain tissue obtained at autopsy from
tors. In all cases, these parameters were found to be present in patients with decompensated cirrhosis who died in stage IV HE
normal amounts in animal models of acute and chronic liver (coma) [49]. ALLO concentrations in control material from
failure and, more importantly, in autopsied brain tissue from patients with no liver disorders, patients with decompensated
patients with decompensated cirrhosis who died in hepatic coma cirrhosis but no encephalopathy, as well as one patient who died
[46]. These findings led to a change in focus of research in this in uremic coma, were all within normal limits. Furthermore,
area and a renewal of interest in the multiple modulatory sites ALLO concentrations in brain extracts from HE patients were
and associated subunit proteins of the GRC itself.
The GRC is composed of an active binding site for GABA, Ammonia Cholesterol
often referred to as the GABA recognition site that, together with
allosteric sites, modulate the activity of the GRC. These sites are
targets for a range of substances including benzodiazepines and TLP
neuroactive steroids also known as neurosteroids (NS). Perineuronal GABA
Pregnenolone
The benzodiazepine modulatory site on the GRC has been astrocyte MM Nerve
or
extensively studied in brain tissue obtained at autopsy from SER
Terminal
microglia
patients with cirrhosis who died in stage IV HE as well as in
living patients with mild HE by use of the selective benzodi- Astrocyte
azepine modulatory site antagonist PET ligand Ro15‐1788
(flumazenil). No significant alterations of binding site affini-
ties or densities of these sites were observed when compared Allopregnanolone GABA
to data from age‐matched control material [46]. Furthermore,
subsequent studies failed to demonstrate any significant alter- GABA
NS
ations of the allosteric coupling between the benzodiazepine site
site
modulatory site and the GABA recognition in the HE patient
material [47]. Together, these findings clearly demonstrate Post-Synaptic
that the GRC with respect to the benzodiazepine modulatory neuron
Cl
site is functioning normally in patients with HE related to
cirrhosis. Neuroinhibition
The reports of no significant alterations in expression or cou-
pling of the GABA receptor machinery of the brain in HE led to HE
a renewed interest in the search for so‐called “endogenous Figure 48.3  Pathophysiologic link between ammonia toxicity and
ligands” for these receptors. Several such substances were “increased GABAergic tone” in HE. A schematic representation of the
detected. Quantitative mass spectrometric analysis followed to sequence of steps whereby activation of the translocator protein (TLP)
unequivocally establish the identity of these substances. They situated on the mitochondrial membrane (MM) of the astrocyte or
were identified as well‐established pharmaceutical benzodiaz- microglial cell occurs. This results in increased transport of cholesterol
epines or their metabolites [48] that were traced to their admin- into the cell and its conversion to pregnenolone and the NS allopregna-
nolone. Stimulation of the NS modulatory site on the GABA‐A receptor
istration as muscle relaxants during prior endoscopic work‐up complex by allopregnanolone then amplifies the signal produced by the
or as sedatives. neurotransmitter (GABA) acting on the adjacent GABA recognition
In contrast to the disappointing findings in relation to the site. The net result is increased entry of chloride (Cl−) and enhanced
benzodiazepine modulatory site and its potential role in the neuroinhibition that contributes to the pathogenesis of HE in cirrhosis.
624 THE LIVER:  MANAGEMENT AND THERAPEUTIC STRATEGIES FOR HE

present in sufficiently high concentrations to result in a 53% [54] has now been widely discontinued following the report [55]
increase in binding of the GABA agonist ligand 3H‐muscimol to that patients with cirrhosis benefit from normal protein (1.2–1.5 g
brain membrane preparations thus satisfying the requirement kg−1 daily). Diets rich in vegetable and dairy protein are
that concentrations of ALLO in HE patient material were suffi- encouraged and patients intolerant to dietary protein may con-
cient to result in “increased GABAergic tone” [50]. Novel com- sider branched chain amino acid supplements as an alternative
pounds with antagonist action at the NS modulatory site on the to protein [53].
GRC have recently been synthesized. Further information and Vitamin deficiencies in cirrhosis result from impaired hepatic
results of preliminary clinical trials are summarized in the sec- function and diminished hepatic reserves as well as inadequate
tion, GABA receptor modulators in this chapter. dietary intake and malabsorption. A neuropathologic study
examining brain tissue from patients with cirrhosis who died in
stage IV HE revealed lesions in thalamus and mamillary bodies
that are characteristic of Wernicke’s encephalopathy together
MANAGEMENT AND THERAPEUTIC with cerebellar degeneration consistent with vitamin B1 defi-
STRATEGIES FOR HE ciency [11]. The B1 deficiency in these patients was attributed to
a loss of hepatic stores of the vitamin and the prevalence of
The ultimate test of the pertinence of the diverse theories of the lesions was threefold higher than that reported in material from
pathogenesis of HE rests in the clinic where the impact of non‐liver disease controls. It was recommended that B1 supple-
appropriate treatments based upon the nutritional, metabolic ments be administered to all patients with decompensated cir-
and toxic factors identified in this chapter may be undertaken. rhosis particularly those with alcoholic etiology [51].
Deficiencies of other vitamins including B2, B6, and B12 have
also been reported in cirrhosis associated with alcoholic liver
Nutritional management
damage and/or diminished hepatic storage but causal relation-
The pathogenesis of malnutrition in liver disease is complex and ships with complications of cirrhosis were not established. Zinc
multifactorial involving reduction of dietary intake due to ano- deficiency is common in cirrhosis but evidence for beneficial
rexia and dietary restriction, alterations of nutrient biosynthesis, effects of zinc supplements on HE is equivocal [56].
impaired intestinal absorption, disturbances of substrate utiliza- It has been suggested that malnutrition is a factor that
tion, increased protein loss, and abnormal metabolism resulting increases both costs and post‐transplant complications [57].
from, for example, the proinflammatory state. Neurological complications post transplantation are numerous
The functional integrity of the liver is essential for the provi- and may include diffuse encephalopathy, seizures, intracranial
sion, inter‐organ transfer, and metabolism of essential nutrients. hemorrhage, stroke, progressive neurological deterioration, cen-
As outlined in the present review, nitrogen metabolism plays a tral pontine myelinolysis, ataxia, psychosis, confabulation, and
key role in the development of HE in cirrhosis and in conditions peripheral neuropathy. While some of these complications
of liver failure, skeletal muscle adapts metabolically to serve as likely result from specific nutritional and metabolic deficits
the major back‐up organ for ammonia removal. Sarcopenia or alluded to above, further studies are required to determine the
loss of muscle mass is commonly encountered in cirrhosis where precise nutritional factors implicated and facilitate specific
it has adverse effects on survival, health‐related quality of life, as nutritional recommendations.
well as outcome following liver transplantation [51]. Sarcopenia
may also lead to worsening of the complications of cirrhosis
including portal hypertension, ascites, and HE. Indeed, results of Ammonia‐lowering strategies
a prospective study reveal that muscle depletion in patients with Given the wealth of evidence derived from biochemical, neuro-
cirrhosis increases the risk of both MHE and OHE [52]. imaging, and spectroscopic studies in support of a central role of
Obesity and its associated alterations of nutrient intake are ammonia in the pathogenesis of HE, it is not surprising that the
frequent occurrences in patients with non‐alcoholic fatty liver therapeutic strategy that has so far stood the test of time relates
disease (NAFLD) and play an important role in determining to the reduction of hyperammonemia. Several treatments aimed
rates of progression to non‐alcoholic steatohepatitis (NASH) at reducing the production of ammonia and/or its removal have
and cirrhosis. A recent consensus document from the been found to be effective in this regard.
International Society for Hepatic Encephalopathy and Nitrogen
Metabolism (ISHEN) on guidelines for the nutritional manage-
Non‐absorbable disaccharides
ment of HE in cirrhosis has appeared [53]. The stated objectives
of nutritional management in patients with cirrhosis include the Non‐absorbable disaccharides such as lactulose and lactitol
correction of specific nutritional deficiencies, prevention and remain first‐line agents for the lowering of the production and
treatment of the complications of cirrhosis as well as the com- absorption of ammonia. They are metabolized by bacteria in the
plications of liver transplantation, and the support of liver colon with production of acetic and lactic acids and the conse-
regeneration. quent acidification of the colon creates a hostile environment for
Energy requirements of 35–45 kcal g−1 daily are recom- urease‐rich intestinal bacteria that are responsible for the pro-
mended for patients with cirrhosis with small meals evenly dis- duction of ammonia. In addition, these agents facilitate the for-
tributed throughout the day and a late‐night snack of complex mation of NH4+ leading to reduced ammonia production and
carbohydrates to minimize protein utilization. The practice of their cathartic properties may result in up to fourfold increases
dietary protein restriction that was prevalent until the late 1990s of fecal nitrogen excretion [58].
48:  Hepatic Encephalopathy 625

Although non‐absorbable disaccharides are widely employed an ability to increase urea production by residual hepatocytes
as first‐line therapy for HE in cirrhosis, systematic reviews and (L‐ornithine is a urea cycle substrate) and increased glutamine
meta‐analyses of trials have yielded mixed results. In an analy- synthesis in skeletal muscle [70]. There is evidence that LOLA
sis of ten studies, it was concluded that there was insufficient has direct hepatoprotective actions in patients with cirrhosis
evidence to recommend them for standard therapy until they and HE [71].
were shown to confer benefit over placebo [59]. The analysis Beneficial effects of LOLA for the management and treat-
was subsequently challenged based upon shortcomings in ment of HE in cirrhosis have been reported in over 20 rand-
experimental design and methodology [60] and further studies omized clinical trials over the last 20 years. The intravenous
including a second meta‐analysis of lactulose revealed signifi- formulation of LOLA is particularly effective in the treatment of
cant improvements in cognitive function and health‐related low‐grade [72] and bouts of overt [73] HE in cirrhosis. A recent
quality of life in MHE [61] as well for the primary prophylaxis meta‐analysis demonstrated that the oral formulation of LOLA
of overt HE [62] in patients with cirrhosis. A subsequent meta‐ is particularly effective for the treatment of MHE [74].
analysis showed that, compared with placebo/no intervention, LOLA is reportedly equivalent or superior in efficacy com-
the non‐absorbable disaccharides were associated with signifi- pared to rifaximin [69], probiotics [68, 69], and lactulose [68]
cant overall benefits on HE [63]. for the treatment of MHE and a network meta‐analysis con-
firmed these findings [75]. However, in the only trial conducted
Antibiotics to date, LOLA did not appear to be effective for the treatment of
HE in ALF [76].
The aminoglycoside antibiotic neomycin is still used for the
treatment of patients with cirrhosis and HE despite its poten- Branched chain amino acids (BCAAs)
tially serious side‐effects that include ototoxicity and nephro-
toxicity since the drug is partially absorbed. Rifaximin is a Results of a Cochrane review on the efficacy of BCAAs for
poorly absorbed, broad spectrum antibiotic that is effective for improvement of mental state concluded that patients treated
lowering of blood ammonia and improvement of mental state in with BCAAs are more likely to recover from HE compared to
patients with MHE as well as in those with overt HE where its patients receiving control regimens that included lactulose and
efficacy has been found to exceed that of neomycin. Results of neomycin [77]. Two large studies revealed beneficial effects of
a large, randomized, double‐blind, placebo‐controlled trial BCAAs on nutrition and survival, effects that may outweigh any
demonstrated that rifaximin reduced the risk of repeat episodes effects on HE per se [77, 78]. Mechanisms responsible for the
of overt HE and reduced the time to first hospitalization with no beneficial effects of BCAAs have not been fully elucidated
serious adverse events [64]. A subsequent systematic review and despite their continuing use for over 35 years. Suggested mech-
meta‐analysis of 19 trials with 1390 patients confirmed that anisms include beneficial effects on their use as substrates for
rifaximin is useful for the management of HE with a beneficial hepatic protein synthesis as well as stimulation of liver regen-
effect on HE severity and for treatment of acute episodes as well eration and increases of cerebral perfusion.
as on patient mortality [65]. Rifaximin also improves psycho-
metric test scores and HRQOL in patients with MHE [66]. Benzoate, phenylacetate
Benzoate and phenylacetate are agents that have been used
Probiotics extensively for the lowering of blood ammonia in children
Urease‐producing bacteria are plentiful in the gut where they with congenital urea cycle disorders. Phenylacetate is gener-
convert urea to carbamate and ammonia. Modifications of the ally given orally as its precursor phenylbutyrate. Phenylacetate
gut microflora balance by the introduction of bacteria that com- condenses with glutamine to form phenylacetyl glutamine
pete with those expressing urease (probiotics) are effective whereas benzoate condenses with glycine to form hippurate.
ammonia‐lowering agents. Treatment of patients with non‐alco- Both condensation products are excreted in the urine. The
holic cirrhosis and MHE with probiotic yoghurt results in slow- ammonia‐lowering action results from removal of these
ing of progression to overt HE [67]. Furthermore, randomized ammoniagenic substrates as well as the diversion of available
controlled trials comparing probiotics with placebo reveal ammonia for the replenishment of the amino acid pools.
improvements in arterial ammonia concentrations, psychomet- Sodium benzoate was reported to manifest comparable ­efficacy
ric test scores, and HRQOL in MHE patients [68, 69]. The to lactulose in a prospective study in patients with cirrhosis
degree of improvement was equivalent to that of lactulose [68] and acute HE [79].
or rifaximin [69]. A subsequent meta‐analysis of 14 studies Mixtures of phenylacetate with other agents such as glycerol
totaling 1152 patients confirmed the comparable beneficial or L‐ornithine have recently been introduced as possible ammo-
effects of probiotics and lactulose and went on to demonstrate nia‐lowering strategies for possible use in the treatment of HE in
efficacy of probiotics in reducing hospitalization rates as well as cirrhosis. In the case of glycerol phenylbutyrate (GBP), a multi-
rates of progression to overt HE. center phase 2 trial, patients in the GBP arm manifested lower-
ing of blood ammonia and experienced less HE events as well as
less HE hospitalizations compared to placebo [80].
L‐ornithine L‐aspartate (LOLA)
Ornithine phenylacetate (OP), a mixture of L‐ornithine and
LOLA is a mixture of two endogenous amino acids with estab- phenylacetate has been shown to reduce blood ammonia in
lished ammonia‐lowering properties. Multiple mechanisms experimental animal models of HE, a result that is not surprising
responsible for the ammonia‐lowering actions of LOLA include given that both the L‐ornithine and phenylacetate moieties are
626 THE LIVER:  MANAGEMENT AND THERAPEUTIC STRATEGIES FOR HE

independently well‐established ammonia‐lowering compounds. Anti‐inflammatory agents


A study on the safety, tolerability, and pharmacokinetics of OP
Recognition of the key roles of systemic inflammation and neuro-
in patients with acute liver injury or failure showed significant
inflammation in the pathogenesis of HE in both ALF and in cir-
urinary ammonia excretion, good tolerability, and no safety sig-
rhosis has resulted in renewed interest in the search for agents with
nals [81]. Randomized, controlled studies are now required to
anti‐inflammatory properties as potential therapies. Ibuprofen
determine its use as an ammonia‐scavenging agent in patients
attenuates the neurologic deficits in learning ability in an experi-
with HE.
mental animal model of type B HE [89]. A central role of TNFα has
been proposed to explain HE in cirrhosis based upon the observa-
L‐carnitine
tions that many precipitating factors including infection, gastroin-
L‐carnitine has been shown to manifest a significant protective testinal bleeding, constipation, and renal failure are associated with
effect in patients with cirrhosis and ammonia‐precipitated increased circulating levels of the cytokine. Moreover, many thera-
encephalopathy [82]. The protective effect of L‐carnitine was pies known to lower ammonia such as lactulose, antibiotics, and
evident for patients with overt HE grades I and II as well as for probiotics also effectively reduce TNFα production [28].
patients with MHE as assessed by improvements in scores for Following up on the report that deletion of the gene coding
NCT‐A. Studies in both experimental congenital hyperam- for the TNFα receptor resulted in delayed onset of HE and pre-
monemia [83] and portal‐systemic encephalopathy [84] showed vention of brain edema in ALF due to toxic liver injury [35], a
a significant protective effect of L‐carnitine on brain energy more recent study showed that the systemic sequestration of
metabolism characterized by improved brain glucose oxidative TNFα by administration of the TNFα‐neutralizing compound
capacity and prevention of cerebrospinal fluid lactate accumula- etanercept resulted in reduction of microglial activation and
tions [84]. As alluded to earlier in this chapter, decreased glu- delayed progression of HE in mice with acute liver injury [90].
cose oxidation in brain in liver failure is the consequence of Minocycline is a semi‐synthetic tetracycline antibiotic that
ammonia‐induced inhibition of the TCA cycle by ammonia limits microglial activation by a mechanism that is independent of
leading to lactate accumulation [25]. its antimicrobial properties and studies in the liver ischemic rat
These novel beneficial effects of L‐carnitine are the conse- model of ALF showed that minocycline prevents microglial acti-
quence of inhibition of one of the major metabolic consequences vation and attenuates the increased synthesis of proinflammatory
of exposure to ammonia namely that of decreased brain glucose cytokines including TNFα in the brain while leading to a slowing
oxidation rather than a direct effect on ammonia production or of progression of HE and prevention of brain edema [91].
removal. The beneficial effect of mild hypothermia in ALF involves
anti‐inflammatory mechanisms at both hepatic and cerebral
Neuropharmacological approaches levels [92].

GABA receptor modulators Other centrally‐acting agents


A clinical trial of the efficacy of flumazenil, a potent antagonist Evidence of a dopaminergic deficit together with neurological
of the benzodiazepine modulatory site on the GABA receptor symptoms characteristic of Parkinson’s disease are well established
complex was undertaken in patients with cirrhosis and moderate findings in patients with HE associated with cirrhosis. It is there-
to severe HE. The trial demonstrated clinical improvement of fore not surprising that attempts have been made to treat these
HE in a significant subset of patients [85] and this was subse- patients with agents with the potential to reactivate the dopaminer-
quently confirmed in a systematic review and meta‐analysis gic system. Dopamine‐replacement therapy by the precursor amino
[86] although, as expected, the benefit was relatively short‐act- acid L‐DOPA was found to be ineffective for improvement of cog-
ing due to the short half‐life of flumazenil. nitive symptoms [93] but, subsequently, using a more appropriate
The findings of increased brain concentrations of allopreg- Parkinson disease rating scale for assessment of clinical efficacy,
nanolone (described in the section, Inhibitory neurotransmis- L‐DOPA was found to result in improvements in motor perfor-
sion in the current chapter) the potent agonist of the NS site on mance [16]. Clinical trials with the dopamine receptor agonist bro-
the GABA‐A receptor in material from HE patients resulted in mocriptine have yielded conflicting results [94, 95] but, again,
a renewed interest in the discovery of compounds with the nec- clinical endpoints such as the portal‐systemic encephalopathy
essary configuration to block the actions of NS on the NS (PSE) index rather than tests aimed at assessment of dopaminergic
­modulatory site on the GRC [47]. So‐called “GABA‐receptor‐ function and motor coordination were used making interpretation
modulating steroid antagonists” (GAMSA) were synthesized of the findings problematic. Therapeutic options for the treatment
and tested in animal models. One example, 3β‐20β‐ of Parkinsonism in cirrhosis have been reviewed [96].
dihydroxy‐5α‐pregnane (UC1011) was shown previously to
inhibit the ALLO‐induced increases of GABA‐induced
increases of chloride uptake into cerebral cortical and hip-
Liver transplantation and HE
pocampal preparations making it an ideal candidate for further Liver transplantation (LT) is the ultimate treatment for decom-
study [87]. More recently, a second member of the GAMSA pensated cirrhosis. Although LT generally confers a level of
family, GR 3027 has been shown to improve spatial learning, reversibility of some of the symptoms of HE, there is a growing
motor coordination, and circadian rhythms in an experimental body of evidence suggesting that this reversibility is far from
model of HE in chronic liver disease [88]. Controlled clinical complete [97] and patients with a history of overt HE have an
trials are now required. increased probability of persistent neurocognitive dysfunction
48:  Hepatic Encephalopathy 627

post‐LT compared to those without prior overt HE. This finding Health Research (CIHR) and the Canadian Association for the
begs the question whether it is the episode(s) of overt HE or the study of the Liver (CASL). The author is grateful to Mr Jonas
LT that is responsible for the persistent deficits. This issue was Eric Pilling for the design of Figure 48.3.
subsequently addressed in a prospective study in which it was
demonstrated that even a single episode of overt HE was accom-
panied by a learning deficit and that multiple episodes led to
deficits in reaction times, attention, psychomotor speed, and REFERENCES
working memory [98]. Overt HE patients may suffer a loss of
1. Mullen, K.D. and Prakash, R.K. (Eds.) Hepatic Encephalopathy, Clinical
“cognitive reserve” that this could have important implications
Gastroenterology, Humana Press/Springer New York, pp. 97–102
for the assignment of priority for LT. Other possible causes of 2. Rossle, M. and Euringer, W. Hepatic encephalopathy in patients with tran-
post‐LT cognitive dysfunction include anoxic intraoperative sjugular intrahepatic portosystemic shunt (TIPS), in Hepatic encephalopa-
complications, immunosuppressant medication, comorbidities, thy: Clinical Gastroenterology, (eds. K.D. Mullen and R.K. Prakash),
manganese neurotoxicity [17], and thiamine deficiency [11]. Humana Press/Springer, New York, 2012, pp. 211–20.
3. Bajaj, J.S., Wade, J.B., Gibson, D.P. et al. The multi‐dimensional burden of
In a study comparing the incidence and severity of neurologi-
cirrhosis and hepatic encephalopathy on patients and caregivers. Am J
cal complications after cadaveric versus living donor LT, the most Gastroenterol, 2011;10:6646–53.
common complications (encephalopathy and seizures) were sig- 4. Bajaj, J.S. Introduction and setting the scene: new nomenclature of hepatic
nificantly lower following living donor LT. These findings could encephalopathy and American Association for the study of liver diseases/
be accounted for, at least in part, to the shorter cold ischemia time. European Association for the study of the liver guidelines. Clin Liv Dis,
2017;9:48–51.
The incidence of neurological complications post‐LT was not 5. Blei, A.T. and Cordoba, J. Hepatic encephalopathy practise guidelines. Am J
dependent on the nature of the immunosuppressant used [99]. Gastroenterol, 2001;96:1968–75.
6. Bajaj, J.S., Cordoba, J., Mullen, K.D. et  al. Review article: the design of
clinical trials in hepatic encephalopathy‐an ISHEN consensus statement.
Future directions for the field of HE Aliment Pharmacol Ther, 2011;3:339–47.
7. Clemmesen, J.O., Larsen, F.S., Kondrup, J. et  al. Cerebral herniation in
When the author of this chapter began his research in HE some patients with acute liver failure is correlated with arterial ammonia concen-
four decades ago, there were two major alternative agents tration. Hepatology, 1999;2:948–53.
employed for the management and treatment of HE in cirrhosis; 8. Kircheis, G., Wettstein, M., Timmermann, L. et al. Critical flicker frequency
a non‐absorbable disaccharide and an aminoglycoside antibi- for quantification of low‐grade hepatic encephalopathy. Hepatology,
2002;35:357–66.
otic. Both acted at the level of the gut, both resulted in lowering
9. Kato, M., Hughes, R.D., and Keays, R.T. Electron microscopic study of
of blood ammonia; both had serious side‐effects. Many of us brain capillaries in cerebral edema from fulminant hepatic failure.
wonder if anything has changed. Well, the disaccharide is still Hepatology, 1992;15:1060–6.
with us but at least we appear to have found a better antibiotic. 10. Butterworth, R.F., Giguere, J.F., Michaud, J. et al. Ammonia: key factor
It appears that the search for agents, some new, some recycled, in the pathogenesis of hepatic encephalopathy. Neurochem Pathol,
1987:6–12.
targeting the lowering of blood ammonia is here to stay. 11. Kril, J.J. and Butterworth, R.F. Diencephalic and cerebellar pathology in
Progress in the search for more rational therapies for HE is alcoholic and non‐alcoholic patients with end‐stage liver disease.
slow. Despite evidence to the contrary, it is widely believed that Hepatology, 1997;2:637–41.
HE results principally, if not entirely, from the neurotoxic 12. Cordoba, J., Alonso, J, Rovira, A. et  al. The development of low‐grade
effects of ammonia. Consequently, search for new therapies cerebral edema in cirrhosis is supported by the evolution of 1H‐magnetic
resonance abnormalities after liver transplantation. J Hepatol,
remains focused on the gut. 2001;3:598–604.
HE is, after all, a brain disorder and advances in our under- 13. Butterworth, R.F. Hepatic encephalopathy: a central neuroinflammatory dis-
standing of intercellular signalling involving defined neuro- order? Hepatology, 2011;53:372–6
transmitter and neuromodulators in brain areas implicated in the 14. Zemtsova, I., Gorg, B., Keitel, V. et  al. Microglial activation in hepatic
encephalopathy in rats and humans. Hepatology, 2011;5:404–15.
maintenance of consciousness, cognitive, and motor function in
15. Butterworth, R.F., Neuronal cell death in hepatic encephalopathy. Metab
liver disorders continue to be elucidated as do novel concepts Brain Dis, 2007;2:209–20.
involving central anti‐inflammatory molecules. Yet translation 16. Burkhard, P.R., Delavelle, J., Du Pasquier, R. et al. Chronic Parkinsonism
of these discoveries, and sometimes the use of new molecules associated with cirrhosis. Arch Neurol, 2003;60:521–8.
themselves, remains sluggish. One exception presented in this 17. Butterworth, R.F., Spahr, L., Fontaine, S. et al. Manganese toxicity, dopa-
minergic dysfunction and hepatic encephalopathy. Metab Brain Dis,
chapter relates to the discovery of the effectiveness of an antag- 1995;10:559–67.
onist of the neurosteroid modulatory site on the GABA‐A recep- 18. Ahluwalia, V., Wade, J.B., Moeller, F.G. et al. The etiology of cirrhosis is a
tor complex of brain. Clinical trials are ongoing. strong determinant of brain reserve: a multi‐modal MR imaging study. Liver
The future directions of HE research and patient care will Transpl, 2015;2:1123–32.
undoubtedly be based to an increasing extent on targeting of the 19. Desjardins, P., Rao, K.V., Michalak, A. et al. Effect of portacaval anastomo-
sis on glutamine synthetase protein and gene expression in brain, liver and
brain per se. skeletal muscle. Metab Brain Dis, 1999;14:273–80.
20. Sorensen, M. and Ott, P. Cerebral ammonia metabolism in cirrhosis. Funct
Mol Imaging Hepatol, 2012:153–9.
21. Bernal, W., Hall, C., Karvellas, C.J. et al. Arterial ammonia and clinical risk
ACKNOWLEDGMENTS factors for encephalopathy and intracranial hypertension in acute liver fail-
ure. Hepatology, 2007;46:1844–52.
22. Laubenberger, J., Haussinger, D., Boyer, S. et al. Proton magnetic resonance
Studies and related publications from the author’s research unit spectroscopy of brain in symptomatic and asymptomatic patients with liver
were funded by operating grants from the Canadian Institutes of cirrhosis. Gastroenterology, 1997;112:1610–6.
628 THE LIVER:  REFERENCES

23. Zwingmann, C., Chatauret, N., Liebfritz, D. et al. Selective increase of brain 50. Butterworth, R.F. Neurosteroids in hepatic encephalopathy: novel insights
lactate synthesis in experimental acute liver failure: results of a 13C/1H and new therapeutic opportunities. J Steroid Biochem Mol Biol,
nuclear magnetic resonance study. Hepatology, 2003;3:420–8. 2015;160:94–7.
24. Felipo, V. and Butterworth, R.F. Neurobiology of ammonia. Prog Neurobiol, 51. Bemeur, C. and Butterworth, R.F. Nutrition in the management of cirrhosis
2002;67:259–79. and its neurological complications. J Clin Exp Hepatol, 2014;4(2):141–50.
25. Lai, J.C.K. and Cooper, A.J.L. Brain α‐ketoglutarate dehydrogenase: kinetic 52. Merli, M., Giusto, M., Lucidi, C. et al. Muscle depletion increases the risk of
properties, regional distribution and effects of inhibitors. J Neurochem, overt and minimal hepatic encephalopathy: results of a prospective study.
1986;47:376–86. Metab Brain Dis, 2013;28:281–84.
26. Pomier Layrargues, G., Spahr, L. and Butterworth, R.F. Increased manganese 53. Amodio, P., Bemeur, C., Butterworth, R.F. et al. The nutritional management
concentrations in pallidum of cirrhotic patients, Lancet, 1995;34:735–41. of hepatic encephalopathy in patients with cirrhosis: International Society
27. Rolando, N., Wade, J., Davalos, M. et  al. The systemic inflammatory for Hepatic Encephalopathy and Nitrogen Metabolism consensus.
response syndrome in acute liver failure. Hepatology, 2000;3:734–9. Hepatology, 2013;58:325–36.
28. Odeh, M. Pathogenesis of hepatic encephalopathy: the tumour necrosis fac- 54. Soulsby, C.T. and Morgan, M.Y. Dietary management of hepatic encepha-
tor‐alpha theory. Eur J Clin Invest, 2007;3:291–304. lopathy in cirrhotic patients; survey of current practice in United Kingdom.
29. Chastre, A., Belanger, M., Nguyen, B.N. et al. Lipopolysaccharide precipi- BMJ, 1999;31:8391.
tates hepatic encephalopathy and increases blood‐brain barrier permeability 55. Cordoba, J., Lopez‐Hellin, J., Planas, M. et al. Normal protein diet for epi-
in mice with acute liver failure. Liver Int, 2013;3:353–61 sodic hepatic encephalopathy: results of a randomized study. J Hepatol,
30. Zieve, L., Doizaki, W.M., and Zieve, F.J. Synergism between mercaptans and 2004;41(1):38–43.
ammonia or fatty acids in the production of coma: a possible role of mercap- 56. Bresci, G., Parisi, G., and Bant, S. Management of hepatic encephalopathy
tans in the pathogenesis of hepatic coma. J Lab Clin Med, 1974;8:26–8. with oral zinc supplementation: a long‐term treatment. Eur J Med,
31. Shawcross, D.L., Wright, G., Olde Daminck, S.W.M. et al. Role of ammonia 1993;2(7):414–6.
and inflammation in minimal hepatic encephalopathy. Metab Brain Dis, 57. Merli, M., Giusto, M., Gentili, F. et al. Nutritional status: its influence on the
2007;2:125–38. outcome of patients undergoing liver transplantation. Liver Int,
32. Butterworth, R.F. Pathogenesis of hepatic encephalopathy in cirrhosis: the 2009;30:208–14.
concept of synergism revisited. Metab Brain Dis, 2016;31(6):1211–15. 58. Sharma, P. and Sharma, B.J. Management of overt hepatic encephalopathy. J
33. Schliess, F., Gorg, B., Fischer, R. et al. Ammonia induces MK801‐sensitive Clin Exp Hepatol, 2015;5:S82–7.
nitration and phosphorylation of protein tyrosine residues in rat astrocytes. 59. Als‐Nielsen, B., Gluud, L.L., and Gluud, C. Non‐absorbable disaccharides
FASEB J, 2002;1:739–41. for hepatic encephalopathy: systematic review of randomized trials. BMJ,
34. Butterworth, R.F. The liver‐brain axis in liver failure: neuroinflammation 2004. doi:10.1136/bmj.38048.506134.EE
and encephalopathy. Nat Rev Gastroenterol Hepatol, 2013;1:522–8. 60. Morgan, M.Y., Blei, A., Grungreiff, K. et  al. The treatment of hepatic
35. Bemeur, C., Qu, H., Desjardins, P. et al. IL‐1 or TNF receptor gene deletion encephalopathy. Metab Brain Dis, 2007;22:389–405.
delays onset of encephalopathy and attenuates brain edema in experimental 61. Prasad, S., Dhiman, R.K., Duseja, A. et  al. Lactulose improves cognitive
acute liver failure. Neurochem Int, 2010;5:213–5. functions and health‐related quality of life in patients who have minimal
36. Cagnin, A., Taylor‐Robinson, S.D., Forton, D.M. et al. In vivo imaging of hepatic encephalopathy. Hepatology, 2007;4:549–59.
cerebral “peripheral benzodiazepine binding sites” in patients with hepatic 62. Sharma, P., Sharma, B.C., Agrawal, A. et al. Primary prophylaxis of overt
encephalopathy. Gut, 2006;5:547–53. hepatic encephalopathy in patients with cirrhosis: an open‐labelled rand-
37. Duchini, A., Govindarajan, S., Santucci, M. et al. Effects of tumor necrosis omized controlled trial of lactulose versus no lactulose. J Gastroenterol
factor‐a and interleukin‐6 on fluid‐phase permeability in CNS‐derived Hepatol, 2012;27:1329–35.
endothelial cells. J Invest Med, 1996;4:474–82. 63. Gluud, L.L., Vilstrup, H., and Morgan, M.Y. Non‐absorbable disaccharides
38. Andersson, A.K., Adermark, L., Persson, M. et  al. Lactate contributes to versus placebo/no intervention and lactulose versus lactitol for the preven-
ammonia‐mediated astroglial dysfunction during hyperammonemia, tion and treatment of hepatic encephalopathy in people with cirrhosis.
Neurochem Res, 2009;3:555–65. Cochrane Database Syst Rev, 2016;4, CD003044.
39. Lockwood, A.H., Weissenborn, K., and Butterworth, R.F. An image of the 64. Bass, N.M., Mullen, K.D., Sanyal, A. et al. Rifaximin treatment in hepatic
brain in patients with liver disease. Curr Opin Neurol, 1997;1:525–33. encephalopathy. N Engl J Med, 2010;362:1071–81.
40. Yao, H., Sadoshima, S., Fijii, K. et al. Cerebrospinal fluid lactate in patients 65. Kimer, N., Krag, A., Moller, S. et al. Systematic review with meta‐analysis:
with hepatic encephalopathy. Eur Neurol, 1987;2:182–7. the effects of rifaximin in hepatic encephalopathy. Aliment Pharmacol Ther,
41. Tofteng, F., Jorgensen, L., Hansen, B.A. et  al. Cerebral microdialysis in 2014;40:123–32.
patients with fulminant hepatic failure, Hepatology, 2002;3(6):1333–40. 66. Sidhu, S.S., Goyal, O., Mishra, B.P. et al. Rifaximin improves psychometric
42. Larsen, F.S., Olsen, K.S., Ejlersen, E. et al. Cerebral blood flow autoregula- performance and health‐related quality of life in patients with minimal
tion and transcranial Doppler sonography in patients with cirrhosis. hepatic encephalopathy (the RIME trial). Am J Gastroenterol,
Hepatology, 1995;22(3):730–6. 2011;106:307–16.
43. Aggarwal, S., Obrist, W., Yonas, H. et al. Cerebral hemodynamic and meta- 67. Bajaj, J.S., Saoian, K., Christensen, K.M. et al. Probiotic yogurt for the treat-
bolic profiles in fulminant hepatic failure: relationship to outcome. Liver ment of minimal hepatic encephalopathy. Am J Gastroenterol,
Transpl, 2005;11(11):1353–60. 2008;103:1707–15.
44. Haussinger, D. Low grade cerebral edema and the pathogenesis of hepatic 68. Mittal, V.V., Sharma, B.C., Sharma, P. et al. A randomized controlled trial
encephalopathy in cirrhosis. Hepatology, 2006;43:1187–90. comparing lactulose, probiotics and L‐ornithine L‐aspartate in treatment of
45. Jones, E.A. Ammonia, the GABA neurotransmitter system and hepatic minimal hepatic encephalopathy. Eur J Gastroenterol Hepatol,
encephalopathy. Metab Brain Dis, 2002;17:275–81. 2011;23:725–32.
46. Butterworth, R.F., Lavoie, J., Giguere, J.F. et al. Affinities and densities of 69. Sharma, K., Pant, S., Misra, S. et al. Effect of rifaximin, probiotics and L‐
high‐affinity [3H] muscimol (GABA‐A) binding sites and of central benzo- ornithine L‐aspartate on minimal hepatic encephalopathy: a randomized
diazepine receptors are unchanged in autopsied brain tissue from cirrhotic controlled trial. Saudi J Gastroenterol, 2014;20:225–32.
patients with hepatic encephalopathy. Hepatology, 1988;8(5):1084–8. 70. Rose, C., Michalak, A., Pannunzio, P. et al. L‐ornithine L‐aspartate in experi-
47. Ahboucha, S. and Butterworth, R.F. Pathogenesis of hepatic encephalopa- mental portal‐systemic encephalopathy: therapeutic efficacy and mechanism
thy: a new look at GABA from the molecular standpoint. Metab Brain Dis, of action. Metab Brain Dis, 1998;13:147–57.
2004;19;331–43. 71. Butterworth, R.F. and Grüngreiff, K. L‐ornithine L‐aspartate (LOLA) for the
48. Butterworth, R.F. and Wells, J. Detection of benzodiazepines in hepatic treatment of hepatic encephalopathy in cirrhosis: evidence for novel hepato-
encephalopathy: reply. Hepatology, 1995;21:604–5. protective mechanisms. JSM Liver Clin Res, 2019;3:5.
49. Ahboucha, S., Pomier Layrargues, G., Mamer, O. et al. Increased brain con- 72. Kircheis, G., Nilius, R., Held, C. et al. Therapeutic efficacy of L‐ornithine
centrations of a neuroinhibitory steroid in human hepatic encephalopathy. L‐aspartate infusions in patients with cirrhosis and hepatic encephalopathy.
Ann Neurol, 2005;58:169–70. Hepatology, 1997;25:1351–60.
48:  Hepatic Encephalopathy 629

73. Sidhu, S.S., Sharma, B.C., Goyal, O. et al. L‐ornithine L‐aspartate in bouts 87. Turkmen, S., Lundgren, P., Birzniece, V. et  al. 3Beta‐20beta‐dihydroxy‐5‐
of overt hepatic encephalopathy. Hepatology, 2018;67:700–10. alpha‐pregnane (UC1011) antagonism of the GABA potentiation and the
74. Butterworth, R.F., Kircheis, G., Hilger, N. et al. Efficacy of L‐ornithine L‐ learning impairment induced in rats by allopregnanolone. Eur J Neurosci,
aspartate for the treatment of hepatic encephalopathy and hyperammonemia 2004;20:1604–12.
in cirrhosis: systematic review and meta‐analysis of randomized controlled 88. Johansson, M., Agusti, A., Llansola, M. et  al. GR3027 antagonizes
trials. J Clin Exp Hepatol. 2018;8(3):301–13. GABA‐A receptor‐potentiating neurosteroids and restores spatial learn-
75. Zhu, G.Q., Shi, K.Q., Huang, S. et al. Systematic review with network meta‐ ing and motor coordination in rats with chronic hyperammonemia and
analysis: the comparative effectiveness and safety of interventions in patients hepatic encephalopathy. Am J Physiol Gastrointest Liver Physiol,
with overt hepatic encephalopathy. Aliment Pharmacol Ther, 2015;41:624–35. 2015;309:G400–9.
76. Acharya, S.K., Bhatia, V., Sreenivas, V. et  al. Efficacy of L‐ornithine L‐ 89. Cauli, O., Rodrigo, R., Piedrafita, B. et al. Inflammation and hepatic enceph-
aspartate in acute liver failure: a double‐blind, randomized, placebo‐con- alopathy: ibuprophen restores learning ability in rats with portacaval shunts.
trolled study. Gastroenterology, 2009;136:2159–68. Hepatology, 2007;46:514–9.
77. Marchesini, G., Bianchi, G., Merli, M. et  al. Nutritional supplementation 90. Chastre, A., Belanger, M., Beauchesne, E. et  al. Inflammatory cascades
with branched chain amino acids in advanced cirrhosis: a double‐blind rand- driven by tumor necrosis factor alpha play a major role in the progression of
omized trial. Gastroenterology, 2003;124:1792–801. acute liver failure and its neurological complications. PloS One,
78. Muto, Y., Sato, S., Watanabe, A. et al. Effects of oral branched‐chain amino 2012;7:e49670.
acid granules on event‐free survival in patients with liver cirrhosis. Clin 91. Jiang, W., Desjardins, P., and Butterworth, R.F. Cerebral inflammation con-
Gastroenterol Hepatol, 2005;3(7):705–13. tributes to encephalopathy and brain edema in acute liver failure: protective
79. Sushma, S., Dasarathy, S., Tandon, R.K. et al. Sodium benzoate in the treat- effect of minocycline. J Neurochem, 2009;109:485–93.
ment of acute hepatic encephalopathy: a double‐blind randomized trial, 92. Vaquero, J. and Butterworth, R.F. Mild hypothermia for the treatment of
Hepatology, 1992;16:138–44. acute liver failure: what are we waiting for? Nat Clin Pract Gastroenterol
80. Rockey, D.C., Vierling, J.M., Mantry, P. et  al. Randomized double‐blind, Hepatol, 2007;10:528–9.
controlled study of glycerol phenylbutyrate in hepatic encephalopathy, 93. Michel, H., Solere, M., Granier, P. et  al. Treatment of cirrhotic hepatic
Hepatology, 2014;59:1073–83. encephalopathy with L‐DOPA. A controlled trial. Gastroenterology,
81. Stravitz, R.T., Gottfried, M., Durkalski, V. et al. Safety, tolerability and phar- 1980;29:555–61.
macokinetics of L‐ornithine phenylacetate in patients with acute liver injury/ 94. Uribe, M., Farca, A., Marquez, M.A. et al. Treatment of chronic portal‐sys-
failure and hyperammonemia. Hepatology, 2018;67:1003–13. temic encephalopathy with bromocriptine: a double‐blind controlled trial.
82. Malaguarnera, M., Pisone, G., Astutu, M. et al. L‐carnitine in the treatment Gastroenterology, 1979;76:1347–51.
of mild or moderate encephalopathy. Digest Dis, 2003;21:271–75. 95. Morgan, M.Y., Jacobovits, A.W., James, I.M. et al. Successful use of bro-
83. Ratnakumari, L., Qureshi, I.A., and Butterworth, R.F. Effect of L‐carnitine mocriptine in the treatment of chronic hepatic encephalopathy.
on cerebral and hepatic energy metabolites in congenitally hyperammone- Gastroenterology, 1980;78:663–70.
mic sparse‐fur mice and its role during benzoate therapy. Metabolism, 96. Butterworth, R.F. Parkinsonism in cirrhosis: pathogenesis and current thera-
1993;42:1039–46. peutic options. Metab Brain Dis, 2013;28(2):261–7.
84. Therrien, G., Butterworth, J., Rose, C. et al. Protective effect of L‐Carnitine 97. Sotil, E.U., Gottstein, J., Ayala, E. et al. Impact of preoperative overt hepatic
in ammonia‐precipitated encephalopathy in portacaval‐shunted rats: evi- encephalopathy on neurocognitive function after liver transplantation. Liver
dence for a central mechanism of action. Hepatology, 1997;25:551–6. Transpl, 2009;15:184–92.
85. Pomier Layrargues, G., Giguere, J.F., Lavoie, J. et al. Flumazenil in cirrhotic 98. Bajaj, J.S., Schubert, C.M., Heuman, D.M. et  al. Persistence of cognitive
patients in hepatic coma: a randomized double‐blind placebo‐controlled impairment after resolution of overt hepatic encephalopathy.
crossover trial. Hepatology, 1994;19:32–7. Gastroenterology, 2010;138:2332–40.
86. Als‐Nielsen, B., Gluud, L.L., and Gluud, C. Benzodiazepine receptor antag- 99. Saner, F.H., Gu, Y., Minouchehr, S. et al. Neurological complications after
onists for hepatic encephalopathy. Cochrane Database Syst Rev, cadaveric and living donor transplantation. J Neurol, 2006;253:612–7.
2004;2:CD002798.
The Kidney in Liver Disease
49 Moshe Levi1, Shogo Takahashi1, Xiaoxin X. Wang1, and Marilyn E. Levi2
1
Department of Biochemistry and Molecular and Cellular Biology, Georgetown University,
Washington, DC, USA
2
Department of Medicine, Division of Infectious Diseases, University of Colorado, Aurora, CO, USA

IMPAIRED RENAL FUNCTION kidney’s ability to counterbalance the effects of vasoconstrictors


IN LIVER DISEASE on the renal circulation, (v) increased gut permeability resulting
in bacterial translocation, and (vi) systemic inflammation with
Impaired renal function is common in liver diseases, either as release of cytokines, damage‐associated molecular patterns
part of multi‐organ involvement in acute illness or secondary to (DAMPS), and reactive oxygen species (ROS).
advanced liver disease [1–4]. These hemodynamic factors can accelerate kidney injury in
the presence of intrinsic kidney disease caused by various
disorders.

ACUTE KIDNEY INJURY


IN LIVER DISEASE HEPATORENAL SYNDROME
Acute kidney injury (AKI) previously known as acute renal Hepatorenal syndrome (HRS) is a unique form of functional
failure (ARF) or acute tubular necrosis (ATN) occurs in renal failure that often complicates advanced liver disease,
approximately 20% of hospitalized patients with cirrhosis. In hepatic failure, or portal hypertension. The International Club of
contrast, the incidence of AKI in acute liver failure varies from Ascites has recently updated the diagnostic criteria for AKI and
40 to 80%, especially in the setting of infections such as viral HRS‐AKI [5, 6], which are outlined in Table 49.1.
hemorrhagic fever, leptospirosis, bacterial peritonitis, and
­ In cirrhotic patients with ascites, most of the cases of acute
toxin‐induced injuries such as acetaminophen poisoning, renal failure are mediated by: (i) pre‐renal failure and (ii) acute
administration of nephrotoxic antibiotics (such as aminogly- kidney injury (acute tubular necrosis).
coside antibiotics), or high doses of non‐steroidal anti‐inflam- HRS accounts for approximately 20% of renal failure. HRS
matory drugs. is characterized by: (i) an extreme expression of the profound
The mechanisms leading to acute renal injury are multiple and circulatory dysfunction in cirrhosis, with marked splanchnic
additive and include: (i) changes in the systemic arterial circula- arterial and systemic vasodilatation, insufficient cardiac output,
tion, including systemic arterial vasodilatation and sometimes in severe reduction of effective blood volume, homeostatic activa-
certain alcoholic patients the concomitant presence of cardio- tion of vasoactive systems, and intense renal vasoconstriction,
myopathy, (ii) portal hypertension, (iii) activation of renal vaso- resulting in a critical decrease in renal blood flow and glomeru-
constrictive hormones, including the renin–angiotensin system, lar filtration rate, (ii) absence of pathological changes in renal
the sympathetic nervous system (SNS), arginine ­vasopressin, tissue (vasculitis, glomerulonephritis, or tubulointerstitial
endothelin (ET), thromboxane A2, and leukotrienes, which all ­fibrosis), and (iii) preserved renal tubular function [7].
additively impair renal blood flow, (iv) suppression of renal vas- Precipitating events that cause the development of HRS
odilatory factors including prostacyclins which impair the include: (i) bacterial peritonitis and/or sepsis of other origin,

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
49:  The Kidney in Liver Disease 631

Table 49.1  Current diagnostic and treatment criteria for HRS‐AKI predictors of HRS reversal. Patients received intravenous terli-
(A) Diagnostic criteria of HRS‐AKI pressin 1–2 mg every 6 hours plus albumin or placebo plus albu-
Cirrhosis and ascites min up to 14 days. The pooled analysis comprised 308 patients
Diagnosis of AKI according to ICA‐AKI criteria: increase in SCr (terlipressin: n = 153; placebo: n = 155). HRS reversal was
greater than or equal to 0.3 mg dL−1 within 48 hours
Absence of shock significantly more frequent with terlipressin versus placebo
­
No response after two consecutive days of diuretic withdrawal and (27% versus 14%; P = 0.004). Terlipressin was associated with
plasma volume expansion with albumin (1 g kg−1 of body weight) a more significant improvement in renal function from baseline
No current or recent use of nephrotoxic drugs (NSAIDs,
until end of treatment, with a mean between‐group difference in
aminoglycosides, iodinated contrast media, etc.)
No macroscopic signs of structural kidney injury, defined as: SCr concentration of −53.0 μmol L−1 (P < 0.0001). Lower SCr,
• absence of proteinuria (greater than 500 mg d−1) lower mean arterial pressure, and lower total bilirubin and
• absence of microhematuria (greater than 50 RBCs per high absence of known precipitating factors for HRS were independ-
power field), ent predictors of HRS reversal and longer survival in terlipres-
• normal findings on renal ultrasonography
sin‐treated patients.
(B) Treatment criteria of HRS‐AKI
Terlipressin plus albumin therefore resulted in a significantly
Meeting all the diagnostic criteria of HRS‐AKI
AKI stage greater than or equal to 1B (after plasma albumin expansion) higher rate of HRS reversal versus albumin alone in patients with
No contraindication to vasoconstrictor therapy HRS‐1 (ClinicalTrials.gov identifier: OT‐0401, NCT00089570;
Criteria for treatment individualized REVERSE, NCT01143246).
AKI, acute kidney injury; HRS‐AKI, hepatorenal syndrome‐AKI; ICA, International While terlipressin is widely used in Europe and Asia, it is
Club of Ascites; NSAIDs, non‐steroidal anti‐inflammatory drugs; RBCs, red blood cells. not yet available in the United States and instead intravenous
Reproduced with permission of John Wiley & Sons [5].
albumin infusion plus noradrenaline has been used. The com-
bination vasoconstrictive therapy of midodrine plus octreotide
(ii) increased doses of diuretics, (iii) large‐volume paracentesis
has also been used. However, the combination of midodrine
without volume expansion, and (iv) gastrointestinal hemor-
plus octreotide is not as effective as terlipressin or noradrena-
rhage, which all result in further compromise of the systemic
line [9, 11].
and renal circulation.
Clinically, two different forms of HRS can be distinguished:
type 1 HRS (HRS‐1) is characterized by rapid progression of
renal failure, whereas type 2 HRS (HRS‐2) is characterized by a VIRAL INFECTIONS
less severe and more stable form of renal impairment.
In recent years, there have been several potentially life‐saving There are several viral diseases that affect both the liver and the
and/or survival‐prolonging therapies aimed at patients with kidney, including hepatitis B and hepatitis C. Viral infections
HRS type 1 (HRS‐1) [8–10]. can cause renal injury by a number of mechanisms: (i) circulat-
ing immune complexes that consist of viral antigens and the
Liver transplantation host antiviral antibodies; (ii) in situ immune‐mediated mecha-
Liver transplantation, when possible, is the ideal form of treat- nisms involving viral antigens bound to glomerular structures;
ment for HRS and end‐stage liver disease. Although subjects (iii) expression of viral proteins or pathogenic proinflammatory
with HRS do have more frequent complications both pre‐ and factors including cytokines, chemokines, and growth factors in
post‐operation, including AKI requiring dialysis, nevertheless renal tissue; and (iv) direct cytopathogenic effects on glomeru-
the overall survival rate averages over 50% at 3 years. However, lar and tubulointerstitial cells.
due to ongoing organ shortage, most patients with end‐stage
liver disease and HRS die before a liver becomes available. Hepatitis B
Therefore, alternative temporizing measures become very
important [10]. Aside from hepatitis, hepatitis B causes several extrahepatic
manifestations, including reactive arthritis, skin rashes, vasculi-
tis, and glomerulonephritis.
Use of vasoconstrictors and albumin Three forms of glomerulonephritis are associated with ­hepatitis
There is increasing evidence that shows that the use of vasocon- B: (i) membranous glomerulonephritis is most frequently reported
strictors especially terlipressin in combination with plasma in the Asian population and in children, (ii) membranoprolifera-
expanders such as intravenous albumin may reverse HRS by tive glomerulonephritis, and (iii) IgA nephropathy, most com-
decreasing splanchnic vasodilatation and increasing effective monly seen in adults. The causal role of hepatitis B virus (HBV)
arterial blood volume, resulting in decreases in endogenous infection in glomerulonephritis has been suggested by the finding
renal vasoconstrictors and increased renal perfusion and of viral antigens in immune complex deposits in the glomeruli
­glomerular filtration rate. and virus‐specific cytotoxic T lymphocytes. This is particularly
Recently two randomized (phase 3) studies OT‐0401 and relevant in kidney transplant recipients, where graft function may
REVERSE compared the efficacy of terlipressin plus albumin be impacted.
versus placebo plus albumin in patients with HRS‐1 [8]. Hepatitis B glomerulonephritis has a more favorable course
Pooled patient‐level data were analyzed for HRS reversal in children than adults. Adults with nephrotic syndrome and
(serum creatinine [SCr] value less than or equal to 133 μmol abnormal liver function tests have the worst prognosis and are
L−1], 90‐day survival, need for renal replacement therapy, and associated with progression to end‐stage renal disease (ESRD).
632 THE LIVER: VIRAL INFECTIONS

Lamivudine, emtricitabine, tenofovir, and entecavir inhibit hep- different classes is used to decrease the development of resist-
atitis B DNA polymerase and are used for treatment, with the ance and enhance potency of the regimen. Regimens are deter-
caveat that resistance may develop, particularly with lamivu- mined by the HCV genotype, specifically genotypes 1, 4, 5, or
dine. Therefore, treatment of hepatitis B with these agents will 6. Some DAAs are potentially nephrotoxic such as sofosbuvir
require monitoring for breakthrough viremia. Lamivudine and which is not recommended for patients with chronic kidney dis-
emtricitabine, nucleoside reverse transcriptase inhibitors used ease stage 4 or 5 with eGFR less than 30 or ESRD.
for the treatment of hepatitis B and HIV requires a lower dose Hepatitis C infection prevalence in patients with ESRD ranges
for the treatment of hepatitis B monoinfection. Tenofovir diso- between 3–70% depending on the country. Therefore, all dialysis
proxil fumarate (TDF) which is also used for the treatment of patients require initial HCV antibody screening and if negative,
both hepatitis B and HIV is potentially nephrotoxic. The prod- repeated every six months. If the HCV antibody is positive, hep-
rug TDF is converted rapidly to tenofovir in the gut and circu- atitis C RNA should be obtained to look for evidence of viremia
lates exclusively in plasma and filtered at the glomerulus. TNF as well as a genotype. Treatment should be initiated in the pres-
is also actively transported into proximal renal tubular cells ence of HCV viremia with regimens based on the genotype.
from the interstitial fluid and leads to mitochondrial toxicity and Patients who are infected with hepatitis C will suppress hepa-
proximal tubular injury, resulting in Fanconi’s syndrome. titis B viremia if coinfected. When these patients undergo treat-
Manifestation of Fanconi’s syndrome includes proteinuria, ment of hepatitis C with either PEG‐IFN or directly acting
hypophosphatemia and phosphaturia, normoglycemic glycosu- antivirals, the underlying hepatitis B may reactivate and pro-
ria, uricosuria, hyperuricemia, and aminoaciduria. By contrast, gress [14]. Liver failure from hepatitis B resulting in death and
a newer prodrug of tenofovir, tenofovir alafenamide (TAF) has liver transplantation has been reported. Consequently, the Food
more potent anti‐HIV activity and is not considered nephrotoxic and Drug Administration placed a black box warning on DAAs,
as it does not accumulate in proximal tubular cells [12]. with recommendations to screen all patients with hepatitis C for
Entecavir has similar mechanisms of nephrotoxicity although to any positive serologies for hepatitis B. The most common posi-
a lesser degree. All of these agents are cleared by the kidney and tive serology to be consistent with active hepatitis B is hepatitis
may require dose adjustment. B surface antigen (HBsAg), and if detected, indicates a risk for
progression of hepatitis B when HCV is successfully treated.
Therefore, if hepatitis B treatment is indicated, this should be
Hepatitis C
initiated prior to DAA therapy for hepatitis C to avoid active
More than 80% of subjects acutely infected with hepatitis C hepatitis B ­infection. In the presence of a positive hepatitis B
develop chronic infection and chronic liver disease. Hepatitis C core antibody and negative HBsAg, the risk of hepatitis B
virus (HCV) infection is also associated with high prevalence of reactivation is less likely. In this scenario, HBV DNA moni-
renal disease. HCV‐positive subjects have 40% higher odds for toring is performed monthly and continued for three months
development of chronic kidney disease as compared with HCV‐ after DAA is completed [15].
negative individuals.
Hepatitis C is commonly associated with membranoprolifer-
ative glomerulonephritis, with and without cryoglobulinemia,
Human immunodeficiency virus
membranous glomerulonephritis, focal segmental glomerulo- While HIV‐1 causes a wide spectrum of renal diseases, the role of
sclerosis (FSGS), and polyarteritis nodosa. Hepatitis C infection HIV‐1 in causing primary hepatitis is less certain. Liver function
has also been associated with diabetes‐related nephropathy. abnormalities in HIV‐1‐infected individuals have been associated
Common laboratory findings in subjects with hepatitis with drug adverse reactions and interactions, alcohol abuse, HCV
C‐­associated glomerulonephritis include cryoglobulinemia or HBV coinfections, opportunistic infections such as HHV‐8
(mixed type II), increased monoclonal rheumatoid factor (IgM (associated with Kaposi’s sarcoma), and Hodgkin’s or non‐
kappa), decreased early complements C4, C1q, and CH50, and Hodgkin’s lymphoma. With initiation of antiretroviral therapy
normal or slightly decreased C3. (ART), autoimmune hepatitis and liver failure may develop as a
Previous treatment of hepatitis C included the use of pegylated complication of immune reconstitution syndrome (IRIS), a par-
interferon (PEG‐IFN) and ribavirin with associated adverse adoxical worsening of the individual patient’s HIV‐associated
reactions including fevers, anemia, and limited efficacy. More condition at the time of initiation of antiretroviral therapy and may
recently, PEG‐IFN has been replaced by directly acting antivi- be responsive to corticosteroids. IRIS may develop in response
rals (DAAs) that have revolutionized the treatment of hepatitis to autoantigens associated with autoimmune conditions that may
C, achieving cure in 95–98% of cases, defined as a sustained manifest after antiretroviral therapy is initiated such as systemic
virologic response at twelve weeks, or SVR12. Successful treat- lupus erythematosus, Graves’ disease, sarcoidosis, and rheuma-
ment of HCV often leads to remission of cryoglobulinemic glo- toid arthritis. IRIS may also develop in reaction to underlying
merulonephritis or mixed cryoglobulinemia. In the presence of pathogens such as Mycobacterium tuberculosis or opportunistic
severe vasculitis, rapidly progressive glomerulonephritis or infections such as Pneumocystis jiroveci (formerly P. carinii),
nephrotic syndrome, immunosuppression should be initiated Mycobacterium avium complex, Toxoplasmosis gondii,
early to prevent progression of renal disease [13], including cor- Cryptococcus spp., Histoplasmosis capsulatum, Cytomegalovirus
ticosteroids, rituximab, and plasma exchange [13]. There are (CMV), and other Herpesviridae and JC virus [16]. IRIS has also
three classes of DAAs: (i) NS3/4 protease inhibitors (PIs), been reported to unmask underlying Hodgkin’s and non‐Hodgkin’s
(ii) NS5B polymerase inhibitors, and (iii) NS5A inhibitor class. lymphoma with potential liver infiltration within six months of
Commonly, combination therapy using 1 or more DAAs from initiation of ART. Management of IRIS includes treatment of
49:  The Kidney in Liver Disease 633

opportunistic infections, hepatitis B or C or lymphoma, and con- Treatment of HIV‐associated nephropathy centers around use
tinuation of antiretroviral therapy with or without steroids. of highly active antiretroviral therapy and renin angiotensin
Antiretroviral drug hepatotoxicity has been well described. A aldosterone antagonists.
previous study conducted in subjects with unexplained transam-
inase (ALT and AST) elevations in HIV‐1 monoinfected patients
on antiviral therapy found a high rate of liver lesions, mostly Viral infections in kidney transplant recipients
consistent with non‐alcoholic steatohepatitis (NASH). It is now Prevention of kidney graft rejection requires immunosuppres-
recognized that nucleoside reverse transcriptase inhibitors sive agents that primarily inhibit T‐lymphocyte function.
(NRTIs) including AZT, lamivudine, tenofovir, and emtricit- Consequently, these patients are at risk for viral complications
abine may induce hepatitis steatosis. The NRTI abacavir induces so that vaccinations, including live vaccines should be adminis-
a unique hypersensitivity reaction in patients who are positive tered prior to transplantation such as Zostavax for herpes zoster.
for HLA B5701 and are at risk for liver failure and death. Following transplantation, live vaccines are contraindicated
Therefore, a screening blood test to identify the presence of due  to risk for vaccine‐induced infection in the presence of
HLA B5701 should be performed prior to initiation of abacavir. immunosuppression.
Non‐nucleoside reverse transcriptase inhibitors (NNRTI) such
as efavirenz may be associated with severe liver injury. Despite
Specific issues related to transplantation
these observations, a more recent study following a cohort of
10 083 HIV infected patients in the United States between 1. Reactivation of previous infections that have remained in the
2004–2010 showed infrequent elevations in aminotransferases. latent phase such as cytomegalovirus, herpes simplex, herpes
However, this study also noted that the risk of hepatotoxicity zoster, and Epstein‐Barr viruses, all potentially manifest as
was greater with HIV protease inhibitor use in hepatitis B or C acute hepatitis. Epstein‐Barr virus is also associated with post‐
coinfected patients [17]. transplant lymphoproliferative disorder (PTLD) that may
In HIV positive patients who are coinfected with hepatitis B involve the liver and responds to lowering of immunosuppres-
and/or C, more rapid progression of hepatic decompensation sion. Specific cases with CD20+ markers may respond to
including drug‐induced liver injury is seen compared to HIV rituximab +/− chemotherapy including cyclophosphamide,
monoinfected individuals. Treatment with antiretroviral therapy doxorubicin, vincristine, and prednisone (CHOP). Reactivation
has been shown to control the advancement of hepatic failure may also occur with the use of antilymphocyte antibodies
and is an integral part of HBV and HCV management. Initial such as antithymocyte globulin (ATG) and thymoglobulin and
antiretroviral regimens in coinfected patients are the same as in requires the use of valganciclovir or ganciclovir for CMV
HIV monoinfected patients with the caveat that drug selection prophylaxis. Reactivation may be predicted by the net state of
may require adjustment due to potential drug interactions with immunosuppression, whereby excessive lowering of T and B
directly acting agents. In addition, when treating HCV with cell function may impair the ability to control latent infections,
DAA regimens that include PIs, HIV coinfected patients who resulting in acute reactivation [22]. Vaccination for herpes
are maintained on HIV protease inhibitors should be switched to zoster reactivation
an HIV non‐PI containing regimen. Reactivation of BK virus, a polyomavirus is seen primar-
Kidney disease is a common complication of HIV infection. ily in post‐kidney transplant recipients. This virus is believed
The spectrum of HIV‐associated renal disease includes direct to be acquired during childhood and is maintained in a latent
association with infection, diseases that are linked with the sys- phase in the genitourinary tract. During periods of immuno-
temic response to infection, diseases that develop because of suppression, BK virus may reactivate and cause multiple
superinfections, and diseases that are associated with treatment of complications including ureteral stenosis, hematuria, kidney
HIV infection [12]. Specifically, AKI may result from acute tubu- graft dysfunction, and organ loss. Monitoring of BK in
lar necrosis, acute tubulointerstitial nephritis, rhabdomyolysis, plasma and urine is used to predict graft dysfunction and is
antiretroviral therapy, HIV‐associated nephropathy, HIV immune treated preemptively with lowering of immunosuppression.
complex disease, thrombotic microangiopathies, and urinary tract 2. Donor derived transmission of viral infections: this is com-
obstruction. Chronic kidney disease (CKD) may also result from monly related to the presence of viral antibodies (e.g. CMV
these causes. In addition, there are several potential opportunistic IgG) in the donor and absence of protective antibody in the
infections of the kidney parenchyma, including viruses such as recipient termed mismatches resulting in primary infections.
cytomegalovirus (CMV) and BK, fungal, both typical and atypical Use of prophylaxis to prevent primary infections with the
mycobacterial infections, as well as infiltrative lesions of the kid- appropriate antivirals may prevent donor transmission such
ney that may result from lymphoma and Kaposi’s sarcoma [12]. as valganciclovir or ganciclovir for cytomegalovirus for six
Current antiretroviral therapy regimens may result in chronic months and acyclovir for herpes simplex and zoster.
inflammation, premature aging, and metabolic syndrome and its Prophylaxis for EBV is controversial, and EBV DNA moni-
complications including chronic kidney disease [18]. toring is recommended.
HIV patients of African descent have increased predisposi- 3. Community acquired infections: The most common viral
tion to HIV‐associated nephropathy due to variants in ApoL1 pathogens observed include Herpesviridae such as cytomeg-
gene encoding apolipoprotein L1 [19, 20]. In addition to HIV‐ alovirus, herpes simplex, herpes zoster, and Epstein‐Barr, all
associated nephropathy, APOL1 risk variants are also associated of which may present as primary infections and involve the
with focal segmental glomerulosclerosis and hypertension‐ liver and kidneys. Adenovirus, a cause of hemorrhagic
associated arterionephrosclerosis [19, 21]. ­cystitis, nephritis, and graft failure in addition to hepatitis
634 THE LIVER:  NAFLD, NASH, AND CKD

responds to lowering of immunosuppression and use of cido- population. The rising disease prevalence is expected to result in
fovir, although nephrotoxicity is a significant issue. A prod- increasing number of patients with cirrhosis and end stage liver
rug of cidofovir, brincidofovir is currently available on a disease requiring liver transplantation and in addition an
compassionate use basis for adenoviral infections with no increase in HCC [31].
significant nephrotoxicity reported, although gastrointestinal
intolerance is frequently seen. Influenza infections are a sig-
nificant cause of morbidity and mortality in transplant recipi-
ents, and vaccination is imperative. NAFLD, NASH, AND CKD
There is now increasing evidence that NAFLD, is associated
with a nearly 40% increase in the long‐term risk of incident
METABOLIC SYNDROME AND OBESITY CKD [1–4]. The mechanisms by which NAFLD increases CKD
risk are not fully known. However, several factors including
It is now increasingly recognized that metabolic syndrome and insulin resistance, altered fatty acid and cholesterol metabolism,
obesity affect both the kidney and the liver. The proportion of mitochondrial dysfunction, oxidative stress, ER stress, inflam-
obese adults worldwide has increased to 37% in men and simi- mation, microbiome, altered bile acid metabolism, and increased
larly to 38% in women. In the United States the prevalence of activity of profibrogenic factors can result in progression of
obesity has similarly increased to 39.8% in adults [23]. NAFLD and play a role on increased incidence of CKD
(Figures 49.1 and 49.2).

OBESITY AND KIDNEY DISEASE Insulin resistance


The obesity epidemic has resulted in an increased incidence of Insulin resistance is a common feature of NAFLD that also
obesity‐related glomerulopathy (ORG), a distinct entity pre- contributes to its pathogenesis. For example, insulin resist-
senting as proteinuria, enlarged glomeruli, progressive glomeru- ance in adipose tissue leads to release of fatty acids through
losclerosis, and decline in renal function [24–26]. Pathologically dysregulated lipolysis of triglycerides. Insulin resistant adi-
glomerular hypertrophy and adaptive focal segmental glomeru- pocytes also secrete cytokines like IL6 that has proinflamma-
losclerosis are characteristics of ORG. tory effects in the liver [32, 33]. This results in further
Obesity‐induced increases in glomerular filtration rate, renal augmentation of insulin resistance. In addition, the fatty acids
plasma flow, filtration fraction, and renal tubular sodium reab- released by adipocytes are taken up by the liver, which con-
sorption results in glomerular hypertrophy. tributes to increased lipogenesis and lipid droplet formation
In addition to the hemodynamic changes, insulin resistance, in the liver [32, 34].
altered fatty acid and cholesterol metabolism, mitochondrial dys- Insulin resistance also plays an important role in pathogen-
function, oxidative stress, ER stress, inflammation, microbiome, esis of CKD by modulating renal hemodynamics, glomerular
altered bile acid metabolism, and increased activity of profibro- mesangial cells and podocytes, and renal tubular function [33,
genic factors also play a role in the pathogenesis of ORG. 35]. In insulin resistance despite increased renovascular resist-
Approximately one third of the patients develop progressive ance due to impaired nitric oxide (NO) generation, reduced
increase in proteinuria and decline in renal function, resulting in tubuloglomerular feedback, and dilation of afferent arterioles
ESRD. Renin angiotensin aldosterone system blockade, and due to increased reabsorption of glucose and sodium result in
especially weight loss, and bariatric surgery result in reductions glomerular hyperfiltration. Reduced insulin signaling in
of proteinuria and progression of kidney disease. mesangial cells results in mesangial cell hypertrophy, hyper-
Studies in experimental models of diet‐induced obesity and trophy, and extracellular matrix deposition. Insulin resistance
kidney disease indicate that activation of the bile acid regulated in podocytes results through lysosomal and proteasomal deg-
nuclear receptor farnesoid X receptor (FXR) and/or G protein‐ radation of the insulin receptor by a nephrin dependent mecha-
coupled receptor TGR5 protects against kidney injury. The nism [36]. This induces podocyte apoptosis, effacement of its
mechanisms of action include prevention of lipid accumulation, foot processes, albuminuria, thickening of the glomerular
oxidative stress, inflammation, ER stress, mitochondrial basement membrane, and increased glomerulosclerosis. The
­dysfunction, and fibrosis [27–30]. renal tubular consequences of insulin resistance are more com-
plex but include gluconeogenesis in the proximal tubule and
increased sodium absorption in the distal tubule [33, 35].
While the effects of insulin resistance per se in regulation of
OBESITY AND LIVER DISEASE proximal tubule glucose reabsorption via SGLT‐2 is debatable,
nevertheless SGLT‐2 expression and activity are increased in
Obesity and metabolic syndrome, especially in the presence of humans and animal models of obesity and diabetes [37]. SGLT‐2
diabetes mellitus, has become the leading cause of non‐alco- inhibitors including empagliflozin and canagliflozin have renal
holic fatty liver disease (NAFLD), in its whole spectrum of dis- and cardiovascular disease beneficial effects in human subject.
ease ranging from simple steatosis to NASH, cirrhosis, and Furthermore SGLT‐2 inhibitors decrease renal disease and liver
hepatocellular carcinoma (HCC). The prevalence of NAFLD is disease in an animal model of diet‐induced obesity and insulin
projected to increase from the current estimate of 25% of the resistance [38].
49:  The Kidney in Liver Disease 635

Figure 49.1  Organ crosstalk in the pathophysiology of non‐alcoholic fatty liver disease (NAFLD) and chronic kidney disease (CKD). Many
­factors, such as caloric intake, dietary factors (such as high fructose consumption and low vitamin D levels), genetic factors, and inflammation of
visceral adipose tissue can increase an individual’s risk of NAFLD. Increased amounts of non‐esterified fatty acids (NEFAs) resulting from the
expansion of intra‐abdominal visceral adipose tissue are associated with an increase in NF‐κB and inflammatory pathways, dysregulation of
­adipokine production leading to decreased adiponectin levels, and impaired insulin signaling. Progression of liver dysfunction triggers pathways
that might influence the development of CKD. For example, insulin resistance and atherogenic dyslipidemia as well as proinflammatory factors,
prothrombotic factors, and profibrogenic molecules can promote vascular and renal damage. Reduced activation of the energy sensor, 5′‐AMP
activated protein kinase (AMPK), in response to reduced adiponectin levels further stimulates proinflammatory and profibrogenic mechanisms.
Activation of the renin‐angiotensin system (RAS) and endothelial cells might also contribute to liver and kidney dysfunction by increasing oxidative
stress, inflammation, and coagulation pathways. Increased production of uremic toxins by intestinal microbiota in the setting of CKD might induce
further renal, liver, and cardiovascular damage through inflammatory, oxidative, and fibrotic pathways. Dysbiosis of the intestinal microbiota,
which often occurs with obesity, potentially influences NAFLD, CKD, and type 2 diabetes mellitus (T2DM) through complex mechanisms. Finally,
cardiovascular disease (CVD), the risk of which is increased in the setting of NAFLD, T2DM, or intestinal dysbiosis, can influence the development
of renal dysfunction (and vice versa) through cardiorenal interactions. AGE, advanced glycation end‐product; NASH, non‐alcoholic steatohepatitis;
ROS, reactive oxygen species; SCFA, short‐chain fatty acid; TMAO, trimethylamine oxide. Reproduced with permission of Springer Nature, Figure
2 in [1].

Lipid metabolism mammalian FATP isoforms, FATP2 and FATP5 are found pri-
Excessive lipid accumulation can occur ectopically in nonfat marily in the liver and their knockdown in mice decreases uptake
tissue, contributing to their damage through toxic processes of fatty acids and ameliorates hepatic steatosis. CD36 facilitates
named lipotoxicity. Liver lipid accumulation results from an the transport of long‐chain fatty acids and is regulated by peroxi-
imbalance between lipid acquisition (uptake of circulating fatty some proliferator‐activated receptor (PPAR) γ, ­ pregnane X
acids and de novo lipogenesis [DNL]) and lipid removal (fatty receptor, and liver X receptor (LXR). CD36 expression in the
acid oxidation [FAO], export as a component of very low‐­ liver is increased in NAFLD and is thought to ­stimulate increased
density lipoproteins [VLDL] particles, lipid droplet formation uptake of fatty acids. The caveolins comprise a family of three
and lipolysis) (Figure 49.3). membrane proteins for lipid trafficking and lipid droplets forma-
The hepatic lipid uptake is largely dependent on fatty acid tion. Caveolin 1 is increased in the liver of mice with NAFLD,
transporters [39], predominately mediated by fatty acid transport and mainly localized in the centrilobular zone three, where the
proteins (FATPs), cluster of differentiation 36 (CD36), and cave- steatosis is most severe. Following uptake, fatty acids in the cyto-
olins located in the hepatocyte plasma membrane. Of the six sol need specific intracellular fatty acid binding proteins (FABP)
636 THE LIVER:  NAFLD, NASH, AND CKD

Figure 49.2  Potential mechanisms by which intestinal dysbiosis might promote the development of non‐alcoholic fatty liver disease (NAFLD)
and chronic kidney disease (CKD). An imbalance in intestinal microbiota (dysbiosis) resulting in an increase in Gram‐negative bacteria leads to an
increase in lipopolysaccharide (LPS) production, which can damage the intestinal epithelium resulting in increased gut permeability. This increased
gut permeability enables further egress of bacterial content, LPS, and small molecules into the portal and systemic circulations causing inflamma-
tion. Dysbiosis also promotes the increased production of secondary bile acids such as deoxycholic acid. Following their return to the liver as part
of the enterohepatic circulation, secondary bile acids have been linked to chronic inflammation, cholestasis and carcinogenesis through their actions
on farnesoid X Receptor (FXR; with downstream effects on cholesterol metabolism) and by inducing cellular senescence and causing damage to
mitochondrial and cell membranes. The intestinal microbiota also generates molecules such as trimethylamine (TMA), p‐cresol, and indole from
dietary choline, phenylalanine/tyrosine, and tryptophan, respectively. TMA is oxidized in the liver to trimethylamine oxide (TMAO), which pro-
motes atherosclerotic vascular disease. Indole and p‐cresol are metabolized in the liver to indole sulfate and p‐cresol sulfate, which are cleared by
the proximal tubules and are potentially nephrotoxic. Other molecules produced by microbiota metabolism, such as phenylacetic acid and hippuric
acid, are also potentially toxic to the kidney. Reduced levels of short‐chain fatty acids (SCFAs) as a consequence of dysbiosis can also induce
decreased lipogenesis and increased gluconeogenesis, leading to insulin resistance, further dysfunction of the liver and kidney, and to the develop-
ment of type 2 diabetes mellitus (T2DM). Reproduced with permission of Springer Nature, Figure 3 in [1].

to facilitate their transportation, s­ torage, and utilization. FABP1, LXRα, and carbohydrate regulatory element‐binding protein
also known as liver‐type FABP, is the predominant isoform in the (ChREBP), which is activated by carbohydrates, play critical roles
liver. Its expression is increased in NAFLD but declines as the in this process. SREBP1c expression is enhanced in patients with
disease progresses to NASH. NAFLD, while SREBP1c knockout mice display decreased
DNL results from the synthesis of new fatty acids from acetyl‐ expression of lipogenic enzymes. ChREBP is the bona fide tran-
CoA subunits produced primarily through glycolysis and the scription ­factor that is primarily responsive to glucose. ChREBP
metabolism of carbohydrates. In addition to glucose which most ­regulates not only enzymes in glucose metabolism, but also lipo-
commonly supplies carbon units for DNL, fructose also produces genic enzymes. Therefore, ChREBP causes complex metabolic
acetyl‐CoA that can enter the lipogenic pathway, important when changes in loss‐ and gain‐of function studies. ChREBP expression
considering the increasing use of fructose in corn syrup as a sweet- in liver biopsies from patients with NASH is increased when stea-
ener. DNL starts with acetyl‐CoA that is converted to malonyl‐ tosis is found to be greater than 50% but decreased in the presence
CoA by acetyl‐CoA carboxylase (ACC) and further converted to of severe insulin resistance [40], indicating ChREBP might disso-
palmitate by fatty acid synthase (FASN). New fatty acid may then ciate hepatosteatosis from insulin resistance. Recent studies fur-
undergo a range of desaturation, elongation, and esterification ther illustrate how different post‐­translational modifications, such
steps before ultimately being stored as triglycerides in lipid drop- as phosphorylation, acetylation, O‐GlcNAcylation, or ubiquitina-
lets or exported as VLDL particles. Dysregulated DNL is a central tion, operate together in regulating lipogenic gene transcription.
feature of liver lipid accumulation in NAFLD patients [31]. For example, deacetylation of SREBP‐1c by SIRT1 inhibited
Lipogenic genes are coordinately regulated at the transcription binding to its target lipogenic promoters [41]. Insulin signaling
level. Transcription factors, such as sterol regulatory element‐ triggers activation of series of kinases downstream of PI3K, such
binding protein 1c (SREBP1c), which is activated by insulin and as Akt and mTORC1/2, to regulate SREBP‐1c ­activity [42].
49:  The Kidney in Liver Disease 637

Figure 49.3  The substrate‐overload liver injury model of NASH pathogenesis. Free fatty acids are central to the pathogenesis of NASH. Free fatty
acids that originate from lipolysis of triglyceride in adipose tissue are delivered through blood to the liver. The other major contributor to the free
fatty acid flux through the liver is DNL, the process by which hepatocytes convert excess carbohydrates, especially fructose, to fatty acids. The two
major fates of fatty acids in hepatocytes are mitochondrial beta‐oxidation and re‐esterification to form triglyceride. Triglyceride can be exported
into the blood as VLDL or stored in lipid droplets. Lipid droplet triglyceride undergoes regulated lipolysis to release fatty acids back into the hepato-
cyte free fatty acid pool. PNPLA3 participates in this lipolytic process, and a single‐nucleotide variant of PNPLA3 is strongly associated with
NASH progression, underscoring the importance of the regulation of this lipolysis. When the disposal of fatty acids through beta‐oxidation or for-
mation of triglyceride is overwhelmed, fatty acids can contribute to the formation of lipotoxic species that lead to ER stress, oxidant stress, and
inflammasome activation. These processes are responsible for the phenotype of NASH with hepatocellular injury, inflammation, stellate cell activa-
tion, and progressive accumulation of excess extracellular matrix. Lifestyle modifications that include healthy eating habits and regular exercise
reduce the substrate overload through decreased intake and diversion of metabolic substrates to metabolically active tissues and can thereby prevent
or reverse NASH. SCD, stearoyl CoA‐desaturase; FAS, fatty acid synthase; NKT, natural killer T cell; Tregs, regulatory T cells; PMNs, polymor-
phonuclear leukocytes. Reproduced with permission of Springer Nature, Figure 1 in [31].

FAO is controlled by nuclear receptor PPARα and occurs mitochondrial ultrastructure were increased alongside FAO in
mainly in the mitochondria, providing a source of energy to gen- patients with NASH. Antioxidant markers, glutathione, glu-
erate ATP especially when circulating glucose concentrations are tathione peroxidase, and superoxide dismutase were decreased
low [31]. Activation of PPARα induces the transcription of a in liver biopsies from NAFLD patients and in mitochondria from
range of genes related to FAO thereby reducing hepatic lipid lev- animal models of NAFLD [43].
els. PPARα expression modulates not only lipid ­homeostasis, but Hepatic lipid can be exported from the liver into the blood after
inflammation as well. Expression of genes related to FAO was being packed into water‐soluble very low‐density lipoprotein
higher in patients with more severe steatosis compared to patients (VLDL) particles alongside cholesterol, phospholipids, and apoli-
with less severe steatosis or non‐steatotic controls. FAO, meas- poproteins [44]. The assembly of VLDL particles occurs in the ER
ured indirectly as plasma β‐hydroxybutyrate levels, was higher where apolipoprotein B100 (apoB100) is lipidated by the enzyme
in patients with NASH compared to steatosis or normal controls. microsomal triglyceride transfer protein (MTTP). apoB100 and
Increased FAO may be an adaptive response in patients with MTTP are key components in hepatic VLDL secretion. In patients
NAFLD attempting to reduce the lipid overload and lipotoxicity, with genetic defects in the apoB or MTTP gene, hepatic steatosis
but it also produces ROS and excessive FAO may exhaust the is common due to compromised triglyceride export.
capacity of the antioxidant defense system and induce oxidative Excessive lipid not secreted into the blood can form lipid
stress. Accordingly, hepatic oxidative stress and changes in droplets in hepatocytes, a defining feature of NAFLD. The
638 THE LIVER:  NAFLD, NASH, AND CKD

triglycerides stored in lipid droplets have to be mobilized via expression is also found in the glomeruli of patients with obesity
hydrolysis to release fatty acids for utilization [45]. Triglyceride related glomerulopathy [51]. In a search for in vivo inhibitors of
hydrolysis consists of sequential reactions of adipose triglycer- SREBP‐1, the FXR agonists have been demonstrated to inhibit
ide lipase (ATGL, annotated as patatin‐like phospholipase SREBP‐1 expression, lipid accumulation, inflammation, and
domain containing protein 2, PNPLA2 or desnutrin) hydrolyz- fibrosis in animal models of diet‐induced obesity and insulin
ing triglycerides to diacylglycerol (DG) and fatty acid, resistance [28, 52, 53]. Renal triglyceride accumulation can also
­hormone‐sensitive lipase (HSL) breaking down DG to monoa- occur by increased uptake via CD36 or FATPs. Renal CD36
cylglycerol (MG) and fatty acid and monoacylglycerol lipase expression is upregulated in patients with CKD, particularly those
(MAGL) completing the lipolytic reaction of MG to glycerol with diabetic nephropathy [54]. Blockade or knockout of CD36
and fatty acid [45]. Various studies suggest a critical role of can prevent kidney injury in experimental animals. Decreased
intracellular lipid droplet homeostasis in regulating hepatic fat FAO mediated by PPARα and its target enzymes is also responsi-
content. A tight regulation between lipid synthesis, hydrolysis, ble for renal triglyceride accumulation and is reported in kidney
secretion, and FAO is required to prevent hepatic lipid overload biopsy samples from patients with ORG and diabetes [55]. In this
and subsequent intracellular lipid accumulation, leading to regard, PPARα agonist fenofibrate has been shown to prevent
steatosis, lipotoxicity, and liver damage, and promoting disease development of kidney disease [24].
progression and fibrosis. Studies in mice with high fat‐induced obesity have shown
The specific lipotoxic lipids that promote NASH phenotype increased cholesterol synthesis and accumulation as well as
include DG, ceramides, and lysophosphatidylcholine (LPC) the  development of renal disease [49, 56] associated with
species [31]. Hepatic free cholesterol is also considered a criti- increased expression and activity of SREBP‐2, a master regulator
cal lipotoxic molecule in NASH [46]. The main pathways by of cholesterol synthesis and cholesterol metabolism. Statin treat-
which hepatocytes acquire cholesterol, include endogenous ment has also been shown to reduce lipid accumulation in the
synthesis of cholesterol via master regulator SREBP‐2 and its proximal tubules in high‐fat diet‐fed mice. Increased cholesterol
target 3-hydroxy-3-methylglutaryl-CoA reductase (HMGCR), accumulation can also occur because of decreased cholesterol
the rate‐limiting enzyme; uptake of LDL and chylomicron rem- efflux. The nuclear receptor LXR plays a major role in cholesterol
nants via LDL receptor‐mediated endocytosis and subsequent efflux and its expression and target enzymes that regulate choles-
processing through the endosomal/lysosomal compartment; and terol efflux, including ABCA1 and ABCG1, are decreased in
uptake of HDL cholesterol directly via scavenger receptor class human kidney biopsy samples from diabetic patients with ORG
B type I (SR-BI). The main pathways by which cholesterol can [55] and in ­animal models of diabetes [57]. Treatment with LXR
be removed from hepatocytes, include conversion to bile acids agonists or induction of cholesterol efflux with cyclodextrin can
by the rate‐limiting enzymes CYP7A1 and CYP27A1 and reduce cholesterol accumulation and improves renal disease in
excretion of bile acids into bile by bile salt export pump (BSEP); diabetic mice.
excretion of cholesterol into bile by ATP‐binding cassette sub‐
family G member 5 and 8 (ABCG5/G8); incorporation into
VLDL and secretion into the circulation; and efflux of choles-
Mitochondrial dysfunction
terol onto circulating apolipoprotein AI and nascent HDL parti- Lipotoxicity is mediated by different cellular mechanisms
cles. Extensive dysregulation of cholesterol homeostasis has including direct damage to mitochondria [58]. Mitochondria,
been documented in NAFLD, causing both increased synthesis the powerhouse of the cells, play a pivotal role in the final oxi-
and uptake of cholesterol as well as decreased removal of cho- dation of metabolites such as fatty acids and glucose. The mito-
lesterol, and leading to increased hepatic cholesterol levels. chondrial oxidative phosphorylation (OXPHOS) is responsible
Among the agents that target lipotoxicity [31, 47], aramchol, for the production of most cellular total energy. In addition, it is
is an antagonist for stearoyl-CoA desaturase-1 (SCD-1), which is also involved in the regulation of intrinsic signaling pathways of
an enzyme that catalyzes a rate‐limiting step in the synthesis of apoptosis. Since the first studies on NAFLD, much evidence
monounsaturated fatty acids such as oleic acid. This agent also pointed out that it was primarily characterized by the presence
activates cholesterol efflux by stimulating the ABCA1 trans- of mitochondrial dysfunction [59]. The alteration of mitochon-
porter, a widely expressed cholesterol export pump. A multi- drial functions is evident with electron microscopy analysis by
center phase 2b trial (NCT02279524) in NASH patients is some ultrastructural changes such as megamitochondria, loss of
ongoing. For another enzyme in DNL, ACC, two inhibitors cristae, and paracrystalline inclusion bodies in the matrix.
PF‐05221304 and GS‐0976 are in clinical trials in patients with When the disposal of fatty acids through FAO or formation of
NAFLD and NASH. triglyceride is overwhelmed in NAFLD, fatty acids can contribute
In the kidney, abnormal lipid metabolism promotes increased to the formation of lipotoxic species that lead to oxidative stress
triglyceride and cholesterol accumulation [24]. Renal triglyceride and inflammation [31]. Oxidative stress can lead to the peroxida-
accumulation can occur because of increased fatty acid synthesis tion of phospholipids, such as cardiolipin (diphosphatidyl glyc-
mediated by SREBP‐1c and its target enzymes including ACC, erol), one of the major phospholipids in the inner mitochondrial
FASN, and SCD‐1; and/or mediated by ChREBP and its target membrane. Since cardiolipin is assumed to enhance the respira-
enzymes including liver‐pyruvate kinase (L‐PK). In the murine tory chain activity, particularly of the complex I, the oxidation of
high‐fat diet‐induced obesity model SREBP‐1 expression and cardiolipin leads to an imbalance of OXPHOS. Oxidative stress‐
activity results in increased renal lipid accumulation and renal induced damage to mitochondrial DNA can also result in the
disease [48–50]. The effects of high‐fat diet on kidney disease are impairment of OXPHOS and further increases the generation
prevented in SREBP‐1c knockout mice. Increased SREBP‐1 of  ROS. The mitochondrial damage may eventually lead to
49:  The Kidney in Liver Disease 639

apoptotic death of hepatocytes. An increased expression of apop- incomplete or suboptimal FAO leads to accumulation of long
totic proteins and enzymes, such as Bcl‐2, was found in the chain acylcarnitines, ceramides, and diacylglycerols, lipotoxic
murine models of fatty liver disease. The apoptosis along with the intermediates that may act as indirect sources of ROS. In addi-
generated cytokines from the liver stellate and Kupffer cells fur- tion to pro‐oxidant mechanisms, in an experimental model of
ther augment the fibrotic changes to advance the disease. NASH, a decreased activity of several detoxifying enzymes
Sirtuins are a group of nicotinamide adenine dinucleotide was observed. Glutathione peroxidase (GPx) activity is
(NAD)‐dependent deacetylases, share multiple cellular functions reduced probably in consequence of GSH depletion and
related to mitochondrial energy homeostasis and antioxidant impaired transport of cytosolic GSH into the mitochondrial
activity in liver and kidney diseases [60, 61]. SIRT1, the most matrix [64]. The polymorphism C47T of the SOD2 gene,
studied of these enzymes, has an indirect regulatory effect on oxi- encoding for manganese superoxide dismutase, is associated
dative stress, activating forkhead proteins and PGC‐1α, transcrip- with a reduction of activity of this enzyme resulting in an
tion factors involved in transcription of mitochondrial biogenesis increased ROS production and a high susceptibility to devel-
genes and antioxidant enzyme genes and in ROS‐detoxifying oping NASH and advanced fibrosis in NAFLD [65].
capacity. SIRT1 level is decreased in a rat model of NAFLD. The oxidative stress‐induced impairment of mitochondrial
SIRT3 localizes in the mitochondrial matrix and can increase permeability can generate an influx of calcium (Ca2+) and
FAO by activation of long‐chain acyl‐CoA dehydrogenases, and iron into the mitochondria. Iron in the presence of H2O2
its activity is found decreased in animal models with fatty liver. favors Fenton’s reaction to generate more hydroxyl radicals.
NAD+, as the co‐substrate for SIRT1 and SIRT3, controls various Further, Ca2+ can stimulate the inducible NO synthase‐medi-
metabolic improvements. In this way, NAD+ depletion may con- ated nitric oxide radical (NO●) production and apoptotic cell
tribute to mitochondrial dysfunction, obstructing the adaptive death. A significant increase in the NO● level during severe
response mediated by sirtuins to high hepatic lipid levels. hepatocyte injury may induce the progression to necrotic
Hence, alleviation of the mitochondrial impairment, particu- cell  death [66]. The peroxynitrite produced from NO● and
larly in the early stages of NAFLD, may prevent the progression of superoxide radical is one of the important mediators of free
the disease. Among the various experimentally studied mitochon- radical toxicity.
drial‐targeted agents, triphenylphosphonium cation ligated ubiqui-
none Q10 and vitamin E, Szeto‐Scheller peptides, and superoxide
dismutase mimetic‐salen manganese complexes (EUK‐8 and
Endoplasmic reticulum stress
EUK‐134) have been found to be most promising [47]. The ER stress and the unfolded protein response (UPR) play
Mitochondrial dysfunction is also the main cause of renal important roles in NAFLD/NASH and the associated kidney
pathology induced by high fat diet. Given that the kidney is an disease. The activation of ER stress may initially be protective,
organ that demands continuous high‐energy provision, mostly while chronic ER stress may result in apoptosis and fibrosis.
from FAO, lipid overload and impaired FAO lead to a distur- The UPR activation in the ER is mediated by three major sign-
bance in fatty acid uptake and utilization, further aggravating aling pathways, including activating transcription factor 6
lipid accumulation in kidney cells and tissue [62]. This will cre- (ATF6), inositol‐requiring enzyme 1α (IRE1α), and PRKR‐
ate a vicious cycle whereby lipids can damage mitochondria and like ER kinase (PERK). Upon accumulation of misfolded pro-
generate ROS, further limiting mitochondrial FAO and causing teins in the ER, BiP (GRP78) disassociates from the sensors
more cellular lipid accumulation, resulting in more mitochon- and binds to unfolded proteins, which then activates the
drial ROS levels. As an approach to target mitochondria dys- sensors.
function, activation of membrane bile acid receptor TGR5 has In the liver while lipid accumulation in hepatocytes can ini-
been shown to decrease mitochondrial ROS generation and tiate ER stress, activation of ER stress can also induce lipogen-
increase mitochondrial biogenesis, mitochondrial antioxidant esis and thus creating a vicious cycle for the progression of
generation, and mitochondrial FAO in kidney disease in obesity steatosis. ER stress induces insulin induced gene (INSIG) deg-
and diabetes [29]. radation and activation of the SREBP pathway [32, 67]. In
addition, ER stress also induces increased expression of the
VLDL receptor which results in increased lipoprotein delivery
Oxidative stress to the liver [32]. Increased lipid accumulation and chronic ER
An overload of fatty acids into mitochondria, after an increased stress in the hepatocyte results increased inflammation and
intake or an insulin‐resistance condition, may lead to an causes hepatocyte apoptosis. In addition, ER stress also stimu-
increase in the permeability of the inner mitochondrial mem- lates hepatic stellate cell (HSC) collagen I secretion, which
brane. This occurrence leads to the dissipation of the mem- mediates fibrosis [68]. CHOP, the C/EBP homologous protein,
brane potential and the loss of ATP synthesis capacity, resulting plays an important role in ER stress induced apoptosis and
in mitochondrial function impairment and an enhanced ROS fibrosis [68].
generation. The increase of FAO, inducing an increased elec- In the kidney chronic ER stress and UPR activation [69], as it
tron flux in the electron transport chain (ETC), may generate occurs in the setting of obesity and the metabolic syndrome
an “electron leakage” due to reduction of the activity of ETC [30], contributes to podocyte injury, apoptosis, proteinuria, and
complexes, thus ensuring a direct reaction between electrons CKD. Furthermore, albumin has been shown to induce ER
and oxygen, leading to the formation of ROS, rather than the stress in renal tubular cells by increasing intracellular calcium
normal reaction mediated by cytochrome c oxidase that com- levels, inducing UPR‐dependent upregulation of lipocalin 2,
bines oxygen and protons in order to form water [63]. The resulting in apoptosis.
640 THE LIVER:  NAFLD, NASH, AND CKD

Inflammation Although obesity and NAFLD injure the kidney, growing


e­vidence suggests CKD may also contribute to NAFLD and
Inflammation is a physiological response to tissue injury or insulin resistance. Taken together, increased levels of lipids
infection that leads to secretion of various inflammatory media- and  the increased inflammatory response are both thought to
tors, such as cytokines, chemokines, and eicosanoids, which be important pathogenetic factors that are not only involved in
coordinate cellular defense mechanisms and tissue repair. The the development of NASH but are also significant factors in the
persistence of inflammatory activity over time results in chronic development and progression of CKD.
inflammatory changes that exacerbate tissue injury and may
result in an abnormal wound healing response. In the case of
Microbiome
NAFLD, inflammation contributes to the development of NASH
and liver fibrosis. NAFLD refers to a spectrum of histological The liver is the key metabolic organ exposed to high levels of
abnormalities in the liver, ranging from isolated steatosis with intestinal products through the tight bidirectional linkages in the
no or minimal inflammatory activity and no evidence of cell biliary tract, portal vein, and systemic circulation. An altered
damage (NAFL) to NASH, which is characterized by steatosis, microbiome (or “dysbiosis”) may lead to an increase in intesti-
inflammation, and hepatocellular injury, hallmarked by the nal permeability, which can amplify many of these effects
presence of hepatocyte ballooning, with different degrees of derived from intestine [80].
fibrosis [70]. Many factors are associated with chronic systemic Studies of the microbiome are relatively nascent; however, it
immune response, a phenotype that is common in many indi- is anticipated that there will be considerable progress in linking
viduals with NAFLD, CKD, T2DM, or CVD. A central issue in its role to NASH. The changes in specific microbial products,
this field is identification of those factors that trigger inflamma- secondary to altered gut microbial composition, and the changes
tion, thus fueling the transition from non‐alcoholic fatty liver to in intestinal permeability and function can affect hepatic struc-
NASH [71]. These triggers of liver inflammation may have their ture and function to further increase the risk of NAFLD. Patients
origins inside as well as outside of the liver. with NAFLD have a higher prevalence of microbial dysbiosis.
Lipotoxicity is one of the major mechanisms underlying Human studies have documented that the gut microbiome
hepatocyte dysfunction leading to disease progression in among patients with NASH is less complex than that of healthy
NASH [72]. Increased hepatocyte free fatty acid (FFA) influx subjects [81, 82] and indicate that weight loss also alters the
results in the formation of lipotoxic intermediates in the liver, microbiome. Adults with NAFLD showed increased serum tri-
such as ceramides, diacylglycerols, LPC, and oxidized fatty methylamine N‐oxide (TMAO) [83] and decreased production
acid and cholesterol metabolites, which act as ROS [73]. of phosphatidylcholine [84]. Plausible mechanistic links
Recent studies have revealed that hepatocyte inflammasome between an altered microbiome and fatty liver are emerging and
activation may be an important link between the initial meta- include the potential for bacterial proteins to function as ligands
bolic stress and subsequent hepatocyte death and stimulation for G protein‐coupled receptors [85, 86] (Figure 49.4).
of fibrosis in NASH [74]. The accumulation of free cholesterol Dysbiosis has been described in patients with obesity [87] or
in Kupffer cells and hepatic stellate cells can also activate the other features of metabolic syndrome [88] and in those with
NLRP3 inflammasome [75]. established NAFLD [89], T2DM [90, 91], CVD [92], or CKD
While liver and kidney are important organs for lipid metabo- [93]. Several potential pathways, factors, and processes might
lism, the increased proinflammatory cytokine production by link dysbiosis, or mediators of the gut microbiota, and NAFLD
lipotoxicity, which causes ectopic lipid deposition that occurs in to CKD risk factors and vascular and renal diseases (Figure 49.2)
more advanced forms of NAFLD is also likely to have a patho- [1]. Dysbiosis can potentially influence NAFLD, CKD, and
genetic role in the development of extra‐hepatic complications, obesity through multiple and complex mechanisms. It has been
such as CVD and CKD. reported that changes in the gut microbiome of patients with
Deterioration to NASH results in the production or activa- CKD relate to lower levels of Bifidobacteriaceae and
tion of multiple inflammatory mediators such as NF‐κB path- Lactobacillaceae and to higher levels of Enterobacteriaceae
way, cytokines [76], lipopolysaccharides (LPS), and ROS, [94]. Microbial fermentation of dietary fiber in the intestine by
which can lead to insulin resistance, endothelial dysfunction, anaerobic bacteria such as Bifidobacteria and Lactobacilli
and a tissue inflammatory infiltrate that can amplify systemic results in the formation of short‐chain fatty acids (SFCAs)
chronic inflammation. A study in patients with T2DM, with or including acetate, propionate, and butyrate, which have the
without persistent hepatic inflammation (owing to chronic potential to influence hepatic lipogenesis and gluconeogenesis.
HBV infection) suggested that the presence of liver inflamma- Thus, decreased levels of SCFAs in the setting of NAFLD might
tion is a key mediator in the increased risk of CKD. Although contribute to the development of liver adiposity and hepatic
the presence of T2DM undoubtedly increases the risk of CVD insulin resistance [95]. The intestinal microbiota also produces
in patients with NAFLD, several studies have shown that trimethylamine, p‐cresol, and indole from dietary nutrients such
potential mediators of vascular and renal damage occur more as choline, phenylalanine/tyrosine, and tryptophan, respectively.
frequently in patients with NAFLD regardless of whether or Further metabolism in the liver by oxidation or sulfation pro-
not they also have T2DM [77]. Administration of sera from duces ionically charged water‐soluble molecules, such as
patients with CKD or uninephrectomy to rodents or cultured TMAO, p‐cresol sulfate, and indole sulfate, which can be
adipocytes induces lipodystrophy, ectopic fat redistribution excreted in the urine. Plasma TMAO levels are elevated in
from adipose tissue to liver and muscle, insulin resistance, and patients with CKD and portend poorer long‐term survival [96].
glucose intolerance [78, 79]. Indole sulfate, which is cleared by the proximal tubules, is
Figure 49.4  Interplay between the liver and gut microbiota in alcoholic liver disease and NAFLD. Intestinal dysbiosis and bacterial overgrowth
are observed in both alcoholic liver disease (ALD) (part a) and non‐alcoholic fatty liver disease (NAFLD) (part b). Bacterial overgrowth causes an
increase in secondary bile acids (BAs), which modulates farnesoid X receptor (FXR)‐mediated hepatic synthesis of BA, leading to an overall
increase in hepatic BA synthesis. A reduction in hepatic phosphatidylcholine is also seen in both ALD and NAFLD, which causes triglyceride
accumulation in the liver (fatty liver). While ALD‐associated dysbiosis is characterized by reduction in Lactobacillus and Candida overgrowth,
patients with NAFLD have a higher abundance of Lactobacillus (effects on fungal population remain to be investigated). In ALD and NAFLD,
increased ethanol and its metabolite acetaldehyde in the intestinal lumen mediate weakening of intestinal tight junctions. Consequently, increased
translocation of microbial‐associated molecular patterns (MAMPs) (seen in ALD and NAFLD) and gut metabolites, such as acetaldehyde, acetate
(seen in ALD) and trimethylamine (TMA, seen in NAFLD), elicits intestinal and hepatic inflammatory responses, leading to progressive liver dam-
age. AMP, antimicrobial peptides; EtOH, ethanol; HFD, high‐fat diet; LCFAs, long‐chain fatty acids, TMAO, trimethylamine N‐oxide. Reproduced
with permission of Springer Nature, Figure 3 in [86].
642 THE LIVER:  REFERENCES

proinflammatory and potentially toxic to the kidneys as it protein‐coupled receptor ubiquitously expressed with high
increases the risk of tubulointerstitial fibrosis [97]. expression in gallbladder, placenta, lung, spleen, intestine, liver,
It is now widely accepted that liver and kidney damage can brown and white adipose tissue, skeletal muscle, and bone mar-
result from extensive interaction with the gut microbiota through row [109]. TGR5 is activated mainly by the secondary bile acids
specialized molecules, such as TMA. As the role of the microbiota LCA and DCA [110]. TGR5 signaling controls glucose homeo-
in liver disease progression, prognosis, and treatment is increas- stasis by increased energy expenditure in brown adipose tissue
ingly recognized, it is necessary to make focused and conscious and muscle and by increased GLP‐1 release in intestinal L cells
efforts to study the effect of the microbiome to efficiently address [111]. At present it is unclear whether FXR and TGR5 signal
the socioeconomic burden of this spectrum of liver diseases. directly or indirectly via alterations in the microbiota and bile
Increasing evidence suggests that the liver and kidneys share path- acid metabolism that produce agonistic or antagonistic signaling
ways including microbiome and dysbiosis that are intrinsically through each other.
linked to each other. These indicate that the prevalence of CKD is Glucagon‐like peptide (GLP)‐1 is an intestinal hormone gen-
markedly increased among patients with NAFLD, and that the erated through the proteolytic processing of proglucagon that
presence and severity of NAFLD is associated with an increased stimulates insulin secretion and inhibits secretion of glucagon.
incidence of CKD, and that this association could be independent GLP‐1 is also an insulin sensitizer with additional metabolic
of multiple cardiorenal risk factors. Taken together, more active effects that contribute to its anti‐NASH activity [112]. Liraglutide,
and systematic research for the link between the microbiome and a GLP‐1 agonist requiring daily injection, improved NASH his-
liver and kidney disease are needed to understand the mechanisms tology in a small pilot study [113]. Interestingly, L cells express
and causal association between these diseases. FXR and FXR also regulates GLP‐1 synthesis [114].

Bile acid signaling


REFERENCES
Bile acids are produced in the liver from cholesterol and are
metabolized by enzymes derived from the gut microbiome. 1. Targher, G. and Byrne, C.D. Non‐alcoholic fatty liver disease: an emerging
Evidence shows that bile acids are critically important for main- driving force in chronic kidney disease. Nat Rev Nephrol, 2017;13(5):
taining a healthy gut microbiota [98]. The synthesis of bile acids 297–310.
is tightly regulated by negative feedback inhibition through the 2. Yeung, M.W., Wong, G.L., Choi, K.C. et al. Advanced liver fibrosis but not
steatosis is independently associated with albuminuria in Chinese patients
nuclear receptor FXR [99], which is a transcription factor that with type 2 diabetes. J Hepatol, 2017.
binds to the promoter region and initiates the expression of a 3. Sinn, D.H., Kang, D., Jang, H.R. et al. Development of chronic kidney dis-
wide range of target genes [100]. FXR is expressed in several ease in patients with non‐alcoholic fatty liver disease: a cohort study. J
tissues including the liver, intestine, and kidney [101]. In the Hepatol, 2017;67(6):1274–80.
liver, bile acid‐activated FXR induces the expression of small 4. Mantovani, A., Zaza, G., Byrne, C.D. et al. Nonalcoholic fatty liver disease
increases risk of incident chronic kidney disease: a systematic review and
heterodimer partner (SHP), which binds to liver receptor meta‐analysis. Metabolism, 2018;79:64–76.
homolog‐1 (LRH‐1) and thereby inhibits expression of the 5. Sole, C., Pose, E., Sola, E., and Gines, P. Hepatorenal syndrome in the era of
Cyp7a1 gene [102]. The primary bile acids produced in humans acute kidney injury. Liver Int, 2018;38(11):1891–1901.
are chenodeoxycholic acid (CDCA) and cholic acid (CA), while 6. Angeli, P., Gines, P., Wong, F. et  al. Diagnosis and management of acute
kidney injury in patients with cirrhosis: revised consensus recommendations
rodents produce CA and muricholic acids (MCAs), predomi-
of the International Club of Ascites. J Hepatol, 2015;62(4):968–74.
nantly beta‐MCA (βMCA) [103]. 7. Durand, F., Graupera, I., Gines, P., Olson, J.C., and Nadim, M.K.
The most potent endogenous ligand for FXR is CDCA, fol- Pathogenesis of hepatorenal syndrome: implications for therapy. Am J
lowed by CA, DCA, and LCA [104]. UDCA, TαMCA, TβMCA, Kidney Dis, 2016;67(2):318–28.
and GβMCA inhibit FXR activation [98]. 8. Sanyal, A.J., Boyer, T.D., Frederick, R.T. et al. Reversal of hepatorenal syn-
drome type 1 with terlipressin plus albumin vs. placebo plus albumin in a
A positive phase 2b clinical trial of obeticholic acid (OCA)
pooled analysis of the OT‐0401 and REVERSE randomised clinical studies.
(25 mg per day) was terminated early because of demonstration Aliment Pharmacol Ther, 2017;45(11):1390–402.
of efficacy during an interim analysis planned a priori [105]. 9. Glass, L. and Sharma, P. Evidence‐based therapeutic options for hepatorenal
However, OCA, in 25 mg per day doses, causes both pruritus syndrome. Gastroenterology, 2016;150(4):1031–3.
and moderate increases in low‐density lipoprotein (LDL) cho- 10. Wong, F., Leung, W., Al Beshir, M., Marquez, M., and Renner, E.L.
Outcomes of patients with cirrhosis and hepatorenal syndrome type 1 treated
lesterol in some individuals. Several small‐molecule agonists of with liver transplantation. Liver Transpl, 2015;21(3):300–7.
FXR that do not have a bile acid structural backbone have been 11. Cavallin, M., Kamath, P.S., Merli, M. et al. Terlipressin plus albumin versus
developed on the premise that they will not increase LDL cho- midodrine and octreotide plus albumin in the treatment of hepatorenal syn-
lesterol or cause pruritus, but this has not yet been proven. drome: a randomized trial. Hepatology, 2015;62(2):567–74.
Additionally, FXR induces the release of the growth factor 12. Cohen, S.D., Kopp, J.B., and Kimmel, P.L. Kidney diseases associated with
human immunodeficiency virus infection. N Engl J Med, 2017;377(24):
FGF19 from the intestine upon bile acid binding to the recep- 2363–74.
tor [106], which has beneficial effects on NASH in animal 13. Rutledge, S.M., Chung, R.T., and Sise, M.E. Treatment of hepatitis C virus
models, although some results are conflicting [107, 108]. In infection in patients with mixed cryoglobulinemic syndrome and cryoglobu-
a phase 2a study, FGF19 analog yielded a dramatic reduction linemic glomerulonephritis. Hemodial Int, 2018;22(1):S81–96.
14. Collins, J.M., Raphael, K.L., Terry, C. et al. Hepatitis B virus reactivation
in hepatic fat and liver enzymes in patients with biopsy‐
during successful treatment of hepatitis C virus with sofosbuvir and sime-
confirmed NASH [107]. previr. Clin Infect Dis, 2015;61(8):1304–6.
TGR5 is another bile acid‐responsive receptor involved in 15. Sise, M.E. Hepatitis C virus infection and the kidney. Nephrol Dial
host metabolism. TGR5 is a plasma membrane‐bound G Transplant, 2018;34(3):415–8.
49:  The Kidney in Liver Disease 643

16. Rodrigues, E.M.F., Fernandes, R., Susin, R., and Fior, B. Immune reconstitu- 41. Ponugoti, B., Kim, D.H., Xiao, Z. et  al. SIRT1 deacetylates and inhibits
tion inflammatory syndrome as a cause of autoimmune hepatitis and acute SREBP‐1C activity in regulation of hepatic lipid metabolism. J Biol Chem,
liver failure. Rev Bras Ter Intensiva, 2017;29(3):382–5. 2010;285(44):33959–70.
17. Gowda, C., Newcomb, C.W., Liu, Q. et al. Risk of acute liver injury with 42. Porstmann, T., Santos, C.R., Griffiths, B. et al. SREBP activity is regulated
antiretroviral therapy by viral hepatitis status. Open Forum Infect Dis, by mTORC1 and contributes to Akt‐dependent cell growth. Cell Metab,
2017;4(2):ofx012. 2008;8(3):224–36.
18. Nadkarni, G.N., Konstantinidis, I., and Wyatt, C.M. HIV and the aging kid- 43. Begriche, K., Massart, J., Robin, M.A., Bonnet, F., and Fromenty, B.
ney. Curr Opin HIV AIDS, 2014;9(4):340–5. Mitochondrial adaptations and dysfunctions in nonalcoholic fatty liver dis-
19. Genovese, G., Friedman, D.J., Ross, M.D. et al. Association of trypanolytic ease. Hepatology, 2013;58(4):1497–507.
ApoL1 variants with kidney disease in African Americans. Science, 44. Fabbrini, E., Mohammed, B.S., Magkos, F., Korenblat, K.M., Patterson,
2010;329(5993):841–5. B.W., and Klein, S. Alterations in adipose tissue and hepatic lipid kinetics in
20. Tzur, S., Rosset, S., Shemer, R. et al. Missense mutations in the APOL1 gene obese men and women with nonalcoholic fatty liver disease. Gastroenterology,
are highly associated with end stage kidney disease risk previously attributed 2008;134(2):424–31.
to the MYH9 gene. Hum Genet, 2010;128(3):345–50. 45. Quiroga, A.D. and Lehner, R. Pharmacological intervention of liver triacylg-
21. Kopp, J.B., Nelson, G.W., Sampath, K. et  al. APOL1 genetic variants in lycerol lipolysis: the good, the bad and the ugly. Biochem Pharmacol,
focal segmental glomerulosclerosis and HIV‐associated nephropathy. J Am 2018;155:233–41.
Soc Nephrol, 2011;22(11):2129–37. 46. Ioannou, G.N. The role of cholesterol in the pathogenesis of NASH. Trends
22. Fishman, J.A. Infection in organ transplantation. Am J Transplant, Endocrinol Metab, 2016;27(2):84–95.
2017;17(4):856–79. 47. Fiorucci, S., Biagioli, M., and Distrutti, E. Future trends in the treatment of
23. Hales, C.M., Fryar, C.D., Carroll, M.D., Freedman, D.S., and Ogden, C.L. non‐alcoholic steatohepatitis. Pharmacol Res, 2018;134:289–98.
Trends in obesity and severe obesity prevalence in US youth and adults by 48. Sun, L., Halaihel, N., Zhang, W., Rogers, T., and Levi, M. Role of sterol
sex and age, 2007–2008 to 2015–2016. JAMA, 2018;319(16):1723–5. regulatory element‐binding protein 1 in regulation of renal lipid metabolism
24. D’Agati, V.D., Chagnac, A., de Vries, A.P. et al. Obesity‐related glomeru- and glomerulosclerosis in diabetes mellitus. J Biol Chem, 2002;277(21):
lopathy: clinical and pathologic characteristics and pathogenesis. Nat Rev 18919–27.
Nephrol, 2016;12(8):453–71. 49. Jiang, T., Wang, Z., Proctor, G. et al. Diet‐induced obesity in C57BL/6J mice
25. Lakkis, J.I. and Weir, M.R. Obesity and kidney disease. Prog Cardiovasc causes increased renal lipid accumulation and glomerulosclerosis via a sterol
Dis, 2018;61(2):157–67. regulatory element‐binding protein‐1c‐dependent pathway. J Biol Chem,
26. Whaley‐Connell, A. and Sowers, J.R. Obesity and kidney disease: from 2005;280(37):32317–25.
population to basic science and the search for new therapeutic targets. 50. Kume, S., Uzu, T., Araki, S. et al. Role of altered renal lipid metabolism in
Kidney Int, 2017;92(2):313–23. the development of renal injury induced by a high‐fat diet. J Am Soc Nephrol,
27. Wang, X.X., Wang, D., Luo, Y. et al. FXR/TGR5 dual agonist prevents pro- 2007;18(10):2715–23.
gression of nephropathy in diabetes and obesity. J Am Soc Nephrol, 51. Wu, Y., Liu, Z., Xiang, Z. et  al. Obesity‐related glomerulopathy: insights
2018;29(1):118–37. from gene expression profiles of the glomeruli derived from renal biopsy
28. Wang, X.X., Jiang, T., Shen, Y., Adorini, L. et al. The farnesoid X receptor samples. Endocrinology, 2006;147(1):44–50.
modulates renal lipid metabolism and diet‐induced renal inflammation, fibro- 52. Jiang, T., Wang, X.X., Scherzer, P. et  al. Farnesoid X receptor modulates
sis, and proteinuria. Am J Physiol Renal Physiol, 2009;297(6):F1587–96. renal lipid metabolism, fibrosis, and diabetic nephropathy. Diabetes,
29. Wang, X.X., Edelstein, M.H., Gafter, U. et al. G protein‐coupled bile acid 2007;56(10):2485–93.
receptor TGR5 activation inhibits kidney disease in obesity and diabetes. J 53. Wang, X.X., Jiang, T., Shen, Y. et al. Diabetic nephropathy is accelerated by
Am Soc Nephrol, 2016;27(5):1362–78. farnesoid X receptor deficiency and inhibited by farnesoid X receptor activa-
30. Gai, Z., Gui, T., Hiller, C., and Kullak‐Ublick, G.A. Farnesoid X receptor tion in a type 1 diabetes model. Diabetes, 2010;59(11):2916–27.
protects against kidney injury in uninephrectomized obese mice. J Biol 54. Yang, X., Okamura, D.M., Lu, X. et  al. CD36 in chronic kidney disease:
Chem, 2016;291(5):2397–411. novel insights and therapeutic opportunities. Nat Rev Nephrol, 2017;13(12):
31. Friedman, S.L., Neuschwander‐Tetri, B.A., Rinella, M., and Sanyal, A.J. 769–81.
Mechanisms of NAFLD development and therapeutic strategies. Nat Med, 55. Herman‐Edelstein, M., Scherzer, P., Tobar, A., Levi, M., and Gafter, U.
2018;24(7):908–22. Altered renal lipid metabolism and renal lipid accumulation in human dia-
32. Baiceanu, A., Mesdom, P., Lagouge, M., and Foufelle, F. Endoplasmic retic- betic nephropathy. J Lipid Res, 2014;55(3):561–72.
ulum proteostasis in hepatic steatosis. Nat Rev Endocrinol, 2016;12(12): 56. Wang, X.X., Jiang, T., Shen, Y. et al. Vitamin D receptor agonist doxercalcif-
710–22. erol modulates dietary fat‐induced renal disease and renal lipid metabolism.
33. Artunc, F., Schleicher, E., Weigert, C., Fritsche, A., Stefan, N., and Haring, Am J Physiol Renal Physiol, 2011;300(3):F801–10.
H.U. The impact of insulin resistance on the kidney and vasculature. Nat Rev 57. Proctor, G., Jiang, T., Iwahashi, M., Wang, Z., Li, J., and Levi, M. Regulation
Nephrol, 2016;12(12):721–37. of renal fatty acid and cholesterol metabolism, inflammation, and fibrosis in
34. Samuel, V.T. and Shulman, G.I. Nonalcoholic fatty liver disease as a nexus Akita and OVE26 mice with type 1 diabetes. Diabetes, 2006;55(9):2502–9.
of metabolic and hepatic diseases. Cell Metab, 2018;27(1):22–41. 58. Marra, F. and Svegliati‐Baroni, G. Lipotoxicity and the gut‐liver axis in
35. Gnudi, L., Coward, R.J.M., and Long, D.A. Diabetic nephropathy: perspec- NASH pathogenesis. J Hepatol, 2018;68(2):280–95.
tive on novel molecular mechanisms. Trends Endocrinol Metab, 59. Caldwell, S.H., Swerdlow, R.H., Khan, E.M. et al. Mitochondrial abnormali-
2016;27(11):820–30. ties in non‐alcoholic steatohepatitis. J Hepatol, 1999;31(3):430–4.
36. Lay, A.C., Hurcombe, J.A., Betin, V.M.S. et  al. Prolonged exposure of 60. Ding, R.B., Bao, J., and Deng, C.X. Emerging roles of SIRT1 in fatty liver
mouse and human podocytes to insulin induces insulin resistance through diseases. Int J Biol Sci, 2017;13(7):852–67.
lysosomal and proteasomal degradation of the insulin receptor. Diabetologia, 61. Morigi, M., Perico, L., and Benigni, A. Sirtuins in renal health and disease. J
2017;60(11):2299–311. Am Soc Nephrol, 2018;29(7):1799–809.
37. Wang, X.X., Levi, J., Luo, Y. et al. SGLT2 protein expression is increased in 62. Tang, C., Cai, J., and Dong, Z. Mitochondrial dysfunction in obesity‐related
human diabetic nephropathy: SGLT2 protein inhibition decreases renal lipid kidney disease: a novel therapeutic target. Kidney Int, 2016;90(5):930–3.
accumulation, inflammation, and the development of nephropathy in dia- 63. Grattagliano, I., de Bari, O., Bernardo, T.C., Oliveira, P.J., Wang, D.Q., and
betic mice. J Biol Chem, 2017;292(13):5335–48. Portincasa, P. Role of mitochondria in nonalcoholic fatty liver disease‐‐from
38. Wang, D., Luo, Y., Wang, X. et al. The sodium‐glucose cotransporter 2 inhib- origin to propagation. Clin Biochem, 2012;45(9):610–8.
itor dapagliflozin prevents renal and liver disease in western diet induced 64. Caballero, F., Fernandez, A., Matias, N. et  al. Specific contribution of
obesity mice. Int J Mol Sci, 2018;19(1). methionine and choline in nutritional nonalcoholic steatohepatitis: impact on
39. Koo, S.H. Nonalcoholic fatty liver disease: molecular mechanisms for the mitochondrial S‐adenosyl‐L‐methionine and glutathione. J Biol Chem,
hepatic steatosis. Clin Mol Hepatol, 2013;19(3):210–5. 2010;285(24):18528–36.
40. Benhamed, F., Denechaud, P.D., Lemoine, M. et al. The lipogenic transcrip- 65. Al‐Serri, A., Anstee, Q.M., Valenti, L. et al. The SOD2 C47T polymorphism
tion factor ChREBP dissociates hepatic steatosis from insulin resistance in influences NAFLD fibrosis severity: evidence from case‐control and intra‐
mice and humans. J Clin Invest, 2012;122(6):2176–94. familial allele association studies. J Hepatol, 2012;56(2):448–54.
644 THE LIVER:  REFERENCES

66. Ajith, T.A. Role of mitochondria and mitochondria‐targeted agents in non‐   91. Utzschneider, K.M., Kratz, M., Damman, C.J., and Hullar, M. Mechanisms
alcoholic fatty liver disease. Clin Exp Pharmacol Physiol, 2018;45(5): linking the gut microbiome and glucose metabolism. J Clin Endocrinol
413–21. Metab, 2016;101(4):1445–54.
67. Lebeaupin, C., Vallee, D., Hazari, Y., Hetz, C., Chevet, E., and Bailly‐Maitre,   92. Koopen, A.M., Groen, A.K., and Nieuwdorp, M. Human microbiome as
B. Endoplasmic reticulum stress signaling and the pathogenesis of non‐alco- therapeutic intervention target to reduce cardiovascular disease risk. Curr
holic fatty liver disease. J Hepatol, 2018;69(4):927–47. Opin Lipidol, 2016;27(6):615–22.
68. Kropski, J.A. and Blackwell, T.S. Endoplasmic reticulum stress in the patho-   93. Briskey, D., Tucker, P.S., Johnson, D.W., and Coombes, J.S. Microbiota
genesis of fibrotic disease. J Clin Invest, 2018;128(1):64–73. and the nitrogen cycle: Implications in the development and progression of
69. Cybulsky, A.V. Endoplasmic reticulum stress, the unfolded protein response CVD and CKD. Nitric Oxide, 2016;57:64–70.
and autophagy in kidney diseases. Nat Rev Nephrol, 2017;13(11):681–96.   94. Sampaio‐Maia, B., Simoes‐Silva, L., Pestana, M., Araujo, R., and Soares‐
70. Brunt, E.M., Wong, V.W., Nobili, V. et al. Nonalcoholic fatty liver disease. Silva, I.J. The role of the gut microbiome on chronic kidney disease. Adv
Nat Rev Dis Primers, 2015;1:15080. Appl Microbiol, 2016;96:65–94.
71. Schuster, S., Cabrera, D., Arrese, M., and Feldstein, A.E. Triggering and   95. Scorletti, E. and Byrne, C.D. Extrahepatic diseases and NAFLD: the trian-
resolution of inflammation in NASH. Nat Rev Gastroenterol Hepatol, gular relationship between NAFLD, type 2‐diabetes and dysbiosis. Dig Dis,
2018;15(6):349–64. 2016;34(1):11–8.
72. Wree, A., Broderick, L., Canbay, A., Hoffman, H.M., and Feldstein, A.E.   96. Tang, W.H., Wang, Z., Kennedy, D.J. et al. Gut microbiota‐dependent tri-
From NAFLD to NASH to cirrhosis‐new insights into disease mechanisms. methylamine N‐oxide (TMAO) pathway contributes to both development
Nat Rev Gastroenterol Hepatol, 2013;10(11):627–36. of renal insufficiency and mortality risk in chronic kidney disease. Circ
73. Han, M.S., Park, S.Y., Shinzawa, K. et  al. Lysophosphatidylcholine as a Res, 2015;116(3):448–55.
death effector in the lipoapoptosis of hepatocytes. J Lipid Res, 2008;   97. Nallu, A., Sharma, S., Ramezani, A., Muralidharan, J., and Raj, D. Gut
49(1):84–97. microbiome in chronic kidney disease: challenges and opportunities. Transl
74. Csak, T., Ganz, M., Pespisa, J., Kodys, K., Dolganiuc, A., and Szabo, G. Fatty Res, 2017;179:24–37.
acid and endotoxin activate inflammasomes in mouse hepatocytes that release   98. Wahlström, A., Sayin Sama I., Marschall, H‐U., and Bäckhed, F. Intestinal
danger signals to stimulate immune cells. Hepatology, 2011;54(1):133–44. crosstalk between bile acids and microbiota and its impact on host metabo-
75. Wree, A., McGeough, M.D., Inzaugarat, M.E. et al. NLRP3 inflammasome lism. Cell Metab, 2016;24(1):41–50.
driven liver injury and fibrosis: roles of IL‐17 and TNF in mice. Hepatology,   99. Sinal, C.J., Tohkin, M., Miyata, M., Ward, J.M., Lambert, G., and Gonzalez,
2017. F.J. Targeted disruption of the nuclear receptor FXR/BAR impairs bile acid
76. Sharma, M., Mitnala, S., Vishnubhotla, R.K., Mukherjee, R., Reddy, D.N., and lipid homeostasis. Cell, 2000;102(6):731–44.
and Rao, P.N. The riddle of nonalcoholic fatty liver disease: progression 100. Teodoro, J.S., Rolo, A.P., and Palmeira, C.M. Hepatic FXR: key regulator
from nonalcoholic fatty liver to nonalcoholic steatohepatitis. J Clin Exp of whole‐body energy metabolism. Trends Endocrinol Metab, 2011;
Hepatol, 2015;5(2):147–58. 22(11):458–66.
77. Targher, G., Day, C.P., and Bonora, E. Risk of cardiovascular disease in 101. Lefebvre, P., Cariou, B., Lien, F., Kuipers, F., and Staels, B. Role of bile
patients with nonalcoholic fatty liver disease. N Engl J Med, 2010; acids and bile acid receptors in metabolic regulation. Physiol Rev,
363(14):1341–50. 2009;89(1):147–91.
78. Pelletier, C.C., Koppe, L., Croze, M.L. et al. White adipose tissue overpro- 102. Goodwin, B., Jones, S.A., Price, R.R. et  al. A regulatory cascade of the
duces the lipid‐mobilizing factor zinc alpha2‐glycoprotein in chronic kidney nuclear receptors FXR, SHP‐1, and LRH‐1 represses bile acid biosynthesis.
disease. Kidney Int, 2013;83(5):878–86. Mol Cell. 2000;6(3):517–26.
79. Axelsson, J., Åström, G., Sjölin, E. et al. Uraemic sera stimulate lipolysis in 103. Takahashi, S., Fukami, T., Masuo, Y. et al. Cyp2c70 is responsible for the
human adipocytes: role of perilipin. Nephrol Dial Transplant, 2011; species difference in bile acid metabolism between mice and humans. J
26(8):2485–91. Lipid Res, 2016;57(12):2130–7.
80. Leung, C., Rivera, L., Furness, J.B., and Angus, P.W. The role of the gut 104. Makishima, M., Okamoto, A.Y., Repa, J.J. et al. Identification of a nuclear
microbiota in NAFLD. Nat Rev Gastroenterol Hepatol, 2016;13(7): receptor for bile acids. Science, 1999;284(5418):1362–5.
412–25. 105. Neuschwander‐Tetri, B.A., Loomba, R., Sanyal, A.J. et  al. Farnesoid X
81. Betrapally, N.S., Gillevet, P.M., and Bajaj, J.S. Changes in the intestinal nuclear receptor ligand obeticholic acid for non‐cirrhotic, non‐alcoholic
microbiome and alcoholic and nonalcoholic liver diseases: causes or effects? steatohepatitis (FLINT): a multicentre, randomised, placebo‐controlled
Gastroenterology, 2016;150(8):1745–55.e3. trial. Lancet, 2015;385(9972):956–65.
82. Loomba, R., Seguritan, V., Li, W. et al. Gut microbiome based metagenomic 106. Nies, V.J., Sancar, G., Liu, W. et al. Fibroblast growth factor signaling in
signature for non‐invasive detection of advanced fibrosis in human nonalco- metabolic regulation. Front Endocrinol,2015;6:193.
holic fatty liver disease. Cell Metab, 2017;25(5):1054–62.e5. 107. Harrison, S.A., Rinella, M.E., Abdelmalek, M.F. et al. NGM282 for treat-
83. Chen, Y.M., Liu, Y., Zhou, R.F. et  al. Associations of gut‐flora‐dependent ment of non‐alcoholic steatohepatitis: a multicentre, randomised, double‐
metabolite trimethylamine‐N‐oxide, betaine and choline with non‐alcoholic blind, placebo‐controlled, phase 2 trial. Lancet, 2018;391(10126):1174–85.
fatty liver disease in adults. Sci Rep, 2016;6:19076. 108. Jiang, C., Xie, C., Li, F. et al. Intestinal farnesoid X receptor signaling pro-
84. Arendt, B.M., Ma, D.W., Simons, B. et al. Nonalcoholic fatty liver disease is motes nonalcoholic fatty liver disease. J Clin Invest, 2015;125(1):386–402.
associated with lower hepatic and erythrocyte ratios of phosphatidylcholine 109. Kawamata, Y., Fujii, R., Hosoya, M. et  al. A G protein‐coupled receptor
to phosphatidylethanolamine. Appl Physiol Nutr Metab, 2013;38(3): responsive to bile acids. J Biol Chem, 2003;278(11):9435–40.
334–40. 110. Chen, X., Lou, G., Meng, Z., and Huang, W. TGR5: a novel target for
85. Cohen, L.J., Esterhazy, D., Kim, S.H. et  al. Commensal bacteria make weight maintenance and glucose metabolism. Exp Diabetes Res, 2011;
GPCR ligands that mimic human signalling molecules. Nature, 2011:853501.
2017;549(7670):48–53. 111. Thomas, C., Gioiello, A., Noriega, L. et al. TGR5‐mediated bile acid sens-
86. Tripathi, A., Debelius, J., Brenner, D.A. et  al. The gut‐liver axis and the ing controls glucose homeostasis. Cell Metab, 2009;10(3):167–77.
intersection with the microbiome. Nat Rev Gastroenterol Hepatol, 112. Jinnouchi, H., Sugiyama, S., Yoshida, A. et al. Liraglutide, a glucagon‐like
2018;15(7):397–411. peptide‐1 analog, increased insulin sensitivity assessed by hyperinsuline-
87. Liu, R., Hong, J., Xu, X. et  al. Gut microbiome and serum metabolome mic‐euglycemic clamp examination in patients with uncontrolled type 2
alterations in obesity and after weight‐loss intervention. Nat Med, 2017; diabetes mellitus. J Diabetes Res, 2015;2015:8.
23:859. 113. Armstrong, M.J., Gaunt, P., Aithal, G.P. et al. Liraglutide safety and effi-
88. Mehal, W.Z. The Gordian knot of dysbiosis, obesity and NAFLD. Nat Rev cacy in patients with non‐alcoholic steatohepatitis (LEAN): a multicentre,
Gastroenterol Hepatol, 2013;10(11):637–44. double‐blind, randomised, placebo‐controlled phase 2 study. Lancet,
89. Tilg, H., Cani, P.D., and Mayer, E.A. Gut microbiome and liver diseases. 2016;387(10019):679–90.
Gut, 2016;65(12):2035–44. 114. Trabelsi, M‐S., Daoudi, M., Prawitt, J. et al. Farnesoid X receptor inhibits
90. Qin, J., Li, Y., Cai, Z. et  al. A metagenome‐wide association study of gut glucagon‐like peptide‐1 production by enteroendocrine L cells. Nat
microbiota in type 2 diabetes. Nature, 2012;490(7418):55–60. Commun, 2015;6:7629.
α1‐Antitrypsin Deficiency
50 David A. Rudnick and David H. Perlmutter
Department of Pediatrics, Division of Gastroenterology, Hepatology, and Nutrition and Department of
Developmental Biology, Washington University School of Medicine in St. Louis, St. Louis Children’s
Hospital, St. Louis, MO, USA

INTRODUCTION of cigarette smoking accelerates lung injury [5]. Moreover, the


elastase–antielastase theory for the pathogenesis of emphysema
The classical form of α1‐antitrypsin (α1‐AT) deficiency, is based on the concept that oxidative inactivation of α1‐AT as a
homozygous for the mutant α1‐ATZ allele, is a relatively com- result of cigarette smoking plays a key role in the emphysema of
mon disease. It affects approximately 1 in 1600 to 1 in 2000 live α1‐AT‐sufficient individuals, the vast majority of cases of
births in most populations of Northern European ancestry [1, 2]. emphysema [5]. It is more difficult to explain the pathogenesis
Although only a subgroup of deficient individuals develop liver of liver injury in this deficiency. Results of transgenic animal
disease, it represents the most common metabolic cause of liver experiments provide further evidence that the liver disease does
disease in children [3] and can be associated with chronic liver not result from a deficiency in antielastase activity [6, 7]. Most
disease and hepatocellular carcinoma in adults [4]. This defi- data in the literature corroborate the concept that liver injury in
ciency also causes premature development of pulmonary α1‐AT deficiency results from the hepatotoxicity of mutant­
emphysema in adults. α1‐ ATZ accumulation in hepatocytes.
The α1‐AT molecule is a single‐chain secretory glycoprotein Although it is a single gene defect, there is extraordinary
that inhibits destructive neutrophil proteases including elastase, variation in the phenotypic expression of disease in the classical
cathepsin G, and proteinase 3. It is often referred to as a hepatic form of α1‐AT deficiency. For instance, nationwide prospective
acute phase reactant in that plasma α1‐AT is predominantly screening studies done by Sveger in Sweden showed that only
derived from the liver and plasma levels increase three‐ to five- 8% of an unbiased cohort developed clinically significant liver
fold during the host response to tissue injury/inflammation. It is disease through the fourth decade of life [1, 8]. These data indi-
the archetype of a family of structurally related circulating ser- cate that other genetic traits and/or environmental factors pre-
ine protease inhibitors termed SERPINs. In the deficient state, dispose a subgroup of PIZZ individuals, that is, those
there is an approximately 85% to 90% reduction in serum α1‐ homozygous for the α1‐ATZ mutant allele, to liver injury. There
AT concentrations. A single amino acid substitution results in a is also variation in incidence and severity of lung injury among
mutant protein unable to traverse the secretory pathway. This α1‐AT‐deficient individuals. Environmental factors, such as
α1‐ATZ protein is retained in the endoplasmic reticulum (ER) cigarette smoking, obviously play an important role [2, 5].
rather than secreted into the blood and body fluids. However, there are well‐documented examples of siblings and
The classical deficient state is unique as a genetic disease in other relatives of deficient individuals with severe emphysema
that it causes injury to one target organ, lung, by loss‐of‐func- who have the same genotype, a history of heavy cigarette smok-
tion and injury to another target organ, liver, by what appears to ing and only mild, subclinical pulmonary function abnormali-
be a gain‐of‐function mechanism. Most data in the literature ties even at advanced ages [9].
indicate that emphysema results from a decreased number of Historically, the diagnosis of α1‐AT deficiency was based on
α1‐AT molecules within the lower respiratory tract, allowing the altered migration of the abnormal α1‐ATZ molecule in
unregulated elastolytic attack on the lung’s connective tissue serum specimens subjected to isoelectric focusing gel analysis.
matrix [5]. Oxidative inactivation of residual α1‐AT as a result Contemporary diagnostic strategies may now employ genomic

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
646 THE LIVER:  CLINICAL MANIFESTATIONS OF LIVER DISEASE

analyses for specific gene mutations. Treatment of α1‐AT defi- autophagy function in middle age may play an important role in
ciency‐associated liver disease remains mostly supportive. Liver age‐dependent degenerative diseases in general.
replacement therapy has been used successfully for severe liver Earlier studies also reported incidentally identified signifi-
injury. α1‐AT‐deficient patients with severe emphysema have cant liver pathology at autopsy in some elderly patients with
undergone lung transplantation. Intravenous administration of α1‐AT deficiency without known liver disease. A Swedish study
purified plasma α1‐AT augmentation therapy was approved for showed a higher risk of cirrhosis and primary liver cancer in
α1‐AT deficiency lung disease on the basis of biochemical effi- such subjects than previously suspected [4]. Based on these con-
cacy studies [10]. Recently, a randomized, placebo‐controlled siderations, α1‐AT deficiency should be considered in the dif-
trial provided some initial, albeit limited, evidence that augmen- ferential diagnosis of any adult who presents with chronic
tation therapy slows the progression of lung destruction in α1‐ hepatitis, cirrhosis, portal hypertension, or hepatocellular carci-
AT deficient subjects [11]. noma of unknown origin. Taken together, the overwhelming
clinical experience with this disease indicates wide variation in
liver disease phenotype with many “protected” from or having
slowly progressing liver disease.
CLINICAL MANIFESTATIONS OF It still remains unclear whether liver injury results from the
LIVER DISEASE heterozygous α1‐AT PIMZ (PI genotype heterozygous for the
wild‐type M and mutant Z alleles) state by itself. Liver biopsy
Liver involvement is often noticed in early infancy because of and transplant database studies have identified heterozygous
persistent jaundice, elevated serum transaminases, and increased patients with severe liver disease without other apparent expla-
conjugated bilirubin levels [12, 13]. Some infants are recog- nation (reviewed in [3]). However, these studies are generally
nized because of cholestasis characterized by pruritus and unable to exclude environmental causes of liver disease. Indeed,
hypercholesterolemia. The clinical picture resembles extrahe- one cross‐sectional study of re‐examined α1‐AT deficiency
patic biliary atresia but liver histology shows paucity of intrahe- patients in a referral‐based Austrian university hospital sug-
patic bile ducts [14]. Alternatively, liver disease may first be gested that liver disease in heterozygotes was largely accounted
discovered in later childhood or early adolescence when the for by hepatitis B or C virus, autoimmune disease, or alcohol
affected individual presents with hepatosplenomegaly, ascites, abuse (reviewed in [3]). Other studies have subsequently
or esophageal variceal hemorrhage. In some but not all cases, a reported significant correlations between the MZ state and
history of unexplained prolonged obstructive jaundice during development of, presentation with, or more rapid progression to
the neonatal period is identified. advanced liver disease and liver transplantation in subjects with
The incidence and natural history of liver disease in α1‐AT other chronic liver diseases [16, 17].
deficiency was determined in a nationwide screening study of Liver disease has also been associated with other allelic vari-
newborns in Sweden by Sveger [1]. From 200 000 infants, 127 ants of α1‐AT. For example, children with compound PISZ het-
PIZZ newborns were identified and prospectively followed, erozygosity are affected by liver injury in a manner similar to
now through age 37–40 years [8]. Initial analysis of this cohort PIZZ children [1], and liver disease was reported in subjects
identified 14 PIZZ infants with prolonged obstructive jaundice, with α1‐AT deficiency type PIMmalton (PI genotype for the
another 8 with other reasons for clinical suspicion of liver dis- Mmalton allele, Table 50.1) [18]. These observations are inter-
ease, and the others with no clinical evidence for liver disease esting because the α1‐ATS molecule forms heteropolymers with
[1]. Five PIZZ subjects subsequently died in early childhood, α1‐ATZ [19], and, like α1‐ATZ, the PIMmalton molecule
two with known cirrhosis and two others with incidentally‐ undergoes polymerization and ER retention [20]. Liver disease
detected liver disease at autopsy [15]. At reassessment of this has also been detected in single patients with other α1‐AT allelic
cohort at age 37–40 years, no participating subjects showed evi- variants [3], but in those cases other causes of liver injury have
dence of active liver disease [8]. However, abnormalities in generally not been completely excluded.
liver‐related laboratory and/or imaging studies were reported. Historically, the diagnosis was established by a serum α1‐AT
One issue not addressed by the Sveger study is whether any phenotype determination by isoelectric focusing or by agarose
PIZZ subjects have subclinical histologic abnormalities that electrophoresis at acid pH. Modern genomic methodologies
lead to progressive disease in later adulthood. permit specific analyses for known liver disease‐associated α1‐
It is now evident that this deficiency causes liver disease with AT gene mutations, and analyses for gene duplications, dele-
initial onset in adults. In our retrospective analysis of US liver tions, or previously unidentified mutations in the AT gene
transplantation databases, we found that 77% of α1‐AT‐defi- coding region. Liver histology is characterized by periodic
cient subjects undergoing liver transplantation over the last 20 acid–Schiff‐positive, diastase‐resistant globules in hepatocyte
years were adults, with peak age of transplantation in those sub- ER. These globules are most prominent in periportal hepato-
jects of 50–64 years [16]. Many are probably heterozygotes cytes, and may also be seen in Kupffer cells and cells of biliary
with other potential causes of liver disease, and some may be ductular lineage (reviewed in [3]). There may be variable hepa-
poorly substantiated diagnoses. Nevertheless, our analysis sug- tocellular necrosis, inflammatory cell infiltration, periportal
gests that severe liver disease requiring transplantation in sub- fibrosis, and/or cirrhosis. There is often evidence of bile duct
jects with ZZ or SZ phenotypes is more frequent in adults than epithelial cell destruction, and, occasionally, paucity of intrahe-
children. This observation is consistent with the discovery that patic bile ducts. There can also be an intense autophagic reac-
autophagy plays a critical role in the pathobiology of this liver tion with nascent and degradative‐type autophagic vacuoles
disease and the understanding that a physiological decline in detected by electron microscopy on liver biopsies [21].
α1-ANTITRYPSIN DEFICIENCY
50:  647

Table 50.1  Deficiency variants of α1‐antitrypsin


Clinical disease
Variant Defect Site Liver Lung Cellular defect
Z Single base substitution M1 [Ala213] Glu342‐Lys + + IC accumulation
S Single base substitution Glu264‐Val − − IC accumulation
MHeerlen Single base substitution Pro369‐Leu − + IC accumulation
MProcida Single base substitution Leu41‐Pro − + IC accumulation
MMalton Single base deletion Phe52 + + IC accumulation
MDuarte Unknown Unknown +? + Unknown
MMineral Springs Single base substitution Gly57‐Glu − + No function; EC degradation?
Siiyama Single base substitution Ser53‐Phe − + IC accumulation
PDuarte Two base substitution Arg101‐His +? + Unknown
Asp256‐Val
PLowell Single base substitution Asp256‐Val − + IC degradation?
WBethesda Single base substitution Ala336‐Thre − + Accelerated catabolism?
ZWrexham Single base substitution Ser19‐Leu ? ? Unknown
F Single base substitution Arg223‐Cys − − Unknown
T Single base substitution Glu264‐Val − − Unknown
I Single base substitution Arg39‐Cys − − IC degradation
MPalermo Single base deletion Phe51 − − Unknown
MNichinan Single base deletion and Phe52 − − Unknown
single base substitution Gly148‐Arg
ZAusburg Single base substitution Glu342‐Lys − − Unknown
King’s Single base substitution His334‐Asp + IC accumulation
Mpisa Single base substitution Lys259‐Ile − +? IC accumulation
Etaurisano Single base substitution Lys368‐Glu − + IC accumulation
Yorzinuovi Single base substitution Pro391‐His +? − IC accumulation
PiSDonosti S variant defect + single base S variant site + Ser14‐Phe − +? IC polymerization, degradation; reduced secretion
substitution
PiTijarafe Single base substitution Ile50‐Asn − − IC polymerization, stabilization; reduced secretion
PiSevilla Single base substitution Ala58‐Asp − +? IC polymerization, stabilization; reduced secretion
PiCadiz Single base substitution Glu151‐Lys − +? None
PiTarragona Single base substitution Phe227‐Cys − +? IC polymerization, degradation; reduced secretion
 PiPuerto Real Single base substitution Thr249‐Ala − +? IC polymerization, stabilization; altered glycosylation
PiValencia Single base substitution Lys328‐Glu − − IC stabilization (reduced enzyme activity)

STRUCTURE, FUNCTION, most important determinant of functional specificity for each


AND PHYSIOLOGY SERPIN molecule. This concept was dramatically confirmed
by the discovery of α1‐AT Pittsburgh, a variant in which the
α1‐AT is encoded by a 12.2 kb gene containing seven exons and P1 methionine 358 residue of α1‐AT is replaced by arginine.
six introns on human chromosome 14q31‐32.3 (reviewed in In this variant, α1‐AT functions as a thrombin inhibitor, and
[3]). The first three exons and a short 5’ segment of the fourth severe bleeding diathesis results [25]. α1‐AT is an inhibitor
exon encode variable, tissue‐specific 5’ untranslated regions of serine proteases in general, but under physiologic condi-
(UTRs) of the α1‐AT mRNA, which are generated by alternative tions its targets are probably only neutrophil elastase, cathep-
post‐transcriptional exon splicing [22, 23]. Alternate splicing of sin G, and proteinase 3, proteases released by activated
the α1‐AT 5’ UTR results in variable inclusion or exclusion of neutrophils because the kinetics of association with these
long upstream open reading frames (ORFs) whose presence or enzymes are dramatically more favorable than with other ser-
absence alters the efficiency of α1‐AT translation [24] and could ine protease [26].
potentially affect hepatic phenotypes in PIZZ subjects. Most of α1‐AT acts competitively by allowing target enzymes to bind
the fourth exon and the remaining three exons encode the α1‐AT directly to its P1 Met residue‐containing reactive loop. The
protein sequence. The α1‐AT protein is a single‐chain, 55 kDa resulting complex contains one molecule of each reactant. The
polypeptide with 394 amino acids and 3 asparagine‐linked com- complex of α1‐AT and serine protease is a covalently stabilized
plex carbohydrate side‐chains. Two major serum isoforms differ structure, resistant to dissociation by denaturing compounds.
in the configuration of the carbohydrate side‐chains. The interaction between α1‐AT and serine protease is “suicidal”
α1‐AT is the archetype of the SERPIN family of protease in that the inhibitor is irreversibly modified and no longer able
inhibitors [3]. Most SERPINs function as suicide inhibitors to bind or inactivate enzyme. Studies also show that the irrevers-
of specific target proteases. Others are not inhibitory, but ible trapping of the target enzyme is mediated by a profound
rather form complexes with but do not inactivate their hor- conformational change in α1‐AT, which Carrell and Lomas lik-
mone ligands. The reactive site “P1” residue of α1‐AT is the ened to a “mousetrap, with the active inhibitor circulating in the
648 THE LIVER:  MECHANISM OF DEFICIENCY

metastable stressed‐form and then springing into the stable, (LRP) can also mediate clearance and catabolism of α1‐AT‐
relaxed form to lock the complex with its target protease” [27]. elastase complexes [43]. α1‐AT diffuses into most tissues and is
The functional activity of α1‐AT in vivo may be regulated by found in most body fluids [5].
several factors. For one, it may be rendered inactive as an elastase
inhibitor by active oxygen products, intermediates of activated
neutrophils and macrophages that can oxidize the reactive‐site
methionine of α1‐AT [5]. This effect is thought to constitute the DEFICIENCY VARIANTS
basis for increased susceptibility to emphysema in smokers, OF α1‐ANTITRYPSIN
whether deficient in α1‐AT or not. The α1‐AT molecule may also
be inactivated in vivo by proteases including collagenase and Human α1‐AT variants are classified according to the protease
pseudomonas elastase [26, 28]. Several studies indicate that α1‐ inhibitor (PI1,2) phenotype system as defined by agarose electro-
AT protects experimental animals from the lethal effects of tumor phoresis or isoelectric focusing of plasma in polyacrylamide at
necrosis factor (TNF) [29]. This protective effect is thought to be acid pH [44]. The PI classification assigns a letter to variants,
due to inhibition of the synthesis and release of platelet‐activating according to migration of the major isoform, using alphabetic
factor from neutrophils [30], presumably through the inhibition order from anode to cathode, or from low to high isoelectric
of neutrophil‐derived proteases. α1‐AT also appears to have func- point. The most common normal variant migrates to an interme-
tional activities that do not involve inhibition of neutrophil pro- diate isoelectric point, designated M, and according to PI geno-
teases. The C‐terminal fragment of α1‐AT, which can be generated type nomenclature individuals homozygous for this variant are
during the formation of a complex with serine protease or during designated PIMM. The most common severe deficiency allelic
proteolytic inactivation by thiol‐ or metalloproteases, is a potent variant migrates to a high isoelectric point, designated Z, and
neutrophil chemoattractant [31]. individuals homozygous for this variant are designated PIZZ.
The predominant site of synthesis of plasma α1‐AT is the Several null and dysfunctional variants, with absent or reduced
liver [32]. Tissue‐specific expression of α1‐AT in human serum levels or activity, have been reported, some associated
hepatoma cells is directed by structural elements recognized by with premature development of emphysema.
nuclear transcription factors including HNF‐1α and HNF‐1β, In addition to Z, other deficiency variants with reduced serum
C‐EBP, HNF‐4, and HNF‐3 (reviewed in [3]). Plasma concen- α1‐AT concentrations have been reported (Table  50.1). Some
trations of α1‐AT increase three‐ to fivefold during the host are not associated with clinical disease such as the S variant [45,
response to inflammation and/or tissue injury [3]. The source of 46]. Several are associated with premature lung disease
this additional α1‐AT has always been considered the liver; (Table 50.1). With respect to other AT variants associated with
thus, α1‐AT is classified as a hepatic acute phase reactant. liver disease, hepatocyte α1‐AT inclusions and liver disease
Synthesis of α1‐AT in human hepatoma cells (HepG2, Hep3B) were reported in two persons with MMalton and one with
is upregulated by interleukin 6 (IL‐6) but not interleukin 1 MDuarte [3, 18]. One person with the deficiency variant
(IL‐1) or TNF [33]. Plasma concentrations also increase during Siiyama, with emphysema and hepatocyte inclusions but with-
oral contraceptive therapy and pregnancy [34]. out liver disease, was also reported [47]. Several novel defi-
α1‐AT is expressed in monocytes and macrophages [35] and ciency variants with a tendency to accumulate and polymerize
mRNA has been isolated from multiple tissues in transgenic in cell culture models were recently discovered by gene sequenc-
mice [36, 37], but only in some cases have studies distinguished ing of subjects with discrepancies between serum α1‐AT con-
whether such mRNA is in ubiquitous tissue macrophages or centration and targeted variant genotyping for Z and S variants
other cell types. For instance, α1‐AT is synthesized in entero- [48]. Another recent study reported that the Z variant, which is
cytes and intestinal paneth cells, as determined by studies in the most common deficiency variant and the variant most often
intestinal epithelial cell lines, ribonuclease protection assays of associated with α1‐AT deficiency liver disease, is able to heter-
human intestinal RNA, and in situ hybridization analysis in cry- opolymerize with S, Mmalton, and Mwurzburg, in cell culture
ostat sections of human intestinal mucosa [23, 38]. α1‐AT is models [49].
also synthesized by pulmonary epithelial cells, with such syn-
thesis less responsive to IL‐6 than to a related cytokine, oncos-
tatin M [39].
The half‐life of α1‐antitrypsin in plasma is approximately 5 MECHANISM OF DEFICIENCY
days [40]. It is estimated that the daily production rate of α1‐AT
is 34 mg kg−1 body weight, with 33% of the intravascular pool A point mutation results in the substitution of lysine for gluta-
degraded daily. Several physiologic factors may affect the rate of mate 342 [27] and renders the mutant α1‐ATZ molecule
α1‐AT catabolism. First, desialylated α1‐AT is cleared from the impaired in its capacity to transverse the secretory pathway.
circulation in minutes [3], probably via hepatic asialoglycopro- Newly synthesized α1‐ATZ accumulates predominantly in the
tein receptor‐mediated endocytosis. Second, α1‐AT in complex ER, and perhaps other as yet undefined compartments that do
with elastase or proteolytically modified is cleared more rapidly not stain with conventional markers of the ER [50], with rela-
than native α1‐AT [41]. Because its ligand specificity is similar to tively limited proportions reaching the extracellular fluid [51].
that required for in vivo clearance of serpin‐enzyme complexes This impairment is observed in liver cells, macrophages, trans-
(SECs), the SEC receptor may also be involved in the clearance fected cell lines, and iPS (induced pluripotent stem cell)‐derived
and catabolism of α1‐AT‐elastase and other serpin‐enzyme com- hepatocyte‐like cells (iHeps) [3, 50]. Site‐directed mutagenesis
plexes [42]. The low‐density protein receptor‐related protein studies have shown that the single amino acid substitution
α1-ANTITRYPSIN DEFICIENCY
50:  649

(E342K) is sufficient to produce the defect in secretion [50, 52]. intracellular accumulation of ATZ and, in this conceptualiza-
The mutant α1‐ATZ molecule is partially functionally active, tion, polymerization/aggregation is a result of, rather than, the
having about 50% to 80% of the elastase inhibitory capacity of primary defect itself. First, only 18% of the intracellular pool of
wild‐type α1‐ATM [53]. There is a modest increase in the rate ATZ is in polymers at steady state in a mammalian cell line
of in vivo clearance/catabolism of radiolabeled α1‐ATZ com- model that faithfully recapitulates the intracellular accumula-
pared with wild‐type α1‐ATM when infused into normal indi- tion/fate of ATZ [65]. Second, a naturally occurring variant of
viduals, but this difference does not account for the decrease in ATZ, which has the same E342K mutation as ATZ and also a
blood levels of α1‐AT in deficient individuals [41]. C‐terminal truncation, accumulates in the ER even though it
The point mutation also renders the α1‐ATZ molecule prone does not form polymers, suggesting that misfolding is sufficient
to polymerization and aggregation. Carrell and Lomas have to lead to intracellular accumulation of ATZ [66]. Third, the
demonstrated the presence of polymers and aggregates in liver results of Sidhar et al. [63] and Kang et al. [64] do not exclude
biopsy specimens and plasma of patients with homozygous the possibility that diminished secretion of ATZ is partially
PIZZ α1‐antitrypsin deficiency [27]. Their studies suggested ­corrected by the second engineered mutation because this sec-
that a “loop‐sheet insertion” mechanism was responsible for the ond mutation also prevents the primary misfolding defect.
polymerization [54]. Similar polymers have been found in Furthermore, in a very interesting study a small molecule that
the plasma of patients with the PISiiyama α1‐AT variant and the prevents polymerization in vitro does not correct the secretory
PIMMalton α1‐AT variant [20, 55]. The mutation in α1‐AT defect of ATZ in vivo but rather leads to enhanced degradation
PISiiyama (Ser 53 to Phe) [47], and in α1‐ATPIMMalton (Phe [67]. Taken together, the data suggests that misfolding is the
52 deletion) [18] affect residues that provide a ridge for the slid- primary defect and that polymerization is a time‐dependent
ing movement that opens the A sheet. Thus, these mutations effect of misfolding and accumulation. This conceptualization
would be expected to interfere with the insertion of the reactive is also consistent with the domain‐swapping mechanism of
center loop into the gap in the A sheet, and therefore leave the polymerization described by Huntington [57–59] and older
gap in the A sheet available for spontaneous loop‐sheet polym- studies of folding kinetics by Yu et al. [68] in which polymeriza-
erization. It is indeed interesting that hepatocellular α1‐AT tion is viewed as a “kinetic” result of delayed folding of mono-
globules have been observed in a few patients with these two meric ATZ, and explains how some ATZ gets secreted. The
variants. Recent observations suggest that the α1‐ATS variant distinction between misfolding or polymerization as the pri-
also undergoes loop‐sheet polymerization [19] and that this may mary inciting event for intracellular accumulation/impaired
account for its retention in the ER, albeit a milder degree of secretion is important when considering potential therapeutic
retention than that for α1‐ATZ [45]. Moreover, α1‐ATS can approaches. If misfolding is the primary event a therapy that
apparently form heteropolymer with α1‐ATZ [19], providing a prevents polymerization but not misfolding might fail to alter
potential explanation for liver disease in patients with the SZ the accumulation and/or the impaired secretion.
phenotype [56]. In either case, however, there is powerful evidence that the
More recent studies by Huntington and colleagues have sug- tendency for α1‐ATZ to misfold, polymerize and aberrantly
gested a different mechanism for polymerogenic and aggrega- accumulate within the secretory pathway is integral to its pro-
tion‐prone properties of ATZ that involves at least two different teotoxicity and disease‐causing potential. In studies of two fam-
domain‐swapping phenomena [57–59]. This mechanism seems ilies with autosomal dominantly inherited dementia, unique
most consistent with recent characterization of a putative ATZ neuronal inclusion bodies were shown to be associated with
monomer crystal structure [60]. polymers of a neuron‐specific member of the SERPIN family,
By themselves, however, these data do not prove that the neuroserpin [69]. Moreover, the mutation in neuroserpin in one
polymerization of α1‐ATZ results in retention in the ER. In fact, family is homologous to the mutation in the α1‐AT Siiyama
many polypeptides must assemble into oligomeric or polymeric allele that is associated with polymerization and inclusions in
complexes to traverse the ER and reach their destination within the ER of liver cells.
the cell, at the surface of the plasma membrane, or into the In one of the most interesting studies on the mechanism for
extracellular fluid [61]. Evidence that polymerization results in ATZ accumulation in the ER, Nyfeler et al. provided evidence
the retention of α1‐ATZ in the ER was originally attributed to for the lectin ER Golgi intermediate compartment 53 kDa pro-
studies in which the fate of α1‐ATZ was examined after the tein (ERGIC‐53) as an export receptor for α1‐antitrypsin [70].
introduction of additional mutations into the molecule. For Most importantly, ERGIC‐53 failed to recognize mutant
instance, Kim et al. introduced a mutation into the α1‐AT mol- α1‐ATZ. This study did not address whether polymerization
ecule at amino acid 51, F51L [62]. This mutation is remote from prevents α1‐ATZ from being presented to ERGIC‐53 or the
the Z mutation, E342K, but apparently impedes polymerization altered folding pathway of α1‐ATZ prevents an essential ligand
and prevents insertion of a synthetic peptide into the gap in the domain from being available to bind to ERGIC‐53.
A sheet, implying that the mutation leads to closing of this gap.
The double‐mutated F51L α1‐ATZ molecule was less prone
to  polymerization and folded more efficiently in vitro than
α1‐ATZ. Moreover, the introduction of the F51L mutation par- PATHOGENESIS OF LIVER DISEASE
tially corrected the intracellular retention properties of α1‐ATZ
in microinjected Xenopus oocytes [63] and in yeast [64]. Although there are still many gaps in our understanding of how
Nevertheless, several lines of evidence suggest that misfold- the deficiency causes liver disease in some of the homozygotes,
ing is the primary defect responsible for impaired secretion/ we know that aberrant intracellular accumulation of α1‐ATZ is
650 THE LIVER:  PATHWAYS FOR INTRACELLULAR DEGRADATION OF MUTANT α1-ATZ

the unifying characteristic of what would be called a gain‐of‐ For many years it was conceptually difficult to reconcile the
toxic‐function mechanism. Experimental results in transgenic gain‐of‐toxic function mechanism with the observations of
mice provide the most powerful evidence. Transgenic mice car- Sveger, showing that only a subset of α1‐AT‐deficient homozy-
rying the mutant human α1‐ATZ allele develop periodic acid– gotes develop significant liver damage. In 1994, we published
Schiff‐positive, diastase‐resistant intrahepatic globules and liver a series of experiments investigating the hypothesis that a sub-
injury [6, 7]. Because there are normal levels of antielastases in set of the PIZZ population is more susceptible to liver injury
these animals, as directed by endogenous genes, the liver injury because of one or more additional inherited traits or environ-
cannot be attributed to a loss‐of‐function mechanism. The most mental factors that exaggerate either the intracellular accumula-
extensively characterized transgenic mouse model, the PiZ tion of the mutant α1‐ATZ or the cellular pathophysiological
mouse, generated with a human genomic fragment encompass- consequence of mutant α1‐ATZ accumulation [86]. Furthermore,
ing all of the exons and introns of the human ATZ gene [6], we hypothesized that these putative genetic/environmental
recapitulates the hepatic pathology of the human disease with modifiers would affect intracellular degradation of α1‐ATZ or
intrahepatocytic globules reflecting accumulation of misfolded adaptive signaling pathways that are activated by the intracel-
ATZ in the ER, slowly progressing fibrosis, low‐grade inflam- lular accumulation of α1‐ATZ. This hypothetical model was
mation, dysplasia, and increased prevalence of hepatocellular validated by showing a lag in intracellular degradation of α1‐
carcinoma [71, 72]. In work with this mouse model over many ATZ in skin fibroblasts from human PIZZ homozygotes
years we have found that marked progressive nodular regenera- already known to have severe liver disease (“susceptible
tion is the dominant hepatic pathological characteristic, strongly hosts”) as compared to fibroblasts from homozygotes who had
resembling liver pathology in the human disease. no liver disease (“protected hosts”). The fibroblasts had been
There is still relatively limited information about how pro- engineered to express α1‐ATZ using gene transfer by ampho-
teinopathy/intracellular α1‐ATZ accumulation leads to liver tropic recombinant retroviral particles. More recently we
injury. Although structural and functional alterations in mito- found similar results in iPS‐derived hepatocyte cells (iHeps)
chondria and caspase‐3 activation have been observed in liver [50]. The rate of degradation of α1‐ATZ was slower in iHeps
tissue from the PiZ mouse model and from α1‐antitrypsin defi- from susceptible hosts than in iHeps from protected hosts.
ciency patients [73, 74], mitochondrial dysfunction probably Over the years since the 1994 skin fibroblast experiments we
has mostly a cytostatic effect because apoptosis, necrosis, and have characterized the putative pathways for intracellular
inflammation are not major pathological characteristics of the ­degradation of α1‐ATZ and the putative adaptive signaling
liver in α1‐antitrypsin deficiency. pathways activated by intracellular accumulation of α1‐ATZ
The mechanisms responsible for the hepatic fibrotic response (Figure 50.1).
to proteinopathy are also not well understood. Genomic analysis
of a mouse model with hepatocyte‐specific inducible expression
of mutant α1‐ATZ, ideal for elucidating signal transduction
pathways that are activated by the proteinopathy, has identified PATHWAYS FOR INTRACELLULAR
a relatively rich network of downstream targets of the TGFβ DEGRADATION OF MUTANT α1‐ATZ
pathway, including upregulation of connective tissue growth
factor [75], and this pathway is known to play a central role in Early studies using yeast and mammalian cell lines showed that
fibrosis [76]. We have also demonstrated that activation of the the proteasomal pathway participates in intracellular disposal of
NFκB signaling pathway is an important and specific effect of α1‐ATZ by a process that is now known as ER‐associated deg-
cellular ATZ accumulation [77] and is designed to prevent fibro- radation (ERAD) in which the substrate is extracted retrograde
genesis by activating collagen‐degrading metalloproteases [78]. from the ER to the cytoplasm [87, 88]. In fact, α1‐ATZ was one
Many recent studies have also suggested that fibrosis results of the first identified substrates of the ERAD pathway [88].
from proteinopathies in several other tissues. Lung fibrosis has Autophagy was subsequently identified as a second major
been described in several rare diseases characterized by pro- pathway for disposal of misfolded α1‐ATZ [21]. Autophagy is
teinopathy in respiratory epithelial cells, including surfactant an intracellular catabolic pathway by which cells digest subcel-
protein C deficiency and Hermansky–Pudlak syndrome [79, lular structures and cytoplasm to generate amino acids as a sur-
80]. Similarly myocardial fibrosis has been described for desmi- vival mechanism (Figure  50.2). It is characterized by double
nopathy that affects cardiomyocytes [81] and skeletal muscle membrane vacuoles called autophagosomes, which fuse with
fibrosis in inclusion‐body myositis [82]. Interestingly, by lysosomes for degradation of the internal constituents. Increased
enhancing the degradation of misfolded proteins, autophagy has numbers of autophagosomes were observed in human fibroblast
been shown to mitigate cardiac fibrosis from desminopathy [81] cell lines engineered for expression of mutant α1‐ATZ, in the
and skeletal muscle fibrosis from inclusion‐body myositis [83] liver of PiZ transgenic mice and in liver biopsy specimens from
just as it does for hepatic fibrosis in the PiZ model of α1‐antit- patients with ATD. Definitive evidence for the role of autophagy
rypsin deficiency. in ATZ disposal was provided by genetic studies in which α1‐
The liver of the PiZ mouse also shows glycogen depletion [84] ATZ disposal was delayed in autophagy‐deficient [Atg5‐null]
and defective ureagenesis [85]. The latter has been attributed to murine embryonic fibroblast cell lines [89] and Atg6‐null yeast
downregulation of hepatocyte nuclear factor‐4α [85]. Because strains [90, 91].
these functional abnormalities are seen clinically in severe forms The importance of the autophagic pathway in intracellular ATZ
of liver disease, this report provides additional evidence for the degradation has been further validated by recent studies demon-
validity of the PiZ mouse as a model of the human disease. strating that autophagy enhancer drugs promote intracellular ATZ
α1-ANTITRYPSIN DEFICIENCY
50:  651

Figure 50.1  Conceptual model for mechanisms of proteotoxicity in α1‐AT deficiency liver disease. Cellular factors that determine whether an
AT‐deficient individual is protected or susceptible to liver disease. In the susceptible host, there is greater accumulation of misfolded ATZ in the ER
because of subtle alternations in putative proteostasis network regulatory mechanisms. Here these proteostasis regulatory mechanisms are envi-
sioned as either ER degradation or signaling pathways that represent cellular protective responses. Reproduced with permission of Cold Spring
Harbor Laboratory Press from [122].

Figure 50.2  (a) Intracellular molecules and organelles destined for autophagic degradation (termed “cargo”) are surrounded by a double‐mem-
brane structure termed the “isolation membrane” (IM). The mechanisms regulating formation of the IM and its targeting to cargo remain incom-
pletely defined. The IM matures into the autophagosome vacuole, separating the cargo from other cellular compartments, and then fuses with the
lysosome to expose the cargo to degradative enzymes. Image adapted from [123] and available under the terms of the Creative Commons Attribution
License, CCBY 4.0.

disposal and attenuate hepatic fibrosis in the PiZ mouse model of insoluble polymers. However, it is also possible that autophagy
ATD in vivo [71]. One of the concepts originating from studies of plays a role in the disposal of soluble monomeric ATZ that accu-
ATZ disposal in autophagy‐deficient yeast strains is that mulates at levels of expression that exceed the capacity of the
autophagy becomes particularly important at higher levels of ATZ proteasome. Another important result from the studies by Kruse
expression (reviewed in [92]). These results taken together with et al. in autophagy‐deficient yeast showed that a misfolded fibrin-
the structural constraints of the proteasome have led to the sup- ogen variant associated with liver disease in a rare inherited form
position that the proteasomal pathway degrades soluble mono- of hypofibrinogenemia is degraded by autophagy in a manner
meric forms of ATZ whereas autophagy is needed for soluble and almost identical to that of misfolded ATZ [91].
652 THE LIVER:  HEPATIC CARCINOMA IN α1-ANTITRYPSIN DEFICIENCY

Pathways for intracellular disposal of ATZ other than the pro- c‐Jun N‐terminal kinase activation was recently demonstrated
teasomal system and the canonical macroautophagy system are in the liver of the PiZ mouse model, as well as in liver speci-
highly likely. For example, a sortilin‐mediated pathway from mens from patients with α1‐antitrypsin deficiency [97].
Golgi to lysosome has been described to participate in degrada- Interestingly this signaling effect leads to increased expression
tion of ATZ in yeast and mammalian cell line models [90, 93]. of α1‐ATZ and therefore likely plays a role in amplifying the
Another pathway for ATZ disposal which diverges from the pathobiology of misfolded protein accumulation.
canonical autophagy system was recently identified in a power-
ful Caenorhabditis elegans model of ATD and found to be pre-
sent in a mammalian cell line model as well [94]. This pathway
is particularly interesting because it is suppressed by insulin HEPATIC CARCINOMA
signaling and when upregulated by knocking down components IN α1‐ANTITRYPSIN DEFICIENCY
of the insulin signaling pathway it can completely mitigate ATZ
proteotoxicity. Although increased susceptibility to hepatic cancer was shown
years ago [4], there have been only a few studies of potential
mechanisms for carcinogenesis. Theorizing that hepatocellular
hyperproliferation was likely to be involved, Rudnick et  al.
SIGNALING PATHWAYS ACTIVATED BY investigated the proliferation of liver cells by BrdU labeling in
ACCUMULATION OF α1‐ATZ the PiZ mouse model [98]. In these studies hepatocellular pro-
IN THE ENDOPLASMIC RETICULUM liferation was increased around sevenfold in the PiZ mouse
compared to a wild‐type control and this degree of proliferation
To determine which signaling pathways are activated when appeared to reflect the slowly progressing chronic nature of α1‐
α1‐ATZ accumulated in the ER, we developed cell line and antitrypsin deficiency liver disease. Most importantly, by using
mouse model systems with inducible rather than constitutive double immunohistochemical staining it was shown that divid-
expression of α1‐ATZ because the latter would potentially per- ing hepatocytes were almost exclusively the ones that lacked
mit adaptations that could obscure the primary signaling intracellular α1‐ATZ inclusions, called the globule‐devoid
effects. A series of studies using these kinds of systems have hepatocytes. Furthermore it was shown that hyperproliferation
shown that the autophagic response and the nuclear factor κB of globule‐devoid hepatocytes was driven by the number of
(NFκB) signaling pathway, but not the unfolded protein adjacent globule‐containing hepatocytes. This last conclusion
response, are a­ ctivated when α1‐ATZ accumulates in the ER was based on the observation that the number of globule‐­
[75, 77, 89]. containing hepatocytes was markedly increased in male PiZ
Activation of NFκB appears to be a hallmark of α1‐ATZ mice or in testosterone‐treated female PiZ mice and this corre-
­accumulation [75, 77]. In recent studies in which the PiZ mouse lated directly with the degree of hyperproliferation of globule‐
is mated to two different mouse models engineered for absence devoid hepatocytes.
of NFκB signaling we found more severe inflammation, fibrosis, Taken together these observations led to a theory that hepat-
steatosis, dysplasia, and more hepatocytes with globules [78] ocyte hyperproliferation was elicited by crosstalk between
indicating that NFκB signaling is intended to protect the liver globule‐containing and globule‐devoid hepatocytes ([99] and
from the effects of α1‐ATZ accumulation. Interestingly the effect Figure 50.3). The globule‐devoid hepatocytes were viewed as
of NFκB signaling pathway under these circumstances is through younger cells capable of responding to trans‐acting regenera-
a small number of downstream target genes: Egr‐1, a transcrip- tive signals derived from the globule‐containing hepatocytes.
tion factor that is essential for liver regeneration [95]; RGS16, an The globule‐containing hepatocytes were viewed as having
inhibitor of G‐protein signaling that has been implicated in acti- greater proteotoxic α1‐ATZ accumulation and unable to
vation of autophagy [75]; and matrix metalloproteases MMP7 respond to the existing regenerative signals because of the pro-
and MMP12. Each of these targets appear to be mediating parts of teotoxic effect on cell proliferation. Thus, the globule‐contain-
the hepatic response to α1‐ATZ accumulation: decreased prolif- ing hepatocytes are sick but not dead and stimulate the
eration of hepatocytes with massive ATZ accumulation, known as regeneration of the globule‐devoid hepatocytes which have a
globule‐containing hepatocytes, due to Egr1‐downregulation; selective proliferative advantage. Interestingly, the replicative
increased autophagy due to RGS16 upregulation; and counteract- defect in the globule‐containing hepatocytes was shown to be
ing of fibrogenesis by MMP7 and MMP12 upregulation. relative because these cells could proliferate as well as globule‐
The hepatic transcriptomic analysis of the Z mouse also devoid hepatocytes when the regenerative stimulus was par-
shows changes in gene expression indicative of TGFβ signaling ticularly powerful, such as in PiZ mice that survived
and this is consistent with the fibrotic response that represents experimental partial hepatectomy [98]. The nature of the differ-
the dominant pathological characteristics of the liver in α1‐­ ences between the globule‐containing and globule‐devoid cells
antitrypsin deficiency [75]. Furthermore we have recently found is not well elucidated. A study by Linblad et al. has suggested
that accumulation of the α1‐ATZ variant in respiratory epithelial that the globule‐devoid hepatocytes have lesser accumulation
cells elicits fibrosis in the lungs with evidence for fibrosis in the of ATZ [74] and this would be consistent with younger cells
lungs of α1‐antitrypsin deficiency patients with very severe that have had less time to accumulate α1‐ATZ. It is also possi-
COPD (chronic obstructive pulmonary disease) [96]. The mech- ble that globule‐devoid hepatocytes are derived from globule‐
anism by which accumulation of misfolded proteins in the ER containing hepatocytes as they increase capacity for degradation
elicits TGFβ signaling has not been studied. of α1‐ATZ. Several observations militate against this latter
α1-ANTITRYPSIN DEFICIENCY
50:  653

Figure 50.3  Hypothetical model for hepatocarcinogenesis in ATD. Globule‐containing hepatocytes (pale pink) tend to be periportal. They are
“sick but not dead” and generate chronic regenerative signals which can only be received effectively in “trans” by globule‐devoid hepatocytes (deep
pink). The globule‐devoid hepatocytes tend to be in the centrilobular regions. When regenerative signals are received by globule‐devoid hepatocytes
by this crosstalk, it drives mitosis and ultimately carcinogenesis (dark red) in the globule‐devoid regions. Reproduced with permission of John
Wiley & Sons from [99].

possibility. The number of globule‐containing hepatocytes It is interesting to note that hepatocellular carcinoma develops
decrease with age [98], and Ding et al. [100] showed that trans- with aging in male PiZ mice [72] and males with α1‐antitrypsin
planted hepatocytes have a selective proliferative advantage deficiency were also disproportionately affected by hepatic
that also depends on the number of adjacent globule‐containing ­cancer in the autopsy studies of Eriksson et al. [4]. Moreover,
hepatocytes in that it was much more evident in male PiZ mice in cases of hepatocellular carcinoma associated with α1‐anti-
that had significantly more globule‐containing hepatocytes trypsin deficiency one observes a staining pattern in which the
than female PiZ mice. This is associated with enhanced apop- carcinoma is negative for inclusions but surrounded by adjacent
tosis of the host hepatocytes, hepatic repopulation with donor liver cells that are positive for inclusions which is entirely con-
hepatocytes, and resolution of the liver fibrosis that occurs in sistent with the carcinogenesis theory proposed by Rudnick and
untreated PiZ mice [100]. Perlmutter [99].
654 THE LIVER:  TREATMENT

MODIFIERS OF THE HEPATIC Autophagy was considered an excellent target because it is


PHENOTYPYE OF α1‐ANTITRYPSIN specifically activated when α1‐ATZ accumulates in cells and it
also plays a critical role in intracellular disposal of ATZ. At the
DEFICIENCY time when this approach was first investigated several drugs
which could enhance autophagic degradation of other misfolded
Studies designed to identify genetic and environmental modifiers
proteins, such as mutant polyQ proteins that cause Huntington
of the liver disease phenotype in human populations have begun
disease, were being described [109]. It had also become apparent
to appear in the literature in recent years. In one interesting study,
that α1‐antitrypsin deficiency liver disease could more frequently
a single nucleotide polymorphism (SNP) in the MAN1B1 gene
have its onset at 50–65 years of age [16] coincident with the
was found to be statistically over‐represented in a series of infants
decline in autophagy function that is believed to trigger other age‐
with end‐stage liver disease [101]. The variant was shown to
dependent degenerative diseases associated with misfolded
reduce intracellular levels of the mannosidase [102]. Recent
­proteins. Hidvegi et al. first investigated the drug carbamazepine
experiments have shown that Man1B1 is actually localized to the
(CBZ), which is known for its widespread use in humans as an
Golgi but it plays a role in regulation of protein secretion as a part
anticonvulsant and mood stabilizer and found that it enhanced
of the protein quality control network which is recently recog-
autophagic degradation of α1‐ATZ in mammalian cell line mod-
nized to be localized in the Golgi [103]. Furthermore those
els [71]. Moreover, administration of this drug by oral gavage to
­experiments have provided a basis for how reduced levels of
the PiZ mouse model over a three‐week period significantly
Man1B1 could theoretically lead to greater intracellular
reduced hepatic α1‐ATZ load and hepatic fibrosis in vivo. Because
α1‐ATZ ­accumulation [103]. These results for the Man1B1 vari-
CBZ is already FDA‐approved it could be moved immediately
ant would appear to validate our hypothesis that intracellular deg-
into a phase 2/3 clinical trial for treatment of severe liver disease
radation pathways are targets of liver disease modifiers but further
due to ATD. Several other drugs with autophagy enhancer proper-
population studies of this variant would be reassuring [104, 105].
ties have been identified by high‐throughput screening of drug
A SNP in the upstream flanking region of the α1‐antitrypsin
libraries and these are currently under investigation [92].
gene has also been implicated in susceptibility to liver disease
A number of relatively new gene therapy strategies are being
[106]. However, the nature of that variant could not be recon-
investigated. One of these involves new methods for silencing
ciled with how it might affect liver disease susceptibility and
gene expression using vectors that are also capable of encoding
its  statistical association with variation in the liver disease
wild‐type α1‐antitrypsin to address both gain‐ and loss‐of
­phenotype was dependent on a questionable classification of
function sequelae of α1‐antitrypsin deficiency, respectively
­
population subgroups.
[110, 111]. In one approach, Li et al. utilized adeno‐associated
Our hypothesis for variation in liver disease susceptibility
virus harboring short‐hairpin RNA to knockdown endogenous
also identifies signaling pathways that could increase or
α1‐ATZ expression together with a codon‐optimized wild‐type
decrease α1‐ATZ proteotoxicity as potential targets for disease
α1‐antitrypsin transgene cassette [110]. In another approach,
modifiers. As of yet we have not encountered an example of this
Mueller et  al. utilized an adeno‐associated virus harboring
potential scenario, but we predict that further studies of iHeps
microRNA to silence endogenous α1‐ATZ gene expression
from patients with different forms of α1‐antitrypsin deficiency
together with a microRNA‐resistant wild‐type α1‐antitrypsin
liver disease, in terms of age of onset and type of hepatic pathol-
gene [111]. In each case hepatic α1‐ATZ load was reduced and
ogy, will identify such a mechanism in the near future.
levels of human α1‐antitrypsin increased in the serum of a trans-
genic mouse model. However, the effect on liver fibrosis by this
strategy was not as compelling and hence further studies are
TREATMENT required to test whether more potent and widespread silencing
would be more effective. Another study using antisense oligo-
The most important principle in the treatment of α1‐AT nucleotides by systemic administration to silence α1‐ATZ gene
­deficiency is avoidance of cigarette smoking. Cigarette smoking expression had a more impressive effect in reducing hepatic
markedly accelerates the destructive lung disease that is associ- fibrosis in the PiZ mouse model system [112].
ated with α1‐AT deficiency, reduces the quality of life, and Another potential gene therapy approach being investigated
­significantly shortens the longevity of these individuals [107]. is the transfer of genes that activate autophagy and therein
There is no specific therapy for α1‐AT deficiency‐associated reduce α1‐ATZ accumulation and proteotoxicity. Pastore et al
liver disease. Therefore, clinical care largely involves support- have championed this approach using TFEB, a master transcrip-
ive management of symptoms due to liver dysfunction and for tional activator of the autophagolysosomal system [113]. Using
the prevention of complications. Progressive liver dysfunction helper‐dependent adenovirus for systemic delivery of TFEB and
in α1‐AT‐deficient patients has been treated by orthotopic liver targeting of its expression to liver, this approach significantly
transplantation, with 5‐year survival rates approaching 90% in reduced hepatic α1‐ATZ load and liver fibrosis in the PiZ mouse
children and 80% in adults at 1 year and 80% at 5 years [108]. model. In vitro studies also validated that TFEB reduces cellular
Several novel strategies for treatment of α1‐antitrypsin defi- α1‐ATZ levels in an autophagy‐dependent manner [113].
ciency liver disease that would obviate the need for organ trans- Although it will not address the loss‐of‐function mechanisms
plantation and chronic immunosuppression are currently under associated with α1‐antitrypsin deficiency lung disease, gene
investigation and at various stages of development. One of the therapy with TFEB or drugs that target TFEB activation
relatively newer strategies targets intracellular degradation ­constitute exciting potential therapeutic strategies for liver dis-
­pathways using autophagy enhancer drugs. ease associated with α1‐antitrypsin deficiency.
α1-ANTITRYPSIN DEFICIENCY
50:  655

Ultimately, genomic editing will be considered to definitively can repopulate almost the entire liver of the PiZ mouse model
correct the genetic defect that causes ATD. The most recent [100]. In the PIZ mouse model, the donor cells replaced both
development in this area, CRISPR/Cas‐9‐mediated genome globule‐containing and globule‐devoid cells, indicating that
editing, has been used in the PiZ mouse model in vivo, showing both types of affected hepatocytes have impaired proliferative
reduced α1‐ATZ in the liver and low levels of wild‐type α1AT in capacity compared to wild‐type hepatocytes. Because the trans-
the liver [114] and in serum [115]. planted hepatocytes have a selective proliferative advantage
Several research groups are exploring a “structure‐based” over α1‐ATZ‐containing endogenous hepatocytes and can sub-
screening strategy that aims to generate peptides to prevent stitute for the latter in a diseased liver, this option of therapy
polymerization of the mutant α1‐ATZ with the hypothesis that may be considered for α1‐antitrypsin deficiency lung and liver
this would facilitate secretion. A small molecule compound disease.
designed against a lateral hydrophobic cavity in α1‐ATZ pre- Another exciting therapeutic strategy in which genomic edit-
vented its polymerization, however, further experiments in a cell ing is combined with hepatocyte transplantation has been tested
line model revealed that this compound enhanced intracellular in a transgenic mouse model of α1‐antitrypsin deficiency.
degradation only, with minimal effect on secretion [67]. These Studies by Yusa et  al. have shown that the mutation in the
results provide further evidence that it is the misfolding of α1‐antitrypsin gene could be corrected in human induced pluri-
α1‐ATZ, independent of its tendency to polymerize, which is potent stem (iPS) cells derived from a α1‐antitrypsin deficiency
primarily responsible for impaired secretion. Another small patient using a combination of zinc‐finger nucleases and trans-
molecule based on a peptide that targets the reactive center loop poson techniques [121]. Importantly, the corrected iPS cell lines
of α1‐antitrypsin has been designed and introduced in cell line could then be engrafted into the liver of the transgenic mouse
model systems with evidence for improved secretion of α1‐ATZ model system and, based on the observations of Ding et  al.
[116]. However, the efficacy of this peptide in an animal model [100], the corrected cells should expand significantly because
system for either increasing secretion or reducing liver damage they will have a selective proliferative advantage. This strategy,
in vivo remains to be tested. It also remains possible that this if it proves successful in further preclinical models, has the
type of peptide binding changes the conformation of the mutant potential to address both the loss‐ and gain‐of‐function mecha-
protein in such a way that both misfolding and polymerization nisms of organ damage and the advantage of personalized treat-
are reduced independently. ment options without any need for immunosuppression.
Chemical chaperones that can non‐selectively facilitate fold-
ing of diverse misfolded proteins have also been investigated as
a potential therapeutic option for α1‐antitrypsin deficiency liver
disease. Glycerol and 4‐phenylbutyric acid (PBA) were found to REFERENCES
mediate a robust enhancement in the secretion of α1‐ATZ in a
mammalian cell line model and its oral administration in PiZ   1. Sveger, T. Liver disease in alpha1‐antitrypsin deficiency detected by screen-
ing of 200,000 infants. N Engl J Med, 1976;294(24):1316–21.
mice increased blood levels of human α1‐antitrypsin reaching   2. Silverman, E.K. and Sandhaus, R.A. Clinical practice. Alpha1‐antitrypsin
20–50% of the levels present in PiM mice and normal humans deficiency. N Engl J Med, 2009;360(26):2749–57.
[117]. However, a pilot clinical trial involving 10 patients with   3. Teckman, J.H., Qu, D., and Perlmutter, D.H. Molecular pathogenesis of liver
α1‐antitrypsin deficiency‐associated liver disease, failed to disease in alpha1‐antitrypsin deficiency. Hepatology, 1996;24(6):1504–16.
reveal any significant increase in serum levels of α1‐antitrypsin   4. Eriksson, S., Carlson, J., and Velez, R. Risk of cirrhosis and primary liver
cancer in alpha 1‐antitrypsin deficiency. N Engl J Med, 1986;314(12):736–9.
after 14 days of treatment with PBA [118]. It is not clear why the   5. Crystal, R.G. Alpha 1‐antitrypsin deficiency, emphysema, and liver disease.
drug lacked effect but the large doses required are known to be Genetic basis and strategies for therapy. J Clin Invest, 1990;85(5):1343–52.
quite challenging to tolerate and so it may be worthwhile to test   6. Carlson, J.A., Rogers, B.B., Sifers, R.N. et al. Accumulation of PiZ alpha
in the future if newer, more tolerable formulations are developed. 1‐antitrypsin causes liver damage in transgenic mice. J Clin Invest, 1989;
83(4):1183–90.
Recently, suberoylanilide hydroxamic acid (SAHA), another
  7. Dycaico, M.J., Grant, S.G., Felts, K. et al. Neonatal hepatitis induced by alpha
drug which has many pharmacological similarities to PBA, has 1‐antitrypsin: a transgenic mouse model. Science, 1988;242(4884):1409–12.
been found to enhance secretion of α1‐ATZ in cell line models of   8. Mostafavi, B., Diaz, S., Tanash, H.A., and Piitulainen, E. Liver function in
α1‐antitrypsin deficiency [119]. However, SAHA has not yet alpha‐1‐antitrypsin deficient individuals at 37 to 40 years of age. Medicine
been tested in animal models. Moreover, detailed studies are (Baltimore), 2017;96(12):e6180.
  9. Silverman, E.K., Province, M.A., Rao, D.C., Pierce, J.A., and Campbell, E.J.
needed to delineate whether this effect is due to increased syn-
A family study of the variability of pulmonary function in alpha 1‐antitrypsin
thesis of α1‐ATZ or due to its ability to reduce α1‐ATZ accumu- deficiency. Quantitative phenotypes. Am Rev Respir Dis, 1990;142(5):
lation in cells or both. If increased secretion of α1‐ATZ is because 1015–21.
of, or even associated with, increased synthesis, the treatment 10. Crystal, R.G. Augmentation treatment for alpha1 antitrypsin deficiency.
could produce more rather than less cellular proteotoxicity. Lancet, 2015;386(9991):318–20.
11. McElvaney, N.G., Burdon, J., Holmes, M. et  al. Long‐term efficacy and
Hepatocyte transplantation therapy has also been investigated safety of alpha1 proteinase inhibitor treatment for emphysema caused by
as a potential treatment for α1‐antitrypsin deficiency. It has been severe alpha1 antitrypsin deficiency: an open‐label extension trial (RAPID‐
tested in the past as a treatment for several metabolic liver dis- OLE). Lancet Respir Med, 2017;5(1):51–60.
eases [120]. Compared to orthotopic liver transplantation it has 12. Ibarguen, E., Gross, C.R., Savik, S.K., and Sharp, H.L. Liver disease in
the advantage of being a minimally invasive procedure with lit- alpha‐1‐antitrypsin deficiency: prognostic indicators. J Pediatr, 1990;
117(6):864–70.
tle known morbidity, and is considerably less expensive than 13. Sharp, H.L., Bridges, R.A., Krivit, W., and Freier, E.F. Cirrhosis associated
protein replacement therapy or liver transplantation. Importantly, with alpha‐1‐antitrypsin deficiency: a previously unrecognized inherited dis-
recent studies have revealed that wild‐type donor hepatocytes order. J Lab Clin Med, 1969;73(6):934–9.
656 THE LIVER:  REFERENCES

14. Hadchouel, M. and Gautier, M. Histopathologic study of the liver in the 37. Carlson, J.A., Rogers, B.B., Sifers, R.N., Hawkins, H.K., Finegold, M.J., and
early  cholestatic phase of alpha‐1‐antitrypsin deficiency. J Pediatr, 1976; Woo, S.L. Multiple tissues express alpha 1‐antitrypsin in transgenic mice
89(2):211–5. and man. J Clin Invest, 1988;82(1):26–36.
15. Sveger, T. The natural history of liver disease in alpha 1‐antitrypsin deficient 38. Molmenti, E.P., Perlmutter, D.H., and Rubin, D.C. Cell‐specific expression
children. Acta Paediatr Scand, 1988;77(6):847–51. of alpha 1‐antitrypsin in human intestinal epithelium. J Clin Invest,
16. Chu, A.S., Chopra, K.B., and Perlmutter, D.H. Is severe progressive liver 1993;92(4):2022–34.
disease caused by alpha‐1‐antitrypsin deficiency more common in children 39. Cichy, J., Potempa, J., and Travis, J. Biosynthesis of alpha1‐proteinase inhib-
or adults? Liver Transpl, 2016;22(7):886–94. itor by human lung‐derived epithelial cells. J Biol Chem, 1997;272(13):
17. Schaefer, B., Mandorfer, M., Viveiros, A. et  al. Heterozygosity for the 8250–5.
alpha‐1‐antitrypsin Z allele in cirrhosis is associated with more advanced 40. Laurell, C.B., Nosslin, B., and Jeppsson, J.O. Catabolic rate of alpha1‐antit-
disease. Liver Transpl, 2018;24(6):744–51. rypsin of Pi type M and Z in man. Clin Sci Mol Med, 1977;52(5):457–61.
18. Curiel, D.T., Holmes, M.D., Okayama, H. et al. Molecular basis of the liver 41. Mast, A.E., Enghild, J.J., Pizzo, S.V., and Salvesen, G. Analysis of the plasma
and lung disease associated with the alpha 1‐antitrypsin deficiency allele elimination kinetics and conformational stabilities of native, proteinase‐com-
Mmalton. J Biol Chem, 1989;264(23):13938–45. plexed, and reactive site cleaved serpins: comparison of alpha 1‐proteinase
19. Mahadeva, R., Chang, W.S., Dafforn, T.R. et al. Heteropolymerization of S, inhibitor, alpha 1‐antichymotrypsin, antithrombin II.I., alpha 2‐antiplasmin,
I, and Z alpha1‐antitrypsin and liver cirrhosis. J Clin Invest, 1999;103(7): angiotensinogen, and ovalbumin. Biochemistry, 1991;30(6):1723–30.
999–1006. 42. Joslin, G., Wittwer, A., Adams, S., Tollefsen, D.M., August, A., and
20. Lomas, D.A., Elliott, P.R., Sidhar, S.K. et al. alpha 1‐Antitrypsin Mmalton Perlmutter, D.H. Cross‐competition for binding of alpha 1‐antitrypsin (alpha
(Phe52‐deleted) forms loop‐sheet polymers in vivo. Evidence for the C sheet 1 AT)‐elastase complexes to the serpin‐enzyme complex receptor by other
mechanism of polymerization. J Biol Chem, 1995;270(28):16864–70. serpin‐enzyme complexes and by proteolytically modified alpha 1 A.T. J
21. Teckman, J.H. and Perlmutter, D.H. Retention of mutant alpha(1)‐antitrypsin Biol Chem, 1993;268(3):1886–93.
Z in endoplasmic reticulum is associated with an autophagic response. Am J 43. Kounnas, M.Z., Church, F.C., Argraves, W.S., and Strickland, D.K. Cellular
Physiol Gastrointest Liver Physiol, 2000;279(5):G961–74. internalization and degradation of antithrombin III‐thrombin, heparin cofac-
22. Perlino, E., Cortese, R., and Ciliberto, G. The human alpha 1‐antitrypsin tor II‐thrombin, and alpha 1‐antitrypsin‐trypsin complexes is mediated by
gene is transcribed from two different promoters in macrophages and hepat- the low density lipoprotein receptor‐related protein. J Biol Chem, 1996;
ocytes. EMBO J, 1987;6(9):2767–71. 271(11):6523–9.
23. Hafeez, W., Ciliberto, G., and Perlmutter, D.H. Constitutive and modulated 44. Pierce, J.A. and Eradio, B.G. Improved identification of antitrypsin pheno-
expression of the human alpha 1 antitrypsin gene. Different transcriptional types through isoelectric focusing with dithioerythritol. J Lab Clin Med,
initiation sites used in three different cell types. J Clin Invest, 1992;89(4): 1979;94(6):826–31.
1214–22. 45. Teckman, J.H. and Perlmutter, D.H. The endoplasmic reticulum degradation
24. Corley, M., Solem, A., Phillips, G. et al. An RNA structure‐mediated, post- pathway for mutant secretory proteins alpha1‐antitrypsin Z and S is distinct
transcriptional model of human alpha‐1‐antitrypsin expression. Proc Natl from that for an unassembled membrane protein. J Biol Chem, 1996;
Acad Sci USA, 2017;114(47):E10244–53. 271(22):13215–20.
25. Owen, M.C., Brennan, S.O., Lewis, J.H., and Carrell, R.W. Mutation of anti- 46. Long, G.L., Chandra, T., Woo, S.L., Davie, E.W., and Kurachi, K. Complete
trypsin to antithrombin. alpha 1‐antitrypsin Pittsburgh (358 Met leads to sequence of the cDNA for human alpha 1‐antitrypsin and the gene for the S
Arg), a fatal bleeding disorder. N Engl J Med, 1983;309(12):694–8. variant. Biochemistry, 1984;23(21):4828–37.
26. Mast, A.E., Enghild, J.J., Nagase, H., Suzuki, K., Pizzo, S.V., ND Salvesen, 47. Seyama, K., Nukiwa, T., Takabe, K., Takahashi, H., Miyake, K., and Kira, S.
G. Kinetics and physiologic relevance of the inactivation of alpha 1‐protein- Siiyama (serine 53 (TCC) to phenylalanine 53 (TTC)). A new alpha 1‐antit-
ase inhibitor, alpha 1‐antichymotrypsin, and antithrombin III by matrix met- rypsin‐deficient variant with mutation on a predicted conserved residue of
alloproteinases‐1 (tissue collagenase), ‐2 (72‐kDa gelatinase/type IV the serpin backbone. J Biol Chem, 1991;266(19):12627–32.
collagenase), and ‐3 (stromelysin). J Biol Chem, 1991;266(24):15810–6. 48. Matamala, N., Lara, B., Gomez‐Mariano, G. et al. Characterization of novel
27. Carrell, R.W. AND Lomas, D.A. Conformational disease. Lancet, 1997; missense variants of SERPINA1 gene causing alpha‐1 antitrypsin deficiency.
350(9071):134–8. Am J Respir Cell Mol Biol, 2018;58(6):706–16.
28. Vissers, M.C., George, P.M., Bathurst, I.C., Brennan, S.O., and Winterbourn, 49. Laffranchi, M., Berardelli, R., Ronzoni, R., Lomas, D.A., and Fra, A.
C.C. Cleavage and inactivation of alpha 1‐antitrypsin by metalloproteinases Heteropolymerization of alpha‐1‐antitrypsin mutants in cell models mimick-
released from neutrophils. J Clin Invest, 1988;82(2):706–11. ing heterozygosity. Hum Mol Genet, 2018;27(10):1785–93.
29. Van Molle, W., Libert, C., Fiers, W., and Brouckaert, P. Alpha 1‐acid glyco- 50. Tafaleng, E.N., Chakraborty, S., Han, B. et al. Induced pluripotent stem cells
protein and alpha 1‐antitrypsin inhibit TNF‐induced but not anti‐Fas‐induced model personalized variations in liver disease resulting from alpha1‐antit-
apoptosis of hepatocytes in mice. J Immunol, 1997;159(7):3555–64. rypsin deficiency. Hepatology, 2015;62(1):147–57.
30. Camussi, G., Tetta, C., Bussolino, F., and Baglioni, C. Synthesis and release 51. Perlmutter, D.H., Kay, R.M., Cole, F.S., Rossing, T.H., Van Thiel, D., and
of platelet‐activating factor is inhibited by plasma alpha 1‐proteinase inhibi- Colten, H.R. The cellular defect in alpha 1‐proteinase inhibitor (alpha 1‐PI)
tor or alpha 1‐antichymotrypsin and is stimulated by proteinases. J Exp Med, deficiency is expressed in human monocytes and in Xenopus oocytes
1988;168(4):1293–306. injected with human liver mRNA. Proc Natl Acad Sci USA, 1985;82(20):
31. Joslin, G., Griffin, G.L., August, A.M. et  al. The serpin‐enzyme complex 6918–21.
(SEC) receptor mediates the neutrophil chemotactic effect of alpha‐1 antit- 52. McCracken, A.A., Kruse, K.B., and Brown, J.L. Molecular basis for defec-
rypsin‐elastase complexes and amyloid‐beta peptide. J Clin Invest, tive secretion of the Z variant of human alpha‐1‐proteinase inhibitor: secre-
1992;90(3):1150–4. tion of variants having altered potential for salt bridge formation between
32. Hood, J.M., Koep, L.J., Peters, R.L. et al. Liver transplantation for advanced amino acids 290 and 342. Mol Cell Biol, 1989;9(4):1406–14.
liver disease with alpha‐1‐antitrypsin deficiency. N Engl J Med, 1980; 53. Lomas, D.A., Evans, D.L., Stone, S.R., Chang, W.S., and Carrell, R.W.
302(5):272–5. Effect of the Z mutation on the physical and inhibitory properties of alpha
33. Perlmutter, D.H., May, L.T., and Sehgal, P.B. Interferon beta 2/interleukin 6 1‐antitrypsin. Biochemistry, 1993;32(2):500–8.
modulates synthesis of alpha 1‐antitrypsin in human mononuclear phago- 54. Lomas, D.A., Evans, D.L., Finch, J.T., and Carrell, R.W. The mechanism of Z
cytes and in human hepatoma cells. J Clin Invest, 1989;84(1):138–44. alpha 1‐antitrypsin accumulation in the liver. Nature, 1992;357(6379):605–7.
34. Laurell, C.B. and Rannevik, G. A comparison of plasma protein changes 55. Lomas, D.A., Finch, J.T., Seyama, K., Nukiwa, T., and Carrell, R.W. Alpha
induced by danazol, pregnancy, and estrogens. J Clin Endocrinol Metab, 1‐antitrypsin Siiyama (Ser53‐‐>Phe). Further evidence for intracellular loop‐
1979;49(5):719–25. sheet polymerization. J Biol Chem, 1993;268(21):15333–5.
35. Perlmutter, D.H., Cole, F.S., Kilbridge, P., Rossing, T.H., and Colten, H.R. 56. Dafforn, T.R., Mahadeva, R., Elliott, P.R., Sivasothy, P., and Lomas, D.A. A
Expression of the alpha 1‐proteinase inhibitor gene in human monocytes and kinetic mechanism for the polymerization of alpha1‐antitrypsin. J Biol
macrophages. Proc Natl Acad Sci USA, 1985;82(3):795–9. Chem, 1999;274(14):9548–55.
36. Kelsey, G.D., Povey, S., Bygrave, A.E., and Lovell‐Badge, R.H. Species‐ and 57. Yamasaki, M., Li, W., Johnson, D.J., and Huntington, J.A. Crystal structure
tissue‐specific expression of human alpha 1‐antitrypsin in transgenic mice. of a stable dimer reveals the molecular basis of serpin polymerization.
Genes Dev, 1987;1(2):161–71. Nature, 2008;455(7217):1255–8.
α1-ANTITRYPSIN DEFICIENCY
50:  657

58. Whisstock, J.C., Silverman, G.A., Bird, P.I. et al. Serpins flex their muscle:   82. Doppler, K., Mittelbronn, M., Lindner, A., and Bornemann, A. Basement
I.I. Structural insights into target peptidase recognition, polymerization, and membrane remodelling and segmental fibrosis in sporadic inclusion body
transport functions. J Biol Chem, 2010;285(32):24307–12. myositis. Neuromuscul Disord, 2009;19(6):406–11.
59. Yamasaki, M., Sendall, T.J., Pearce, M.C., Whisstock, J.C., and Huntington,   83. Nogalska, A., D’Agostino, C., Terracciano, C., Engel, W.K., and Askanas,
J.A. Molecular basis of alpha1‐antitrypsin deficiency revealed by the struc- V. Impaired autophagy in sporadic inclusion‐body myositis and in endo-
ture of a domain‐swapped trimer. EMBO Rep, 2011;12(10):1011–7. plasmic reticulum stress‐provoked cultured human muscle fibers. Am J
60. Huang, X., Zheng, Y., Zhang, F. et al. Molecular mechanism of Z alpha1‐ Pathol, 2010;177(3):1377–87.
antitrypsin deficiency. J Biol Chem, 2016;291(30):15674–86.   84. Hubner, R.H., Leopold, P.L., Kiuru, M., De, B.P., Krause, A., and Crystal,
61. Hurtley, S.M. and Helenius, A. Protein oligomerization in the endoplasmic R.G. Dysfunctional glycogen storage in a mouse model of alpha1‐antit-
reticulum. Annu Rev Cell Biol, 1989;5:277–307. rypsin deficiency. Am J Respir Cell Mol Biol, 2009;40(2):239–47.
62. Kim, J., Lee, K.N., Yi, G.S., and Yu, M.H. A thermostable mutation located   85. Piccolo, P., Annunziata, P., Soria, L.R. et al. Down‐regulation of hepatocyte
at the hydrophobic core of alpha 1‐antitrypsin suppresses the folding defect nuclear factor‐4alpha and defective zonation in livers expressing mutant Z
of the Z‐type variant. J Biol Chem, 1995;270(15):8597–601. alpha1‐antitrypsin. Hepatology, 2017;66(1):124–35.
63. Sidhar, S.K., Lomas, D.A., Carrell, R.W., and Foreman, R.C. Mutations  86. Wu, Y., Whitman, I., Molmenti, E., Moore, K., Hippenmeyer, P., and
which impede loop/sheet polymerization enhance the secretion of human Perlmutter, D.H. A lag in intracellular degradation of mutant alpha 1‐antit-
alpha 1‐antitrypsin deficiency variants. J Biol Chem, 1995;270(15):8393–6. rypsin correlates with the liver disease phenotype in homozygous PiZZ alpha
64. Kang, H.A., Lee, K.N., and Yu, M.H. Folding and stability of the Z and 1‐antitrypsin deficiency. Proc Natl Acad Sci USA, 1994;91(19):9014–8.
S(iiyama) genetic variants of human alpha1‐antitrypsin. J Biol Chem,   87. Qu, D., Teckman, J.H., Omura, S., and Perlmutter, D.H. Degradation of a
1997;272(1):510–6. mutant secretory protein, alpha1‐antitrypsin Z., in the endoplasmic reticu-
65. Schmidt, B.Z. and Perlmutter, D.H. Grp78, Grp94, and Grp170 interact with lum requires proteasome activity. J Biol Chem, 1996;271(37):22791–5.
alpha1‐antitrypsin mutants that are retained in the endoplasmic reticulum.   88. Werner, E.D., Brodsky, J.L., and McCracken, A.A. Proteasome‐dependent
Am J Physiol Gastrointest Liver Physiol, 2005;289(3):G444–55. endoplasmic reticulum‐associated protein degradation: an unconventional
66. Lin, L., Schmidt, B., Teckman, J., and Perlmutter, D.H. A naturally occurring route to a familiar fate. Proc Natl Acad Sci USA, 1996;93(24):13797–801.
nonpolymerogenic mutant of alpha 1‐antitrypsin characterized by prolonged   89. Kamimoto, T., Shoji, S., Hidvegi, T. et al. Intracellular inclusions contain-
retention in the endoplasmic reticulum. J Biol Chem, 2001;276(36):33893–8. ing mutant alpha1‐antitrypsin Z are propagated in the absence of autophagic
67. Mallya, M., Phillips, R.L., Saldanha, S.A. et al. Small molecules block the activity. J Biol Chem, 2006;281(7):4467–76.
polymerization of Z alpha1‐antitrypsin and increase the clearance of intra-   90. Kruse, K.B., Brodsky, J.L., and McCracken, A.A. Characterization of an
cellular aggregates. J Med Chem, 2007;50(22):5357–63. ERAD gene as VPS30/ATG6 reveals two alternative and functionally dis-
68. Yu, M.H., Lee, K.N., and Kim, J. The Z type variation of human alpha 1‐anti- tinct protein quality control pathways: one for soluble Z variant of human
trypsin causes a protein folding defect. Nat Struct Biol, 1995;2(5):363–7. alpha‐1 proteinase inhibitor (A1PiZ) and another for aggregates of A1PiZ.
69. Davis, R.L., Shrimpton, A.E., Holohan, P.D., et al. Familial dementia caused Mol Biol Cell, 2006;17(1):203–12.
by polymerization of mutant neuroserpin. Nature, 1999;401(6751):376–9.   91. Kruse, K.B., Dear, A., Kaltenbrun, E.R. et  al. Mutant fibrinogen cleared
70. Nyfeler, B., Reiterer, V., Wendeler, M.W. et al. Identification of ERGIC‐53 from the endoplasmic reticulum via endoplasmic reticulum‐associated pro-
as an intracellular transport receptor of alpha1‐antitrypsin. J Cell Biol, tein degradation and autophagy: an explanation for liver disease. Am J
2008;180(4):705–12. Pathol, 2006;168(4):1299–308; quiz 404–5.
71. Hidvegi, T., Ewing, M., Hale, P. et  al. An autophagy‐enhancing drug pro-   92. Wang, Y. and Perlmutter, D.H. Targeting intracellular degradation pathways
motes degradation of mutant alpha1‐antitrypsin Z and reduces hepatic fibro- for treatment of liver disease caused by alpha1‐antitrypsin deficiency.
sis. Science, 2010;329(5988):229–32. Pediatr Res, 2014;75(1–2):133–9.
72. Marcus, N.Y., Brunt, E.M., Blomenkamp, K. et al. Characteristics of hepato-   93. Gelling, C.L., Dawes, I.W., Perlmutter, D.H., Fisher, E.A., and Brodsky,
cellular carcinoma in a murine model of alpha‐1‐antitrypsin deficiency. J.L. The endosomal protein‐sorting receptor sortilin has a role in trafficking
Hepatol Res, 2010;40(6):641–53. alpha‐1 antitrypsin. Genetics, 2012;192(3):889–903.
73. Teckman, J.H., An, J.K., Blomenkamp, K., Schmidt, B., and Perlmutter, D.   94. Long, O.S., Benson, J.A., Kwak, J.H. et al. A C. elegans model of human
Mitochondrial autophagy and injury in the liver in alpha 1‐antitrypsin defi- alpha1‐antitrypsin deficiency links components of the RNAi pathway to
ciency. Am J Physiol Gastrointest Liver Physiol, 2004;286(5):G851–62. misfolded protein turnover. Hum Mol Genet, 2014;23(19):5109–22.
74. Lindblad, D., Blomenkamp, K., and Teckman, J. Alpha‐1‐antitrypsin mutant   95. Liao, Y., Shikapwashya, O.N., Shteyer, E., Dieckgraefe, B.K., Hruz, P.W.,
Z protein content in individual hepatocytes correlates with cell death in a and Rudnick, D.A. Delayed hepatocellular mitotic progression and
mouse model. Hepatology, 2007;46(4):1228–35. impaired liver regeneration in early growth response‐1‐deficient mice. J
75. Hidvegi, T., Mirnics, K., Hale, P., Ewing, M., Beckett, C., and Perlmutter, Biol Chem, 2004;279(41):43107–16.
D.H. Regulator of G signaling 16 is a marker for the distinct endoplasmic   96. Hidvegi, T., Stolz, D.B., Alcorn, J.F. et al. Enhancing autophagy with drugs
reticulum stress state associated with aggregated mutant alpha1‐antitrypsin or lung‐directed gene therapy reverses the pathological effects of respira-
Z in the classical form of alpha1‐antitrypsin deficiency. J Biol Chem, tory epithelial cell proteinopathy. J Biol Chem, 2015;290(50):
2007;282(38):27769–80. 29742–57.
76. Dooley, S., Hamzavi, J., Ciuclan, L. et  al. Hepatocyte‐specific Smad7   97. Pastore, N., Attanasio, S., Granese, B. et al. Activation of the c‐Jun N‐ter-
expression attenuates TGF‐beta‐mediated fibrogenesis and protects against minal kinase pathway aggravates proteotoxicity of hepatic mutant Z
liver damage. Gastroenterology, 2008;135(2):642–59. alpha1‐antitrypsin. Hepatology, 2017;65(6):1865–74.
77. Hidvegi, T., Schmidt, B.Z., Hale, P., and Perlmutter, D.H. Accumulation of  98. Rudnick, D.A., Liao, Y., An, J.K., Muglia, L.J., Perlmutter, D.H., and
mutant alpha1‐antitrypsin Z in the endoplasmic reticulum activates cas- Teckman, J.H. Analyses of hepatocellular proliferation in a mouse model of
pases‐4 and ‐12, NFkappaB, and BAP31 but not the unfolded protein alpha‐1‐antitrypsin deficiency. Hepatology, 2004;39(4):1048–55.
response. J Biol Chem, 2005;280(47):39002–15.   99. Rudnick, D.A. and Perlmutter, D.H. Alpha‐1‐antitrypsin deficiency: a new
78. Mukherjee, A., Hidvegi, T., Araya, P., Ewing, M., Stolz, D.B., and Perlmutter, paradigm for hepatocellular carcinoma in genetic liver disease. Hepatology,
D.H. NFkappaB mitigates the pathological effects of misfolded alpha1‐anti- 2005;42(3):514–21.
trypsin by activating autophagy and an integrated program of proteostasis 100. Ding, J., Yannam, G.R., Roy‐Chowdhury, N. et  al. Spontaneous hepatic
mechanisms. Cell Death Differ, 2019;26(3)455–69. repopulation in transgenic mice expressing mutant human alpha1‐antit-
79. Bridges, J.P., Wert, S.E., Nogee, L.M., and Weaver, T.E. Expression of a rypsin by wild‐type donor hepatocytes. J Clin Invest, 2011;121(5):
human surfactant protein C mutation associated with interstitial lung disease 1930–4.
disrupts lung development in transgenic mice. J Biol Chem, 2003;278(52): 101. Pan, S., Huang, L., McPherson, J. et al. Single nucleotide polymorphism‐
52739–46. mediated translational suppression of endoplasmic reticulum mannosidase
80. Young, L.R., Gulleman, P.M., Bridges, J.P. et  al. The alveolar epithelium I modifies the onset of end‐stage liver disease in alpha1‐antitrypsin defi-
determines susceptibility to lung fibrosis in Hermansky–Pudlak syndrome. ciency. Hepatology, 2009;50(1):275–81.
Am J Respir Crit Care Med, 2012;186(10):1014–24. 102. Pan, S., Wang, S., Utama, B. et al. Golgi localization of ERManI defines
81. Bhuiyan, M.S., Pattison, J.S., Osinska, H. et al. Enhanced autophagy amelio- spatial separation of the mammalian glycoprotein quality control system.
rates cardiac proteinopathy. J Clin Invest, 2013;123(12):5284–97. Mol Biol Cell, 2011;22(16):2810–22.
658 THE LIVER:  REFERENCES

103. Iannotti, M.J., Figard, L., Sokac, A.M., and Sifers, R.N. A Golgi‐localized tion of hepatic disease in alpha‐1‐anti‐trypsin deficiency. EMBO Mol Med,
mannosidase (MAN1B1) plays a non‐enzymatic gatekeeper role in protein 2013;5(3):397–412.
biosynthetic quality control. J Biol Chem, 2014;289(17):11844–58. 114. Shen, S., Sanchez, M.E., Blomenkamp, K. et al. Amelioration of Alpha‐1
104. Chappell, S., Guetta‐Baranes, T., Hadzic, N., Stockley, R., and Kalsheker, antitrypsin deficiency diseases with genome editing in transgenic mice.
N. Polymorphism in the endoplasmic reticulum mannosidase I (MAN1B1) Hum Gene Ther, 2018;29(8):861–73.
gene is not associated with liver disease in individuals homozygous for the 115. Song, C.Q., Wang, D., Jiang, T. et  al. In vivo genome editing partially
Z variant of the alpha1‐antitrypsin protease inhibitor (PiZZ individuals). restores alpha1‐antitrypsin in a murine model of AAT deficiency. Hum
Hepatology, 2009;50(4):1315;6. Gene Ther, 2018;29(8):853–60.
105. Joly, P., Lachaux, A., Ruiz, M. et al. SERPINA1 and MAN1B1 polymor- 116. Alam, S., Wang, J., Janciauskiene, S., and Mahadeva, R. Preventing and
phisms are not linked to severe liver disease in a French cohort of alpha‐1 reversing the cellular consequences of Z alpha‐1 antitrypsin accumulation
antitrypsin deficiency children. Liver Int, 2017;37(11):1608–11. by targeting s4A. J Hepatol, 2012;57(1):116–24.
106. Chappell, S., Hadzic, N., Stockley, R., Guetta‐Baranes, T., Morgan, K., and 117. Burrows, J.A., Willis, L.K., and Perlmutter, D.H. Chemical chaperones
Kalsheker, N. A polymorphism of the alpha1‐antitrypsin gene represents a mediate increased secretion of mutant alpha 1‐antitrypsin (alpha 1‐AT) Z:
risk factor for liver disease. Hepatology, 2008;47(1):127–32. A potential pharmacological strategy for prevention of liver injury and
107. Janus, E.D., Phillips, N.T., and Carrell, R.W. Smoking, lung function, and emphysema in alpha 1‐AT deficiency. Proc Natl Acad Sci USA,
alpha 1‐antitrypsin deficiency. Lancet, 1985;1(8421):152–4. 2000;97(4):1796–801.
108. Kemmer, N., Kaiser, T., Zacharias, V., and Neff, G.W. Alpha‐1‐antitrypsin 118. Teckman, J.H. Lack of effect of oral 4‐phenylbutyrate on serum alpha‐1‐
deficiency: outcomes after liver transplantation. Transplant Proc, antitrypsin in patients with alpha‐1‐antitrypsin deficiency: a preliminary
2008;40(5):1492–4. study. J Pediatr Gastroenterol Nutr, 2004;39(1):34–7.
109. Sarkar, S., Perlstein, E.O., Imarisio, S. et  al. Small molecules enhance 119. Bouchecareilh, M., Hutt, D.M., Szajner, P., Flotte, T.R., and Balch, W.E.
autophagy and reduce toxicity in Huntington’s disease models. Nat Chem Histone deacetylase inhibitor (HDACi) suberoylanilide hydroxamic acid
Biol, 2007;3(6):331–8. (SAHA)‐mediated correction of alpha1‐antitrypsin deficiency. J Biol
110. Li, C., Xiao, P., Gray, S.J., Weinberg, M.S., and Samulski, R.J. Combination Chem, 2012;287(45):38265–78.
therapy utilizing shRNA knockdown and an optimized resistant transgene 120. Fox, I.J., Chowdhury, J.R., Kaufman, S.S. et al. Treatment of the Crigler‐
for rescue of diseases caused by misfolded proteins. Proc Natl Acad Sci Najjar syndrome type I with hepatocyte transplantation. N Engl J Med,
USA, 2011;108(34):14258–63. 1998;338(20):1422–6.
111. Mueller, C., Tang, Q., Gruntman, A. et  al. Sustained miRNA‐mediated 121. Yusa, K., Rashid, S.T., Strick‐Marchand, H. et al. Targeted gene correction
knockdown of mutant AAT with simultaneous augmentation of wild‐type of alpha1‐antitrypsin deficiency in induced pluripotent stem cells. Nature,
AAT has minimal effect on global liver miRNA profiles. Mol Ther, 2011;478(7369):391–4.
2012;20(3):590–600. 122. Perlmutter, D.H. and Silverman, G.A. Hepatic fibrosis and carcinogenesis
112. Guo, S., Booten, S.L., Aghajan, M. et al. Antisense oligonucleotide treat- in alpha1‐antitrypsin deficiency: a prototype for chronic tissue damage in
ment ameliorates alpha‐1 antitrypsin‐related liver disease in mice. J Clin gain‐of‐function disorders. Cold Spring Harb Perspect Biol, 2011;3(3).
Invest, 2014;124(1):251–61. 123. Juhasz, G. and Neufeld, T.P. Autophagy: a forty‐year search for a missing
113. Pastore, N., Blomenkamp, K., Annunziata, F. et al. Gene transfer of master membrane source. PLoS Biol, 2006;4(2):e36.
autophagy regulator TFEB results in clearance of toxic protein and correc-
Pathophysiology of Portal
51 Hypertension
Yasuko Iwakiri and Roberto J. Groszmann
Section of Digestive Diseases, Yale School of Medicine, New Haven, CT, USA

INTRODUCTION peripheral resistance. The hyperdynamic circulation along with


increased blood flow to portosystemic collaterals results in
Portal hypertension refers to a pathological increase in blood ­clinically devastating complications such as gastroesophageal
pressure within the portal system due to obstruction of portal varices and variceal hemorrhage, hepatic encephalopathy,
blood flow. The most frequent cause of portal hypertension is ascites, and renal failure due to the hepato‐renal syndrome
liver cirrhosis, and many of lethal complications of cirrhosis [8–10] (Figure 51.1).
such as ascites and gastroesophageal variceal hemorrhage are This chapter provides current knowledge of the biology of
related to portal hypertension. With less frequency, portal hyper- portal hypertension with a focus on cellular and molecular
tension also occurs in noncirrhotic conditions, such as portal events in different segments of vasculatures (intrahepatic versus
vein thrombosis, chronic schistosomiasis, heart failure, Budd– extrahepatic circulations) that contribute to the development
Chiari syndrome, and idiopathic portal hypertension. As a disor- and perpetuation of portal hypertension and the subsequent
der of portal venous pressure, portal hypertension can be development of hyperdynamic circulatory syndrome.
conceptualized using the hydraulic derivation of Ohm’s Law
(pressure = flow × resistance) [1], composed of variables
grounded in basic vascular biology. INTRAHEPATIC CIRCULATION
Portal hypertension develops in liver cirrhosis as a result of
increased intrahepatic vascular resistance caused by multiple
An overview
pathological events in the sinusoidal circulation, such as struc-
tural distortion by fibrosis, microvascular thrombosis, dysfunc- Hepatic sinusoids are specialized vascular beds that constitute
tion of liver sinusoidal endothelial cells (LSECs), and hepatic the hepatic microcirculation. Blockage of sinusoids and the
stellate cell activation [2–5]. LSECs and pericyte‐like hepatic resulting increased hepatic vascular resistance to portal venous
stellate cells are closely associated with one another in these flow is the primary cause of portal hypertension. In cirrhosis,
pathological events in paracrine and autocrine manners. blockage of sinusoids is largely a result of massive structural
Portal hypertension then leads to splanchnic and systemic changes in the liver associated with fibrosis/cirrhosis and intra-
arterial vasodilation, which in turn contributes to increases in hepatic vasoconstriction [5, 11, 12]. Phenotypic changes in
splanchnic blood flow to the portal system and thus further hepatic cells, such as hepatic stellate cells and LSECs, are
increases in portal pressure despite the formation of portosys- known to play pivotal roles in increased intrahepatic vascular
temic collaterals [3–7]. This condition aggravates portal resistance and have been studied extensively. This section
­hypertension and facilitates the hyperdynamic circulation char- addresses the cellular and molecular mechanisms underlying
acterized by decreased mean arterial pressure, decreased sys- increased intrahepatic vascular resistance with a focus on
temic vascular resistance, increased cardiac index, and decreased hepatic stellate cells, LSECs, and thrombosis.

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
660 THE LIVER:  INTRAHEPATIC CIRCULATION

studies demonstrating that activated hepatic stellate cells acquire


a myofibroblast‐like phenotype in response to liver injury [20]
Functional Structural
and exhibit a contractile phenotype [21]. Thus hepatic stellate
Increased resistance
Varices cells, through perivascular contraction in the sinusoidal micro-
circulation, are thought to be key contributors to the dynamic
Porto-systemic
collaterals
and reversible component of portal hypertension in cirrhosis.

Regulation of hepatic stellate cell contraction


PORTAL
HYPERTENSION Endothelin signaling is a key pathway that regulates a contractile
phenotype of hepatic stellate cells. Endothelin binds to G‐protein
coupled receptors, endothelin A (ETA) and endothelin B (ETB),
Vasodilatation Increased flow which are generally found on vascular smooth muscle cells and
endothelial cells, respectively. Endothelin‐1 (ET‐1) is the major
subtype found in liver disease and is preferentially bound to ETA
Hypotension Increase in CO
more than the other two subtypes (ET‐2 and ET‐3) [22]. ET‐1
Central hypovolemia protein levels are elevated in liver injury along with ET‐1 mRNA
[23]. While endothelial cells produce the majority of ET‐1 in
normal liver, liver injury shifts this production primarily to
Activation NE, VP, A-II Na and water retention
increase venous return hepatic stellate cells [24], which also profoundly upregulate ETA
to the heart and ETB receptors [25, 26], suggesting increased sensitivity to
this signal. ET‐1 has been shown to induce contraction of the
Figure 51.1  Summary of the pathophysiology of portal hyperten-
sinusoidal vasculature in an experimental model [27], and antag-
sion. The increase in hepatic resistance leads to an increase in portal
pressure. This leads to a cascade of disturbances in the splanchnic and onism of ETA has been shown to reduce portal pressure in an
systemic circulation characterized by vasodilation, sodium and water animal model of cirrhosis [28]. Smooth muscle cell or hepatic
retention, and plasma volume expansion, which are major players in the stellate cell contraction is a result of myosin light chain (MLC).
pathogenesis of ascites and hepatorenal syndrome. Additionally, these A study on the mechanism of ET‐1‐mediated hepatic stellate cell
alterations lead to an increase in portal blood inflow that contributes to contraction has revealed that ET‐1 leads to MLC phosphoryla-
maintaining and aggravating portal hypertension. Another characteris- tion through multiple pathways [29]. One is a Ca2+‐dependent
tic feature is the development of portosystemic collaterals, which are
responsible for complications such as variceal bleeding and hepatic
pathway in which ET‐1 causes a transient increase in intracellu-
encephalopathy. CO, cardiac output; NE, norepinephrine; VP, vasopres- lar Ca2+, leading to MLC kinase activation and MLC phospho-
sin; A‐II, angiotensin II; Na, sodium. rylation. Others include activation of the protein kinase C and
Rho‐kinase pathways, both leading to MLC phosphorylation.
Hepatic stellate cell biology
Fibrosis and architectural alterations Liver sinusoidal endothelial cell biology
Hepatic fibrosis is thought to be the primary factor that contrib- Fenestration and capillarization
utes to increased intrahepatic vascular resistance in cirrhotic
LSECs have a distinct phenotype from endothelial cells else-
liver. In a classic study, Bhathal and Grossman demonstrated
where in the liver, as well as elsewhere in the body. Their most
through vasodilator challenges in the isolated perfused cirrhotic
distinguishing feature is fenestration with fenestrae being
rat liver that 80% of the increased intrahepatic resistance to por-
approximately 0.1 microns in size and organized into groups of
tal venous flow is attributable to structural changes, while the
sieve plates. It is thought that fenestrae facilitate the transport of
remaining 20% is due to a reversible, hypercontractile pheno-
macromolecules from hepatic sinusoids to the space of Disse,
type of the hepatic microcirculation [13]. The key pathway of
where they can interact with hepatic stellate cells and hepato-
fibrosis in the liver is proinflammatory signaling that causes
cytes. Another feature of LSECs distinct from endothelial cells
activation of hepatic stellate cells and thereby leads to extracel-
in other organs is the lack of a basement membrane, which
lular matrix deposition. Understanding this pathway has led in
allows maximum permeability between the lumen of the sinu-
turn to an understanding of its reversal and the regression of
soid and the space of Disse [30].
fibrosis, which holds a great therapeutic potential for chronic
LSECs lose their fenestrae, develop a basement membrane,
liver disease and portal hypertension [14, 15].
and become “capillarized” as a consequence of liver fibrosis [31,
32]. Vascular endothelial growth factor (VEGF) is known as a key
Hepatic stellate cells as pericytes
factor for the maintenance of endothelial fenestrae [33]. Inhibition
Hepatic stellate cells are thought to play an additional role in of VEGF signaling in a transgenic animal model, in which liver‐
portal hypertension beyond fibrosis. Hepatic stellate cells are specific secretion of a soluble VEGF decoy receptor sequesters
located in the space of Disse, directly underneath LSECs, and endogenous VEGF, caused loss of LSEC fenestration and resulted
work as hepatic pericytes, perivascular non‐endothelial cells in portal hypertension and hepatic stellate cell activation inde-
with a variety of functions including regulation of vascular tone pendent of hepatic parenchymal damage, with reversal of portal
through smooth muscle‐like contractility and regulation of hypertension with restoration of VEGF [34]. This VEGF action
endothelial proliferation [16–19]. These notions arise from on LSEC phenotype is nitric oxide (NO)‐dependent. Accordingly,
51:  Pathophysiology of Portal Hypertension 661

an NO synthase inhibitor, Nω‐Nitro‐L‐arginine methyl ester stellate cell migration [42], which is closely connected to their
hydrochloride (L‐NAME), can lead to the loss of the LSEC phe- activated phenotype.
notype [35]. Nitric oxide produced by LSECs not only regulates vascular
Besides VEGF, various factors have been shown to alter tone in the liver, but also plays a key role in the crosstalk between
LSEC fenestration. Collagens in the space of Disse may play a LSECs and hepatic stellate cells. LSECs are able to induce
role in maintenance or loss of LSEC fenestration [36]. A role of hepatic stellate cell reversion from activation to quiescence via
lipid rafts in regulation of fenestration has also been ­investigated. an NO‐dependent mechanism [43], possibly paracrine signaling
Using super‐resolution fluorescence microscopy, researchers via the Kruppel‐like factor 2 (KLF2)‐NO‐guanylate cyclase
showed an inverse relationship between areas of lipid rafts and pathway in LSECs [44]. Another study showed that NO donors
areas of membranes with fenestration in LSECs. They also dem- were able to inhibit proliferation and chemotaxis of activated
onstrated that inhibiting lipid raft formation via 7‐ketocholesterol hepatic stellate cells in response to platelet‐derived growth fac-
or actin disruption increased fenestration, and conversely that tor by disrupting its intracellular signaling pathway in prosta-
increasing lipid raft formation with a low concentration of glandin E2‐mediated manners [45]. Further, NO inhibited
Triton X‐100 decreased fenestration [37]. hepatic stellate cell migration via cGMP‐dependent protein
kinase (PKG)‐mediated inhibition of the Rac1 pathway [46,
Crosstalk between LSECs and hepatic stellate cells 47]. NO signaling has also been shown to induce hepatic stellate
cell apoptosis [48] via a caspase‐independent mechanism pos-
While phenotypic changes of LSEC is known as a consequence
sibly related to increased mitochondrial oxidative stress and
of liver fibrosis/cirrhosis as mentioned above [31, 32], it is also
increased mitochondrial membrane permeability, along with a
thought that loss of the LSEC phenotype can be permissive for
possible lysosomal stress component. These observations may
hepatic stellate cell activation [38] and that the communication
have relevance in diseases such as alcoholic liver injury and
between LSECs and hepatic stellate cells plays a key role in the
non‐alcoholic fatty liver disease, in which loss of the normal
pathogenesis of portal hypertension (Figure 51.2) [39]. Research
LSEC phenotype has been shown to occur before fibrosis [38].
into LSEC and hepatic stellate cell communication demon-
strated that an isoform of fibronectin produced by LSECs in a
Endothelial dysfunction
bile duct ligation model of liver damage was able to activate
hepatic stellate cells [40], though a subsequent study revealed it NO is a critical regulator of normal hepatic vascular tone and
to be a factor important for hepatic stellate cell motility but not portal pressure [49]. The sources of NO in the hepatic vascula-
differentiation to a myofibroblast phenotype [41]. Another study ture are LSECs and endothelial cells of blood vessels. These
has shown that LSECs communicate with hepatic stellate cells cells constitutively express endothelial nitric oxide synthase
via exosomes containing sphingosine kinase‐1 (SK1) and its (eNOS) and produce a low level of NO. Production of NO
product sphingosine‐1 phosphate, providing a signal for hepatic increases in response to an increase in blood flow via shear

Figure 51.2  LSEC/HSC crosstalk. (a) Injured liver sinusoidal endothelial cells (LSECs) lead to activation of hepatic stellate cells (HSCs). Liver
injury leads LSECs to produce the EIIIA isoform of fibronectin, which signals to HSCs through integrin α9β1 to promote motility, which is impor-
tant for their activated phenotype. LSECs also signal to promote HSC motility through sphingosine kinase 1 – sphingosine‐1‐phosphate (SK1‐S1P)‐
containing exosomes, which adhere to HSCs via fibronectin binding to an integrin receptor. Nitric oxide (NO) production by sinusoidal endothelial
cells is important in maintaining HSC quiescence, with a Kruppel‐like factor (KLF) 2 pathway enhancing NO and guanylate cyclase production.
(b) NO production from LSECs also inhibiting the Rac/Rho pathway in HSCs. NO also causes HSC apoptosis, and may thereby limit the number
of activated HSCs in the liver. Vit A: Vitamin A. Modified from McConnell and Iwakiri [39] and reproduced with permission of Springer Nature.
HSC are in green; healthy LSEC are in blue; injured LSEC are in red.
662 THE LIVER:  EXTRAHEPATIC CIRCULATION

stress [50] and can also be increased by VEGF [51]. LSECs in kinase (i.e. impairment of arterial versus venous specification).
cirrhotic liver express eNOS similarly to LSECs in normal liver. Interestingly, animals with lack of Notch1 gene developed portal
However, eNOS activity is lower under pathologic conditions hypertension even before the onset of nodular regenerative hyper-
and NO release is diminished in disease [52]. In addition, plasia, and this was thought to result from intrahepatic vascular
LSECs in cirrhosis have a reduced ability to respond to increases disorders, possibly due to intussusceptive angiogenesis in the
in blood flow compared to the healthy state [53]. These adverse liver microcirculation [63]. Angiogenesis has also been shown to
changes in LSECs result in decreased production of NO and be associated with fibrosis progression in the liver [64, 65].
impaired vasodilation in the hepatic microcirculation in cirrho- However, this relationship is complex and could be causative or
sis and are important contributors to increased intrahepatic vas- correlative due to involvement of hypoxia, a strong inducer of
cular resistance observed in portal hypertension. The regulation angiogenesis. Modulation of angiogenesis does not generally pro-
of eNOS function involves multiple regulators acting in concert vide a predictable effect on liver fibrosis, either [66].
with stimulatory signaling, including its phosphorylation by the
protein kinase Akt [54], which is facilitated by G‐protein cou- Microvascular thrombosis/platelet activation
pled‐receptor kinase interactor‐1 (GIT1) [55, 56]. The function
of eNOS may also be inhibited by binding to caveolin‐1, which The study of intrahepatic portal hypertension is evolving to
can be disrupted by calmodulin [57]. include platelet activation and thrombosis as crucial factors for
Because endothelial dysfunction leads to increased vascular its pathophysiology. Ian Wanless and others were instrumental
resistance in the sinusoidal microcirculation and promotes acti- contributors in this area, observing what they termed “parenchy-
vation of hepatic stellate cells, a pharmacological approach that mal extinction” accounting for fibrosis progression due to intra-
reverses the dysfunctional LSEC phenotype could be an effec- hepatic vascular thrombosis [67]. Further, while cirrhosis had
tive therapeutic strategy. An example is the use of statins. previously been thought to have a bleeding trend, more advanced
Studies have shown that statins ameliorate portal hypertension physiologic tests to assess coagulation status [68] and system-
in cirrhotic patients [58] as well as experimental models of por- atic studies of bleeding complications [69] have led to an impor-
tal hypertension [59, 60]. These studies demonstrated that tant consensus that hemostatic status in cirrhosis is rebalanced
statins improve endothelial dysfunction and increase NO bioa- or perhaps even prothrombotic.
vailability in the sinusoidal microcirculation. Several potential Platelets are a key player of the typical physiology of vascu-
mechanisms for this increased NO bioavailability have been lar thrombosis. Initially, cirrhotic platelets were thought to be
proposed. One is the ability of statins to inhibit synthesis of iso- dysfunctional and predispose patients to a bleeding trend [70,
prenoids, which are critical for membrane anchoring and activa- 71]. More recently, however, it has been found that activity of
tion of small GTPases, such as RhoA. Given that RhoA/ cirrhotic platelets in hemostasis and thrombosis is potentially
Rho‐kinase signaling could decrease eNOS activity [61] and preserved [72] or even increased [73], although conflicting data
expression [62], statins, by decreasing RhoA activity, could also exist [74]. Generally, the function of platelets in various
improve eNOS function, thereby enhance NO bioavailability types of liver injury is quite complex with multiple stage‐spe-
and decrease intrahepatic vascular resistance [60]. Another cific factors influencing the role of platelets as profibrotic or
potential mechanism is that statins increase activity of Akt/pro- antifibrotic [75]. Further experiments and clinical trials will
tein kinase B, which phosphorylates and activates eNOS, continue to expand our knowledge regarding the role of throm-
thereby increasing NO bioavailability [59]. In addition to bosis and platelets in sinusoidal portal hypertension.
improving endothelial cell dysfunction, statins could target the
RhoA/Rho‐kinase pathway in pericytes (e.g. activated hepatic
stellate cells) and decrease their contractility, thereby lowering EXTRAHEPATIC CIRCULATION
intrahepatic vascular resistance and ameliorating portal hyper-
tension [60].
An overview
In addition to the liver vasculature, the mesenteric vasculature
Pathological angiogenesis plays a key role in portal hypertension. Foundational knowledge
Angiogenesis, or the process of new blood vessel formation from of the pathophysiology of this vascular bed comes from seminal
pre‐existing vascular beds, has also been implicated in portal studies by Groszmann and others [76–96]. These studies dem-
hypertension. Hepatic angiogenesis is thought to generate irregu- onstrated that in portal hypertension, even with increased
lar intrahepatic circulatory routes and thus could increase intrahe- hepatic vascular resistance, the splanchnic circulation is hyper-
patic vascular resistance. The Notch1 signaling pathway is known dynamic. Splanchnic arterial vasodilation is a key feature of the
to be important for embryonic vascular development and postna- hyperdynamic circulation, as it can perpetuate increased blood
tal vascular remodeling. In the postnatal liver, inhibition of the inflow to the portal system and thus exacerbate portal hyperten-
Notch1 pathway leads to nodular regenerative hyperplasia, a sion [4, 5]. Arterial vasodilation is attributable to abnormal cell
common etiology of noncirrhotic portal hypertension. Notch1 function in different layers of the vasculature, such as endothe-
deletion in mice resulted in an abnormal sinusoidal microcircula- lial cells, smooth muscle cells, and the adventitial layer that
tion, as indicated by de‐differentiation of LSECs, pathological contains vascular progenitor cells and neuronal termini. This
remodeling of the hepatic sinusoidal microvasculature, intussus- section discusses the mechanisms of arterial vasodilation and
ceptive angiogenesis (also known as splitting angiogenesis), collateral vessel formation in the splanchnic and systemic circu-
and  dysregulation of ephrinB2/EphB4 and endothelial tyrosine lations in cirrhosis with portal hypertension.
51:  Pathophysiology of Portal Hypertension 663

Arterial vasodilation in the splanchnic vascular relaxation. The candidates of EDHF include ­arachidonic
and systemic circulations acid metabolites, epoxyeicosatrienoic acid (EET), potassium
ions (K+), components of gap junctions, and hydrogen peroxide
Arterial vasodilation
[5]. Vasoactive molecules known to be involved in the regula-
NO is probably the most important vasodilator molecule that tion of vascular tone in cirrhosis are summarized in Figure 51.3.
contributes to excessive vasodilation observed in arteries of the An increase in portal pressure triggers eNOS activation and
splanchnic and systemic circulations in portal hypertension. subsequent NO overproduction in the extrahepatic circulation.
Experimental models of portal hypertension with or without cir- Changes in portal pressure are detected at different vascular
rhosis have also shown that other vasodilator molecules, such as beds depending on the severity of portal hypertension [51]. An
carbon monoxide (CO), prostacyclin (PGI2), endocannabinoids, increment increase in portal pressure is sensed first by the intes-
and endothelium‐derived hyperpolarizing factor (EDHF), are tinal microcirculation and increases VEGF production with a
also induced [5, 7, 97]. The induction of arterial vasodilation subsequent increase in eNOS levels in the intestinal microcircu-
despite inhibition of NO, CO, and PGI2 has indicated the pres- lation. When portal pressure further increases and reaches a
ence of other endothelium‐derived vasodilator molecule(s), ­certain level, arterial vasodilation develops in the splanchnic
known as EDHF [98]. As its name implies, the action of EDHF circulation (i.e. the mesenteric arteries). In contrast to arterial
is to hyperpolarize vascular smooth muscle cells, causing vasodilation in the intestinal microcirculation where VEGF

(a) (b) (c)


Normal Cirrhosis Cirrhosis

Intrahepatic circulation Splanchnic and Systemic Circulation

Sinusoidal Endothelial Cells Endothelial Cells


HbCO
Adrenomedullin Hb
VEGF
TNFα Agonists CO
P1 P1
-1 3 -1 3 Shear Stress
ET Ak ET Ak GTP
t GRK2 t 1 Cyclooxygenase
P v-
BR Akt BR Ca IP3
ET ET 1 AK1
P BH4 BH4 v- eNOS BH4
eNOS CaM eNOS CaM Ca PO4
Ca2+
xxx Hsp90 Hsp90
Hsp90
eNOSCa2+ HO-1 COX
COX-1 COX-1

NO CO PGI2
NO
NO NO TXA2
sGC
ETAR sGC sGC AC
cGMP
cGMP cGMP cAMP
Stellate Cells
Smoooth Muscle Cells

Vasoconstriction Vasodilatation
NO
TXA2
CO
NO PGI2
Endothelial hypoactivation Endothelial hyperactivation

Figure 51.3  Vasoactive molecules known to be involved in the regulation of vascular tone in cirrhosis. In the arterial splanchnic and systemic
circulation (right panel), agonists such as adrenomedullin, vascular endothelial growth factor (VEGF), and tumor necrosis factor alpha (TNFα) or
physical stimuli such as shear stress stimulate Akt, which directly phosphorylates and activates endothelial nitric oxide synthase (eNOS). eNOS
requires cofactors such as tetrahydrobiopterin (BH4) for its activity. Heat shock protein 90 (Hsp90) is one of the positive regulators of eNOS. Like
NO, carbon monoxide (CO) produced by hemeoxygenase‐1 (HO‐1) causes vasodilation by activating soluble guanylate cyclase (sGC) to generate
cyclic guanosine monophosphate (cGMP) in vascular smooth muscle cells. Prostacyclin (PGI2) is synthesized by cyclooxygenase (COX) and
elicits smooth muscle relaxation by stimulating adenylate cyclase (AC) and generation of cyclic adenosine monophosphate (cAMP). In the intrahe-
patic circulation in cirrhosis (left panel), decreased NO and increased thromboxane A2 (TXA2) production in SECs results in a net reduction of
vasorelaxation in the intrahepatic circulation. Endothelin‐1 (ET‐1) has dual vasoactive effects, mediating vasoconstriction through binding to
endothelin A (ETA) receptors located on HSCs and causing HSC contraction. Binding of ET‐1 to ETB receptor (ETBR) mediates vasodilation
through Akt phosphorylation and eNOS phosphorylation in normal liver. In cirrhosis, G‐protein‐coupled receptor kinase‐2 (GRK2), an inhibitor of
G protein‐coupled receptor signaling, is upregulated in SECs, leading to the impairment of Akt phosphorylation and a reduction in NO production.
An increased production of COX‐1‐derived vasoconstrictor prostanoid TXA2 is also an example of endothelial dysfunction in cirrhosis. Modified
from Iwakiri and Groszmann [7] and reproduced with permission of Elsevier.
664 THE LIVER:  HYPERDYNAMIC CIRCULATORY SYNDROME

mediates eNOS upregulation, it is thought that mechanical anti‐VEGF (rapamycin)/anti‐PDGF (Gleevec) [120], anti‐PlGF
forces including cyclic strains and shear stress induce eNOS [119], apelin antagonist [121], sorafenib [122, 123], and a can-
activation and lead to NO overproduction in the splanchnic cir- nabinoid receptor 2 agonist [124]. However, the reduction of
culation [51, 92, 93, 99, 100]. Arterial vasodilation in the these collaterals does not necessarily decrease portal pressure
splanchnic circulation facilitates the systemic circulation to be because it does not substantially change the blood flow to the
also hyperdynamic. portal vein. Therefore, the concomitant mitigation of arterial
vasodilation is needed to reduce portal pressure.
Hypocontractility
A decrease in contractility to vasoconstrictors is also typical of
arteries of the splanchnic and systemic circulations in portal HYPERDYNAMIC CIRCULATORY
hypertension. This hypocontractility occurs largely due to the
presence of excessive vasodilator molecules (e.g. NO) in the
SYNDROME
endothelium, but is to some degree attributable to a decrease in
several vasoconstrictive molecules produced in smooth muscle An overview
cells and neurons. Those molecules include neuropeptide Y Excessive arterial vasodilation in the splanchnic and systemic
[101], urotensin II [102, 103], angiotensin [104], and brady- circulations in portal hypertension results in the development of
kinin [105, 106]. In arteries of the splanchnic circulation, it has hyperdynamic circulatory syndrome. This syndrome is charac-
been shown that vasodilators increase and vasoconstrictors terized by increased cardiac index, decreased systemic vascular
decrease. resistance, and decreased mean arterial pressure and eventually
leads to multiple organ failure frequently observed in chronic
Neural factors liver disease. This section discusses multiple organ failure
Neural factors are also thought to be involved in the dysfunction observed in the hyperdynamic circulatory state (Figure 51.4).
of vascular tone in portal hypertension, especially through the
sympathetic nervous system [101, 107, 108]. It is reported that The systemic circulation and the heart
sympathetic nerve atrophy/regression observed in mesenteric
arterial beds of portal hypertensive rats leads to vasodilation Although arterial vasodilation in the splanchnic circulation is an
and/or hypocontractility of those arterial beds [109, 110]. The essential initiating factor, hyperdynamic circulation does not
role of neural factors in decreased contractile responses remains occur without expansion of plasma volume and the development
to be fully elucidated. of portosystemic collaterals [125, 126]. In cirrhotic patients,
blood and plasma volumes are elevated, but are not evenly dis-
tributed among vascular areas [127]. For example, arterial blood
Structural changes of arteries
volumes in the heart, lungs, and central arterial tree are decreased
The thinning of arterial walls is observed in the splanchnic and compared to those in the splanchnic circulation, resulting in
systemic circulations of rats with cirrhotic livers [111, 112]. central hypovolemia. Decreased blood volumes together with
While this arterial wall thinning can be a consequence of the arterial hypotension lead to baroreceptor activation of potent
hyperdynamic circulation, it may also sustain arterial vasodila- vasoconstriction systems such as the sympathetic nervous sys-
tion and exacerbate portal hypertension [5, 6]. While NO plays tem and the renin angiotensin aldosterone system [128], result-
a role at least in part, the molecular mechanisms responsible for ing in water retention and plasma volume expansion. The
arterial wall thinning remain to be understood. combination of an expanded plasma volume and a reduction in
peripheral vascular resistance leads to an increase in cardiac
output.
Collateral vessel formation When portal hypertension persists, the heart results in a high
Portosystemic collateral vessels develop in response to an cardiac output syndrome: initial compensation occurs according
increase in portal pressure. These collateral vessels develop in to the degree of individual cardiac output, followed by some
an attempt to decompress the hypertensive portal system and are degree of cardiac insufficiency. The cardiac index is usually
formed through the opening of pre‐existing vessels or angiogen- higher than normal (greater than 4 L min−1 m2) but insufficient
esis [113, 114]. However, these collateral vessels are also known to maintain arterial pressure in the face of progressive arterial
to cause serious complications, including variceal bleeding and vasodilation [5]. Importantly, high cardiac output failure is
hepatic encephalopathy [5, 115, 116]. A change in portal pres- reversible once the initial cause leading to the high cardiac out-
sure is thought to be detected first by the intestinal microcircula- put is treated. This reversal has also been observed in patients
tory bed, followed by arteries of the splanchnic circulation [51]. with cirrhosis after liver transplantation [129, 130].
These vascular beds subsequently generate various angiogenic
factors, such as VEGF [96, 117, 118] and placental growth fac-
The hyperdynamic splanchnic circulation
tor (PlGF) [119], which promote the formation of portosystemic
collaterals. As mentioned previously, the hyperdynamic splanchnic circula-
Experimental studies of portal hypertension and cirrhosis tion is central to the development of the syndrome. Although
have shown that portosystemic collaterals are reduced by 18 to it  is commonly recognized as a complication of cirrhosis, it
78% with treatment by anti‐VEGFR2 [100], a combination of should be better conceptualized as a complication of portal
51:  Pathophysiology of Portal Hypertension 665

Splanchnic Renal Systemic Pulmonary Brain


Circulation Circulation Circulation Circulation Circulation

* Acute liver failure

Vasodilation Vasoconstriction Vasodilation Vasodilation Vasodilation


Initiated by local Initiated by local Initiated by local
factors factors factors

• ↑Blood flow • Na/H2O retention • ↑Cardiac output • Arterial hypoxemia • Brain edema
• ↑Portal pressure • Hepatorenal • High output • Hepato-pulmonary
syndrome • Heart failure syndrome
• ↓Peripheral O2
utilization

Figure 51.4  Vasodilation: the source of all evils. *Chronic encephalopathy is associated with a reduced brain blood flow. The mechanism is prob-
ably similar to what is observed in the renal circulation. Modified from Iwakiri and Groszmann [5] and reproduced with permission of John
Wiley & Sons.

hypertension. The term “portal venous inflow” is often used for triggers this syndrome into its full expression is not fully known,
splanchnic blood flow entering into the portal system to distin- local vasodilation mediated by several endothelial vasodilators,
guish it from portal blood flow entering into the liver [77]. including NO and CO, plays an important role [131]. Local fac-
Portal hypertension is the only known pathophysiological situa- tors in the pulmonary circulation may determine why only some
tion in which portal blood flow entering into the portal system patients develop hepatopulmonary syndrome. High cardiac out-
(portal venous inflow) is different from portal blood flow enter- put may also contribute to the severity of hepatopulmonary syn-
ing into the liver. Portal venous inflow entering into the portal drome by increasing shear stress in the pulmonary vascular
system significantly increases, while portal blood flow entering endothelium as well as by shortening a pulmonary and tissue
into the liver decreases, because portal blood escapes into porto- transit time of red blood cells [132, 133].
systemic collaterals formed as a result of portal hypertension
[5]. Arterial vasodilation in the splanchnic vascular bed is
thought to contribute to this increased portal venous inflow
The renal circulation
entering into the portal system and also helps to compensate for The hyperdynamic circulatory state indirectly influences the
the blood that should escape into portosystemic collaterals. renal circulatory bed (Figure 51.4). It is thought that splanchnic
Thus, the most accepted concept is that arterial vasodilation arterial vasodilation in patients with portal hypertension results
starts from the splanchnic vascular bed (i.e. intestinal microcir- in redistribution of the blood volume, leading to a reduction in
culation) and then proceeds to the whole splanchnic circulation) central blood volumes (i.e. a relative hypovolemic state). The
and is the key factor that leads to the development of the hyper- kidney responds to this hypovolemic state by renal arterial vaso-
dynamic circulatory syndrome. The increase in portal pressure constriction, a reduction in glomerular filtration, and retention
itself triggers arterial vasodilation in the splanchnic vascular of sodium and water. The central hypovolemic state activates
bed [51, 93]. signals that induce vasoconstrictive and volume retaining neu-
rohumoral conditions, thereby keeping the sodium and water
retentive state [134–136]. In patients with compensated cirrho-
The hyperdynamic pulmonary circulation sis, progressive systemic arterial vasodilation leads to an eleva-
The hyperdynamic circulation also affects the lungs. The pul- tion of intravascular volume and cardiac output, so that arterial
monary vasodilation is associated with the hepatopulmonary perfusion pressure can be maintained. As disease conditions
syndrome, one of the most severe complications of chronic liver progress, cardiac output continues to increase in response to
disease (Figure  51.4). Although the intrinsic mechanism that progressive arterial vasodilation. Eventually, cardiac response
666 THE LIVER:  REFERENCES

starts to fail to maintain perfusion pressure, renal blood flow practice guidance by the American Association for the study of liver dis-
drops, and renal failure develops [137, 138]. This phenomenon eases. Hepatology, 2017;65(1):310–35.
11. Rockey, D.C. Cell and molecular mechanisms of increased intrahepatic
is known as hepatorenal syndrome. Treatment of arterial vasodi- resistance and hemodynamic correlates. Sanyal AJ SV, editor. Totowa, NJ:
lation improves renal function [139, 140]. Humana Press Inc; 2005.
12. Pinzani, M., Vizzutti, F. Anatomy and vascular biology of the cells in the
portal circulation. Sanyal AJ SV, editor. Totowa, NJ: Humana Press Inc;
The cerebral circulation 2005.
13. Bhathal, P.S. and Grossman, H.J. Reduction of the increased portal vascular
The effects of hyperdynamic circulatory syndrome on the cere-
resistance of the isolated perfused cirrhotic rat liver by vasodilators. J
bral circulation are probably the most difficult to define Hepatol, 1985;1(4):325–37.
(Figure 51.4). Both an increase and decrease in cerebral blood 14. Lee, Y.A., Wallace, M.C., and Friedman, S.L. Pathobiology of liver fibrosis:
flow have been described in acute and chronic liver diseases, a translational success story. Gut, 2015;64(5):830–41.
respectively [141]. An increase in cerebral blood flow has been 15. Trautwein, C., Friedman, S.L., Schuppan, D. et al. Hepatic fibrosis: Concept
to treatment. J Hepatol, 2015;62(1 Suppl):S15–24.
mainly associated with acute liver failure, which could poten-
16. Shepro, D. and Morel, N.M. Pericyte physiology. FASEB, 1993;7(11):
tially lead to the development of brain edema [142]. In contrast, 1031–8.
cerebral blood flow is decreased in chronic liver disease, and 17. Bergers, G. and Song, S. The role of pericytes in blood‐vessel formation and
this decrease runs in parallel with the above‐mentioned reduc- maintenance. Neuro Oncol, 2005;7(4):452–64.
tion in renal blood flow, suggesting that the mechanisms of 18. Franco, M., Roswall, P., Cortez, E. et al. Pericytes promote endothelial cell
survival through induction of autocrine VEGF‐A signaling and Bcl‐w
blood flow reductions in chronic liver disease may be similar in expression. Blood, 2011;118(10):2906–17.
these two organs [143]. 19. LaBarbera, K.E., Hyldahl, R.D., O’Fallon, K.S. et al. Pericyte NF‐kappaB
activation enhances endothelial cell proliferation and proangiogenic cytokine
secretion in vitro. Physiol Rep, 2015;3(4).
20. Rockey, D.C., Boyles, J.K., Gabbiani, G. et al. Rat hepatic lipocytes express
CONCLUSION smooth muscle actin upon activation in vivo and in culture. J Submicrosc
Cytol Pathol, 1992;24(2):193–203.
21. Rockey, D.C., Housset, C.N., and Friedman, S.L. Activation‐dependent con-
The pathogenesis of portal hypertension is complex, because tractility of rat hepatic lipocytes in culture and in vivo. J Clin Invest,
portal hypertension involves not only the hepatic circulation, but 1993;92(4):1795–804.
also the splanchnic and systemic circulations of the hyperdy- 22. Rockey, D.C. Vascular mediators in the injured liver. Hepatology,
namic circulatory state. Because of the disparate conditions of 2003;37(1):4–12.
23. Rockey, D.C., Fouassier, L., Chung, J.J. et  al. Cellular localization of
vascular tone in the intrahepatic and extrahepatic circulations
endothelin‐1 and increased production in liver injury in the rat: potential for
(i.e. vasoconstriction in the intrahepatic circulation versus vaso- autocrine and paracrine effects on stellate cells. Hepatology, 1998;27(2):
dilation in the extrahepatic circulation), the organ/tissue or cell‐ 472–80.
specific modulation of vasodilator or vasoconstrictor molecules 24. Shao, R., Yan, W., and Rockey, D.C. Regulation of endothelin‐1 synthesis by
is of paramount importance for therapeutic purposes. endothelin‐converting enzyme‐1 during wound healing. J Biol Chem,
1999;274(5):3228–34.
25. Rothermund, L., Leggewie, S., Schwarz, A. et al. Regulation of the hepatic
endothelin system in advanced biliary fibrosis in rats. Clin Chem Lab Med,
2000;38(6):507–12.
REFERENCES 26. Yokomori, H., Oda, M., Ogi, M. et al. Enhanced expression of endothelin
receptor subtypes in cirrhotic rat liver. Liver, 2001;21(2):114–22.
1. Iwakiri, Y., Shah, V., and Rockey, D.C. Vascular pathobiology in chronic 27. Zhang, J.X., Pegoli, W., Jr., and Clemens, M.G. Endothelin‐1 induces
liver disease and cirrhosis – current status and future directions. J Hepatol, direct  constriction of hepatic sinusoids. Am J Physiol, 1994;266(4 Pt 1):
2014;61(4):912–24. G624–32.
2. Bosch, J., Groszmann, R.J., and Shah, V.H. Evolution in the understanding 28. Feng, H.Q., Weymouth, N.D., and Rockey, D.C. Endothelin antagonism in
of the pathophysiological basis of portal hypertension: How changes in para- portal hypertensive mice: implications for endothelin receptor‐specific sign-
digm are leading to successful new treatments. J Hepatol, 2015;62(1 aling in liver disease. Am J Physiol Gastrointest Liver Physiol,
Suppl):S121–30. 2009;297(1):G27–33.
3. Iwakiri, Y. The molecules: mechanisms of arterial vasodilatation observed in 29. Iizuka, M., Murata, T., Hori, M. et al. Increased contractility of hepatic stel-
the splanchnic and systemic circulation in portal hypertension. J Clin late cells in cirrhosis is mediated by enhanced Ca2+‐dependent and Ca2+‐
Gastroenterol, 2007;41(3):S288–94. sensitization pathways. Am J Physiol Gastrointest Liver Physiol, 2011;
4. Iwakiri, Y. Pathophysiology of portal hypertension. Clin Liver Dis, 300(6):G1010–21.
2014;18(2):281–91. 30. Wisse, E. An electron microscopic study of the fenestrated endothelial lining
5. Iwakiri, Y. and Groszmann, R.J. The hyperdynamic circulation of chronic of rat liver sinusoids. J Ultrastruct Res, 1970;31(1):125–50.
liver diseases: from the patient to the molecule. Hepatology, 2006;43(2 31. Bhunchet, E. and Fujieda, K. Capillarization and venularization of hepatic
Suppl 1):S121–31. sinusoids in porcine serum‐induced rat liver fibrosis: a mechanism to main-
6. Iwakiri, Y. Endothelial dysfunction in the regulation of cirrhosis and portal tain liver blood flow. Hepatology, 1993;18(6):1450–8.
hypertension. Liver Int, 2012;32(2):199–213. 32. Horn, T., Christoffersen, P., and Henriksen, J.H. Alcoholic liver injury:
7. Iwakiri, Y. and Groszmann, R.J. Vascular endothelial dysfunction in cirrho- defenestration in noncirrhotic livers – a scanning electron microscopic study.
sis. J Hepatol, 2007;46(5):927–34. Hepatology, 1987;7(1):77–82.
8. Tetangco, E.P., Silva, R., and Lerma, E. Portal hypertension: etiology, evalu- 33. Funyu, J., Mochida, S., Inao, M. et al. VEGF can act as vascular permeability
ation, and management. Dis Mon, 2016;62(12):411–26. factor in the hepatic sinusoids through upregulation of porosity of endothe-
9. Cavallin, M., Fasolato, S., Marenco, S. et al. The Treatment of Hepatorenal lial cells. Biochem Biophys Res Commun, 2001;280(2):481–5.
Syndrome. Dig Dis, 2015;33(4):548–54. 34. May, D., Djonov, V., Zamir, G. et  al. A transgenic model for conditional
10. Garcia‐Tsao, G., Abraldes, J.G., Berzigotti, A. et  al. Portal hypertensive induction and rescue of portal hypertension reveals a role of VEGF‐mediated
bleeding in cirrhosis: Risk stratification, diagnosis, and management: 2016 regulation of sinusoidal fenestrations. PloS One, 2011;6(7):e21478.
51:  Pathophysiology of Portal Hypertension 667

35. DeLeve, L.D., Wang, X., Hu, L. et al. Rat liver sinusoidal endothelial cell 58. Abraldes, J.G., Albillos, A., Banares, R. et al. Simvastatin lowers portal pres-
phenotype is maintained by paracrine and autocrine regulation. Am J Physiol sure in patients with cirrhosis and portal hypertension: a randomized con-
Gastrointest Liver Physiol, 2004;287(4):G757–63. trolled trial. Gastroenterology, 2009;136(5):1651–8.
36. McGuire, R.F., Bissell, D.M., Boyles, J. et al. Role of extracellular matrix in 59. Abraldes, J.G., Rodriguez‐Vilarrupla, A., Graupera, M. et  al. Simvastatin
regulating fenestrations of sinusoidal endothelial cells isolated from normal treatment improves liver sinusoidal endothelial dysfunction in CCl4 cirrhotic
rat liver. Hepatology, 1992;15(6):989–97. rats. J Hepatol, 2007;46(6):1040–6.
37. Svistounov, D., Warren, A., McNerney, G.P. et al. The relationship between 60. Trebicka, J., Hennenberg, M., Laleman, W. et al. Atorvastatin lowers portal
fenestrations, sieve plates and rafts in liver sinusoidal endothelial cells. PloS pressure in cirrhotic rats by inhibition of RhoA/Rho‐kinase and activation of
One, 2012;7(9):e46134. endothelial nitric oxide synthase. Hepatology, 2007;46(1):242–53.
38. DeLeve, L.D. Liver sinusoidal endothelial cells in hepatic fibrosis. 61. Ming, X.F., Viswambharan, H., Barandier, C. et al. Rho GTPase/Rho kinase
Hepatology, 2015;61(5):1740–6. negatively regulates endothelial nitric oxide synthase phosphorylation
39. McConnell, M. and Iwakiri, Y. Biology of portal hypertension. Hepatology through the inhibition of protein kinase B/Akt in human endothelial cells.
Int, 2018;12(Suppl 1):11–23. Mol Cell Biol, 2002;22(24):8467–77.
40. Jarnagin, W.R., Rockey, D.C., Koteliansky, V.E. et al. Expression of variant 62. Laufs, U., and Liao, J.K. Post‐transcriptional regulation of endothelial nitric
fibronectins in wound healing: cellular source and biological activity of the oxide synthase mRNA stability by Rho GTPase. J Biol Chem, 1998;
EIIIA segment in rat hepatic fibrogenesis. J Cell Biol, 1994;127(6 Pt 273(37):24266–71.
2):2037–48. 63. Dill, M.T., Rothweiler, S., Djonov, V. et  al. Disruption of Notch1 induces
41. Olsen, A.L., Sackey, B.K., Marcinkiewicz, C. et  al. Fibronectin extra vascular remodeling, intussusceptive angiogenesis, and angiosarcomas in
domain‐A promotes hepatic stellate cell motility but not differentiation into livers of mice. Gastroenterology, 2012;142(4):967–77 e2.
myofibroblasts. Gastroenterology, 2012;142(4):928–37 e3. 64. Corpechot, C., Barbu, V., Wendum, D. et  al. Hypoxia‐induced VEGF and
42. Wang, R., Ding, Q., Yaqoob, U. et al. Exosome adherence and internalization collagen I expressions are associated with angiogenesis and fibrogenesis in
by hepatic stellate cells triggers sphingosine 1‐phosphate‐dependent migra- experimental cirrhosis. Hepatology, 2002;35(5):1010–21.
tion. J Biol Chem, 2015;290(52):30684–96. 65. Ehling, J., Bartneck, M., Wei, X. et  al. CCL2‐dependent infiltrating mac-
43. Deleve, L.D., Wang, X., and Guo, Y. Sinusoidal endothelial cells prevent rat rophages promote angiogenesis in progressive liver fibrosis. Gut,
stellate cell activation and promote reversion to quiescence. Hepatology, 2014;63(12):1960–71.
2008;48(3):920–30. 66. Thabut, D. and Shah, V. Intrahepatic angiogenesis and sinusoidal remodeling
44. Marrone, G., Russo, L., Rosado, E. et  al. The transcription factor KLF2 in chronic liver disease: new targets for the treatment of portal hypertension?
mediates hepatic endothelial protection and paracrine endothelial‐stellate J Hepatol, 2010;53(5):976–80.
cell deactivation induced by statins. J Hepatol, 2013;58(1):98–103. 67. Wanless, I.R., Wong, F., Blendis, L.M. et al. Hepatic and portal vein throm-
45. Failli, P., De, F.R., Caligiuri, A. et  al. Nitrovasodilators inhibit platelet‐ bosis in cirrhosis: possible role in development of parenchymal extinction
derived growth factor‐induced proliferation and migration of activated and portal hypertension. Hepatology, 1995;21(5):1238–47.
human hepatic stellate cells. Gastroenterology, 2000;119(2):479–92. 68. Tripodi, A. Hemostasis abnormalities in cirrhosis. Curr Opin Hematol,
46. Routray, C., Liu, C., Yaqoob, U. et al. Protein kinase G signaling disrupts 2015;22(5):406–12.
Rac1‐dependent focal adhesion assembly in liver specific pericytes. Am J 69. De Pietri, L., Bianchini, M., Montalti, R. et al. Thrombelastography‐guided
Physiol Cell Physiol, 2011;301(1):C66–74. blood product use before invasive procedures in cirrhosis with severe coagu-
47. Lee, J.S., Kang Decker, N., Chatterjee, S. et  al. Mechanisms of nitric oxide lopathy: a randomized, controlled trial. Hepatology, 2016;63(2):566–73.
interplay with Rho GTPase family members in modulation of actin membrane 70. Ordinas, A., Escolar, G., Cirera, I. et  al. Existence of a platelet‐adhesion
dynamics in pericytes and fibroblasts. Am J Pathol, 2005;166(6):1861–70. defect in patients with cirrhosis independent of hematocrit: studies under
48. Langer, D.A., Das, A., Semela, D. et al. Nitric oxide promotes caspase‐inde- flow conditions. Hepatology, 1996;24(5):1137–42.
pendent hepatic stellate cell apoptosis through the generation of reactive 71. Rubin, M.H., Weston, M.J., Langley, P.G. et al. Platelet function in chronic liver
oxygen species. Hepatology, 2008;47(6):1983–93. disease: relationship to disease severity. Dig Dis Sci, 1979;24(3):197–202.
49. Mittal, M.K., Gupta, T.K., Lee, F.Y. et  al. Nitric oxide modulates hepatic 72. Lisman, T., Bongers, T.N., Adelmeijer, J. et  al. Elevated levels of von
vascular tone in normal rat liver. Am J Physiol, 1994;267(3 Pt 1):G416–22. Willebrand Factor in cirrhosis support platelet adhesion despite reduced
50. Shah, V., Haddad, F.G., Garcia‐Cardena, G. et al. Liver sinusoidal endothe- functional capacity. Hepatology, 2006;44(1):53–61.
lial cells are responsible for nitric oxide modulation of resistance in the 73. Raparelli, V., Basili, S., Carnevale, R. et  al. Low‐grade endotoxemia and
hepatic sinusoids. J Clin Invest, 1997;100(11):2923–30. platelet activation in cirrhosis. Hepatology, 2017;65(2):571–81.
51. Abraldes, J.G., Iwakiri, Y., Loureiro‐Silva, M. et al. Mild increases in portal 74. Potze, W., Siddiqui, M.S., Boyett, S.L. et al. Preserved hemostatic status in
pressure upregulate vascular endothelial growth factor and endothelial nitric patients with non‐alcoholic fatty liver disease. J Hepatol, 2016;65(5):
oxide synthase in the intestinal microcirculatory bed, leading to a hyperdy- 980–7.
namic state. Am J Physiol Gastrointest Liver Physiol, 2006;290(5):G980–7. 75. Chauhan, A., Adams, D.H., Watson, S.P. et al. Platelets: no longer bystanders
52. Rockey, D.C. and Chung, J.J. Reduced nitric oxide production by endothelial in liver disease. Hepatology, 2016;64(5):1774–84.
cells in cirrhotic rat liver: endothelial dysfunction in portal hypertension. 76. Chojkier, M. and Groszmann, R.J. Measurement of portal‐systemic shunting
Gastroenterology, 1998;114(2):344–51. in the rat by using gamma‐labeled microspheres. Am J Physiol, 1981;
53. Shah, V., Toruner, M., Haddad, F. et  al. Impaired endothelial nitric oxide 240(5):G371–5.
synthase activity associated with enhanced caveolin binding in experimental 77. Groszmann, R.J., Vorobioff, J., and Riley, E. Splanchnic hemodynamics in
cirrhosis in the rat. Gastroenterology, 1999;117(5):1222–8. portal‐hypertensive rats: measurement with gamma‐labeled microspheres.
54. Fulton, D., Gratton, J.P., McCabe, T.J. et  al. Regulation of endothelium‐ Am J Physiol, 1982;242(2):G156–60.
derived nitric oxide production by the protein kinase Akt. Nature, 78. Vorobioff, J., Bredfeldt, J.E., and Groszmann, R.J. Hyperdynamic circula-
1999;399(6736):597–601. tion in portal‐hypertensive rat model: a primary factor for maintenance of
55. Liu, S., Premont, R.T., and Rockey, D.C. G‐protein‐coupled receptor kinase chronic portal hypertension. Am J Physiol, 1983;244(1):G52–7.
interactor‐1 (GIT1) is a new endothelial nitric‐oxide synthase (eNOS) inter- 79. Sikuler, E., Kravetz, D., and Groszmann, R.J. Evolution of portal hyperten-
actor with functional effects on vascular homeostasis. J Biol Chem, sion and mechanisms involved in its maintenance in a rat model. Am J
2012;287(15):12309–20. Physiol, 1985;248(6 Pt 1):G618–25.
56. Liu, S., Premont, R.T., and Rockey, D.C. Endothelial nitric‐oxide synthase 80. Sarin, S.K., Mosca, P., Sabba, C. et  al. Hyperdynamic circulation in a
(eNOS) is activated through G‐protein‐coupled receptor kinase‐interacting chronic murine schistosomiasis model of portal hypertension. Hepatology,
protein 1 (GIT1) tyrosine phosphorylation and Src protein. J Biol Chem, 1991;13(3):581–4.
2014;289(26):18163–74. 81. Colombato, L.A., Albillos, A., and Groszmann, R.J. Temporal relationship
57. Michel, J.B., Feron, O., Sacks, D. et al. Reciprocal regulation of endothelial of peripheral vasodilatation, plasma volume expansion and the hyperdy-
nitric‐oxide synthase by Ca2+‐calmodulin and caveolin. J Biol Chem, namic circulatory state in portal‐hypertensive rats. Hepatology, 1992;
1997;272(25):15583–6. 15(2):323–8.
668 THE LIVER:  REFERENCES

82. Lee, F.Y., Albillos, A., Colombato, L.A. et al. The role of nitric oxide in the 105. Chu, C.J., Wu, S.L., Lee, F.Y. et al. Splanchnic hyposensitivity to glypres-
vascular hyporesponsiveness to methoxamine in portal hypertensive rats. sin in a haemorrhage/transfused rat model of portal hypertension: role of
Hepatology, 1992;16(4):1043–8. nitric oxide and bradykinin. Clin Sci (Lond), 2000;99(6):475–82.
83. Sieber, C.C. and Groszmann, R.J. Nitric oxide mediates hyporeactivity to 106. Chen, C.T., Chu, C.J., Lee, F.Y. et al. Splanchnic hyposensitivity to glypres-
vasopressors in mesenteric vessels of portal hypertensive rats. sin in a hemorrhage‐transfused common bile duct‐ligated rat model of portal
Gastroenterology, 1992;103(1):235–9. hypertension: role of nitric oxide and bradykinin. Hepatogastroenterology,
84. Lee, F.Y., Colombato, L.A., Albillos, A. et  al. N omega‐nitro‐L‐arginine 2009;56(94–95):1261–7.
administration corrects peripheral vasodilation and systemic capillary 107. Heinemann, A., Wachter, C.H., Fickert, P. et al. Vasopressin reverses mes-
hypotension and ameliorates plasma volume expansion and sodium reten- enteric hyperemia and vasoconstrictor hyporesponsiveness in anesthetized
tion in portal hypertensive rats. Hepatology, 1993;17(1):84–90. portal hypertensive rats. Hepatology, 1998;28(3):646–54.
85. Sieber, C.C., Lopez‐Talavera, J.C., and Groszmann, R.J. Role of nitric 108. Song, D., Liu, H., Sharkey, K.A. et al. Hyperdynamic circulation in portal‐
oxide in the in vitro splanchnic vascular hyporeactivity in ascitic cirrhotic hypertensive rats is dependent on central c‐fos gene expression. Hepatology,
rats. Gastroenterology, 1993;104(6):1750–4. 2002;35(1):159–66.
86. Lopez‐Talavera, J.C., Merrill, W.W., and Groszmann, R.J. Tumor necrosis 109. Coll, M., Martell, M., Raurell, I. et al. Atrophy of mesenteric sympathetic
factor alpha: a major contributor to the hyperdynamic circulation in prehe- innervation may contribute to splanchnic vasodilation in rat portal hyper-
patic portal‐hypertensive rats. Gastroenterology, 1995;108(3):761–7. tension. Liver Int, 2010;30(4):593–602.
87. Sieber, C.C., Lee, F.Y., and Groszmann, R.J. Long‐term octreotide treat- 110. Ezkurdia, N., Coll, M., Raurell, I. et al. Blockage of the afferent sensitive
ment prevents vascular hyporeactivity in portal‐hypertensive rats. pathway prevents sympathetic atrophy and hemodynamic alterations in rat
Hepatology, 1996;23(5):1218–23. portal hypertension. Liver Int, 2012;32(8):1295–305.
88. Shah, V., Wiest, R., Garcia‐Cardena, G. et al. Hsp90 regulation of endothe- 111. Fernandez‐Varo, G., Ros, J., Morales‐Ruiz, M. et al. Nitric oxide synthase
lial nitric oxide synthase contributes to vascular control in portal hyperten- 3‐dependent vascular remodeling and circulatory dysfunction in cirrhosis.
sion. Am J Physiol, 1999;277(2 Pt 1):G463–8. Am J Pathol, 2003;162(6):1985–93.
89. Wiest, R., Das, S., Cadelina, G. et al. Bacterial translocation in cirrhotic 112. Fernandez‐Varo, G., Morales‐Ruiz, M., Ros, J. et al. Impaired extracellular
rats stimulates eNOS‐derived NO production and impairs mesenteric vas- matrix degradation in aortic vessels of cirrhotic rats. J Hepatol,
cular contractility. J Clin Invest, 1999;104(9):1223–33. 2007;46(3):440–6.
90. Wiest, R., Shah, V., Sessa, W.C. et al. NO overproduction by eNOS pre- 113. Sumanovski, L.T., Battegay, E., Stumm, M. et al. Increased angiogenesis in
cedes hyperdynamic splanchnic circulation in portal hypertensive rats. Am portal hypertensive rats: role of nitric oxide. Hepatology, 1999;29(4):
J Physiol, 1999;276(4 Pt 1):G1043–51. 1044–9.
91. Iwakiri, Y., Cadelina, G., Sessa, W.C. et al. Mice with targeted deletion of 114. Sieber, C.C., Sumanovski, L.T., Stumm, M. et al. In vivo angiogenesis in
eNOS develop hyperdynamic circulation associated with portal hyperten- normal and portal hypertensive rats: role of basic fibroblast growth factor
sion. Am J Physiol Gastrointest Liver Physiol, 2002;283(5):G1074–81. and nitric oxide. J Hepatol, 2001;34(5):644–50.
92. Iwakiri, Y., Tsai, M.H., McCabe, T.J. et al. Phosphorylation of eNOS initi- 115. Groszmann, R.J., Kotelanski, B., and Cohn, J.N. Different patterns of
ates excessive NO production in early phases of portal hypertension. Am J porta‐systemic shunting in cirrhosis of the liver studied by an indicator
Physiol Heart Circ Physiol, 2002;282(6):H2084–90. dilution technique. Acta Gastroenterol Latinoam, 1971;3(3):111–6.
93. Tsai, M.H., Iwakiri, Y., Cadelina, G. et al. Mesenteric vasoconstriction trig- 116. Bosch, J., Pizcueta, P., Feu, F. et al. Pathophysiology of portal hyperten-
gers nitric oxide overproduction in the superior mesenteric artery of portal sion. Gastroenterol Clin North Am, 1992;21(1):1–14.
hypertensive rats. Gastroenterology, 2003;125(5):1452–61. 117. Fernandez, M., Vizzutti, F., Garcia‐Pagan, J.C. et al. Anti‐VEGF receptor‐2
94. Wiest, R., Cadelina, G., Milstien, S. et al. Bacterial translocation up‐regu- monoclonal antibody prevents portal‐systemic collateral vessel formation
lates GTP‐cyclohydrolase I in mesenteric vasculature of cirrhotic rats. in portal hypertensive mice. Gastroenterology, 2004;126(3):886–94.
Hepatology, 2003;38(6):1508–15. 118. Geerts, A.M., De Vriese, A.S., Vanheule, E. et al. Increased angiogenesis
95. Groszmann, R.J., Garcia‐Tsao, G., Bosch, J. et al. Beta‐blockers to prevent and permeability in the mesenteric microvasculature of rats with cirrhosis
gastroesophageal varices in patients with cirrhosis. N Engl J Med, and portal hypertension: an in vivo study. Liver Int, 2006;26(7):889–98.
2005;353(21):2254–61. 119. Van Steenkiste, C., Geerts, A., Vanheule, E. et al. Role of placental growth
96. Huang, H.C., Haq, O., Utsumi, T. et al. Intestinal and plasma VEGF levels factor in mesenteric neoangiogenesis in a mouse model of portal hyperten-
in cirrhosis: the role of portal pressure. J Cell Mol Med, 2012;16(5): sion. Gastroenterology, 2009;137(6):2112–24 e1–6.
1125–33. 120. Fernandez, M., Mejias, M., Garcia‐Pras, E. et al. Reversal of portal hyper-
97. Iwakiri, Y. The molecules: mechanisms of arterial vasodilatation observed tension and hyperdynamic splanchnic circulation by combined vascular
in the splanchnic and systemic circulation in portal hypertension. J Clin endothelial growth factor and platelet‐derived growth factor blockade in
Gastroenterol, 2007;41(10 Suppl 3):S288–94. rats. Hepatology, 2007;46(4):1208–17.
98. Feletou, M. and Vanhoutte, P.M. Endothelium‐dependent hyperpolariza- 121. Tiani, C., Garcia‐Pras, E., Mejias, M. et  al. Apelin signaling modulates
tions: past beliefs and present facts. Ann Med, 2007;39(7):495–516. splanchnic angiogenesis and portosystemic collateral vessel formation in
99. Iwakiri, Y. The systemic and splanchnic circulation, in Chronic Liver rats with portal hypertension. J Hepatol, 2009;50(2):296–305.
Failure, Mechanisms and Management, (eds. Gines, P. et  al.), Humana 122. Mejias, M., Garcia‐Pras, E., Tiani, C. et al. Beneficial effects of sorafenib
Press, New York, 2011, pp. 305–21. on splanchnic, intrahepatic, and portocollateral circulations in portal hyper-
100. Fernandez, M., Mejias, M., Angermayr, B. et al. Inhibition of VEGF recep- tensive and cirrhotic rats. Hepatology, 2009;49(4):1245–56.
tor‐2 decreases the development of hyperdynamic splanchnic circulation 123. Reiberger, T., Angermayr, B., Schwabl, P. et  al. Sorafenib attenuates the
and portal‐systemic collateral vessels in portal hypertensive rats. J Hepatol, portal hypertensive syndrome in partial portal vein ligated rats. J Hepatol,
2005;43(1):98–103. 2009;51(5):865–73.
101. Moleda, L., Trebicka, J., Dietrich, P. et al. Amelioration of portal hyperten- 124. Huang, H.C., Wang, S.S., Hsin, I.F. et al. Cannabinoid receptor 2 agonist
sion and the hyperdynamic circulatory syndrome in cirrhotic rats by ameliorates mesenteric angiogenesis and portosystemic collaterals in cir-
­neuropeptide Y via pronounced splanchnic vasoaction. Gut, 2011;60(8): rhotic rats. Hepatology, 2012;56(1):248–58.
1122–32. 125. Genecin, P., Polio, J., and Groszmann, R.J. Na restriction blunts expansion
102. Trebicka, J., Leifeld, L., Hennenberg, M. et  al. Hemodynamic effects of of plasma volume and ameliorates hyperdynamic circulation in portal hyper-
urotensin II and its specific receptor antagonist palosuran in cirrhotic rats. tension. Am J Physiol Gastrointest Liver Physiol, 1990;259(22):G498–503.
Hepatology, 2008;47(4):1264–76. 126. Morgan, J.S., Groszmann, R.J., Rojkind, M., and Enriquez, R.
103. Kemp, W., Krum, H., Colman, J. et  al. Urotensin II: a novel vasoactive Hemodynamic mechanisms of emerging portal hypertension caused by
mediator linked to chronic liver disease and portal hypertension. Liver Int, schistosomiasis in the hamster. Hepatology, 1990;11(1):98–104.
2007;27(9):1232–9. 127. Moller, S. and Henriksen, J.H. Cardiovascular complications of cirrhosis.
104. Hennenberg, M., Trebicka, J., Kohistani, A.Z. et al. Vascular hyporespon- Gut, 2008;57(2):268–78.
siveness to angiotensin II in rats with CCl(4)‐induced liver cirrhosis. Eur J 128. Schrier, R.W. Water and sodium retention in edematous disorders: role of
Clin Invest, 2009;39(10):906–13. vasopressin and aldosterone. Am J Med, 2006;119(7 Suppl 1):S47–53.
51:  Pathophysiology of Portal Hypertension 669

129. Aller, R., de Luis, D.A., Moreira, V. et al. The effect of liver transplantation 136. Schrier, R.W., Arroyo, V., Bernardi, M. et al. Peripheral arterial vasodila-
on circulating levels of estradiol and progesterone in male patients: tion hypothesis: a proposal for the initiation of renal sodium and water
­parallelism with hepatopulmonary syndrome and systemic hyperdynamic retention in cirrhosis. Hepatology, 1988;8(5):1151–7.
circulation improvement. J Endocrinol Invest, 2001;24(7):503–9. 137. Cohn, J.N. Renal hemodynamic alterations in liver disease. Perspect
130. Angus, P.W., Vaughan, R.B., and Chin‐Dusting, J.P. Responses to endothe- Nephrol Hypertens, 1976;3:255–34.
lin‐1 in patients with advanced cirrhosis before and after liver transplanta- 138. Ruiz‐del‐Arbol, L., Monescillo, A., Arocena, C. et al. Circulatory function
tion. Gut, 2004;53(5):773. and hepatorenal syndrome in cirrhosis. Hepatology, 2005;42(2):439–47.
131. Fallon, M.B. Mechanisms of pulmonary vascular complications of liver 139. Cohn, J.N., Tristani, F.E., and Khatri, I.M. Systemic vasoconstrictor and
disease: hepatopulmonary syndrome. J Clin Gastroenterol, 2005;39(4 renal vasodilator effects of PLV‐2 (octapressin) in man. Circulation,
Suppl 2):S138–42. 1968;38(1):151–7.
132. Katsuta, Y., Honma, H., Zhang, X.J. et al. Pulmonary blood transit time and 140. Lenz, K., Hortnagl, H., Druml, W. et  al. Beneficial effect of 8‐ornithin
impaired arterial oxygenation in patients with chronic liver disease. J vasopressin on renal dysfunction in decompensated cirrhosis. Gut,
Gastroenterol, 2005;40(1):57–63. 1989;30(1):90–6.
133. Agusti, A.G., Roca, J., Bosch, J. et  al. Effects of propranolol on arterial 141. Blei, A.T. Monitoring cerebral blood flow. A useful clinical tool in acute
oxygenation and oxygen transport to tissues in patients with cirrhosis. Am liver failure? Liver Transplantation, 2005;11(11):1320–22.
Rev Respir Dis, 1990;142(2):306–10. 142. Vaquero, J., Chung, C., and Blei, A.T. Cerebral blood flow in acute liver
134. Moller, S., Henriksen, J.H., and Bendtsen, F. Central and noncentral blood failure: a finding in search of a mechanism. Metab Brain Dis,
volumes in cirrhosis: relationship to anthropometrics and gender. Am J 2004;19(3–4):177–94.
Physiol Gastrointest Liver Physiol, 2003;284(6):G970–9. 143. Guevara, M., Bru, C., Gines, P. et al. Increased cerebrovascular resistance
135. Shapiro, M.D., Nicholls, K.M., Groves, B.M. et  al. Interrelationship in cirrhotic patients with ascites. Hepatology, 1998;28(1):39–44.
between cardiac output and vascular resistance as determinants of
effective arterial blood volume in cirrhotic patients. Kidney Int,
­
1985;28(2):206–11.
Non‐alcoholic Fatty Liver
52 Disease: Mechanisms
and Treatment
Yaron Rotman and Devika Kapuria
Liver Energy and Metabolism Section, Liver Diseases Branch, National Institute of Diabetes and Digestive
and Kidney Diseases, National Institutes of Health, Bethesda, MD, USA

SUMMARY EPIDEMIOLOGY
Non‐alcoholic fatty liver disease (NAFLD), the excess hepatic Prevalence of NAFLD
accumulation of fat in the form of triglycerides (TG), has
become a global epidemic. The disease spans across a spectrum NAFLD is the most common liver disorder in the Western world
ranging from simple steatosis to non‐alcoholic steatohepatitis and its prevalence appears to be increasing [4], concordant with the
(NASH), where the fat accumulation is accompanied by inflam- rise in metabolic syndrome, obesity, and type 2 diabetes. Estimates
mation and injury, and to cirrhosis and hepatocellular carci- of disease prevalence vary depending on the method used to detect
noma. NAFLD is associated with other metabolic diseases it. In the NHANES III survey (1988–1994) the prevalence of unex-
including type 2 diabetes and coronary artery disease and there plained aminotransferase elevation, thought to represent NAFLD,
is increasing evidence of a link between NAFLD and several was 5.4% [5] while in the 1999–2002 NHANES, the rate of ele-
cancers. In this chapter, we review the epidemiology of NAFLD, vated transaminases was 9.8% [6]. More recently, Takyar et al. [7]
discuss mechanisms of injury and progression, the natural his- in an analysis of 3160 “healthy” volunteers reported a 27.9% preva-
tory of the disease, and management options for patients. lence of presumed NAFLD, based on elevated transaminases and
body mass index (BMI). However, it is well known that normal
aminotransferases do not preclude the presence of NAFLD [8].
This was demonstrated when the same 1988–1994 NHANES III
DEFINITIONS cohort was assessed by ultrasonography and the prevalence of stea-
tosis was found to be 21.4% [9]. In a later study using ultrasonog-
NAFLD is defined as the presence of excess hepatic TG by raphy between 2007–2010 [10], 46% of 328 subjects had evidence
imaging or histology in the absence of secondary causes of of steatosis. Finally, using 1H‐magnetic resonance spectroscopy
hepatic fat accumulation [1]. A liver TG content greater than (MRS) in the multiethnic Dallas heart study, steatosis was found in
5.5% is considered abnormal, a threshold that was obtained one third of the study cohort [11].
from population imaging studies [2] and is consistent with data There is a wide variation in the global prevalence and report-
from biochemical assays of autopsy samples [3]. There are two ing of NAFLD. A recent meta‐analysis [4] that included
forms of NAFLD, which can only be distinguished histologi- 57 worldwide studies reported a pooled global NAFLD preva-
cally. Non‐alcoholic fatty liver (NAFL) is defined as the pres- lence of 25.24%. The highest prevalence was reported from
ence of steatosis without evidence for hepatocellular damage South America (30.45%) and the Middle East (31.79%). The
(ballooning), whereas the presence of hepatocyte injury (or bal- lowest prevalence was reported in Africa (13.48%). Asia,
looning) is required to diagnose NASH. Both NAFL and NASH Europe, and North America respectively have a pooled NAFLD
may present with or without fibrosis. prevalence of 27.37%, 23.71% and 24.13% [12].

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
52:  Non‐alcoholic Fatty Liver Disease: Mechanisms and Treatment 671

A similar trend of increasing NAFLD prevalence is seen in Race and ethnicity are major risk factors for NAFLD with a
pediatric populations, where prevalence rose from 3.9% in disproportionately higher prevalence in Hispanics and relative
1988–1994 to 10.7% in 2007–2010 [13]. An autopsy series span- protection in African‐Americans compared to non‐Hispanic
ning 1993–2003 estimates the prevalence of pediatric NAFLD as Caucasians [4, 11]. In Asians, NAFLD is being increasingly
9.6%, with a prevalence of 38% in obese children [14]. ­recognized even in subjects with normal body mass indices.

Prevalence of NASH
In contrast to NAFLD, that can be identified reliably through GENETIC EPIDEMIOLOGY
imaging modalities, the diagnosis of NASH remains a histologi-
cal one. Although several non‐invasive tests and formulas were The prevalence and severity of NAFLD have a strong heritable
suggested as a method to differentiate NAFL from NASH component, as seen in twin [23] and family [24] studies, and
(­discussed below), none is sufficiently reliable to allow accurate reflected by the strong racial impact [11]. Genome‐wide asso-
diagnosis. As a liver biopsy is often not feasible or clinically‐ ciation studies (GWAS) and in recent years, sequencing
required in most patient cohorts, estimating the true incidence approaches, identified several single nucleotide polymorphism
of NASH can be complex. In a recent meta‐analysis [4], the (SNPs) that are associated with the disease. Of most interest are
pooled prevalence of NASH among NAFLD patients that under- variants that associate with NAFLD independently of estab-
went a liver biopsy was 59.1%. However, this is likely overesti- lished risk factors such as obesity or diabetes, reflecting a true
mating true prevalence due to selection bias. In a study that hepatic risk.
performed prospective research liver biopsies in all subjects
with NAFLD irrespective of the clinical indication 29.9% of PNPLA3
subjects with NAFLD were found to have NASH [10].
Extrapolating from these data, the overall prevalence of NASH The non‐synonymous SNP rs738409 (c.444C>G, p.I148M) in
is estimated at 1.5–6.5% [4]. Consistent with that, in a large the patatin‐like phospholipase domain containing‐3 (PNPLA3)
cohort of obese subjects who underwent bariatric surgery gene encoding the adiponutrin protein has consistently shown
(expected to be enriched for NAFLD), 7.5% were found to have a strong association with NAFLD. Initially identified to be
biopsy‐proven NASH [15]. associated in the general population with hepatic TG [25], and
liver enzymes [26], PNPLA3‐I148M has been subsequently
associated with histological severity of NAFLD and NASH
NAFLD and the metabolic syndrome and was shown to promote inflammation, injury and fibrosis in
NAFLD often occurs together with obesity, diabetes, and the addition to its impact on liver fat content [27, 28]. The high‐
metabolic syndrome. There is a clear correlation between BMI risk allele is also associated with earlier presentation of dis-
and hepatic TG content [11]. The prevalence of obesity in ease in pediatric patients [27], risk of hepatocellular carcinoma
NAFLD subjects is reported at 37–71% [4, 16] and over 90% of [29], alcoholic liver disease [30], and has also been shown to
severely obese patients undergoing bariatric surgery were found impact other liver diseases, suggesting it affects a common
to have NAFLD. Similarly, up to 33% of subjects with NAFLD injury pathway.
have been found to have type 2 diabetes, whereas 70% of The mechanism by which the I148M mutation affects hepatic
patients with diabetes have NAFLD; patients with both diabetes lipid metabolism is not clear. Adiponutrin is present in the endo-
and NAFLD have an increased risk of NASH, fibrosis [17], and plasmic reticulum and lipid droplet membranes of hepatocytes
hepatocellular carcinoma [18]. Almost half of NAFLD and up to and adipose cells and demonstrates both TG hydrolase and acyl-
71% of all NASH patients have been found to have the meta- transferase activities, although this may not be its physiological
bolic syndrome [4, 19]. function in vivo. The I148M mutation impairs ubiquination,
Given the co‐occurrence of NAFLD and metabolic abnormali- leading to adiponutrin accumulation on hepatic lipid droplets
ties, it is difficult to ascertain from cross‐sectional studies whether [31] and modulation of the droplet TG and phospholipid com-
NAFLD is a consequence of obesity, diabetes, and metabolic syn- position [32]. Beyond its function in hepatocytes adiponutrin
drome, or vice versa. In a longitudinal study [20], subjects with has retinyl‐palmitate lipase activity in stellate cells which may
no NAFLD at baseline were more likely to develop NAFLD after explain its link to hepatic fibrosis [33].
7 years of follow‐up if they had higher BMI, insulin resistance, or
presence of metabolic syndrome at baseline. Similarly, NAFLD
at baseline was associated with incident diabetes after 6 years of
TM6SF2
follow‐up [21]. Thus, a bidirectional association exists between A SNP in the transmembrane 6 superfamily member 2 (TM6SF2)
NAFLD and diabetes and the metabolic syndrome. gene, rs58542926 (c.499A>G, p.E167K), was found to be asso-
ciated with NAFLD [34]. The minor allele, associated with
increased risk of NAFLD, is also associated with decreased
Additional risk factors serum LDL‐cholesterol and TG and with decreased risk for
Longitudinal studies show a distinctly increased incidence of myocardial infarction [35]. Inhibition of TM6SF2 impairs the
NAFLD in males compared to females. In women, the incidence lipidation of nascent very low‐density lipoprotein (VLDL) in
of NAFLD is higher in menopausal and post‐menopausal the liver, resulting in decreased TG secretion and increased cel-
women compared to pre‐menopausal women [20, 22]. lular TG accumulation [36, 37].
672 THE LIVER:  MECHANISMS OF HEPATIC INJURY AND DEVELOPMENT OF STEATOHEPATITIS

HSD17B13 contributes less than 5% of the total TG synthesis [47]. However,


in NAFLD patients, DNL is markedly increased and accounts
Several SNPs in 17‐beta hydroxysteroid dehydrogenase 13 for nearly a quarter of the FA in hepatic TG [48]. The presence
(HSD17B13) show a strong association with NAFLD‐­associated of increased hepatic DNL, an insulin‐driven process, despite
liver injury. A splice‐site SNP, rs72613567, generates alterna- hepatic IR suggests the presence of selective IR, either through
tive loss‐of‐function isoforms that associate with protection impairment of specific signaling pathways [49] or as a response
from hepatocyte injury, inflammation and fibrosis but con- to altered nutrient fluxes [50]. Thus, the liver in NAFLD is
versely with increased steatosis in NAFLD subjects [38, 39]. exposed to increased levels of NEFA derived from dysregulated
Other loss‐of‐function variants in the HSD17B13 gene have adipose tissue lipolysis and hepatic lipogenesis.
been identified, with a similar protective effect on injury
[39, 40]. The gene product is a hepatic lipid‐droplet protein with
retinol dehydrogenase activity in vitro and in vivo [39]. VLDL excretion
Several other genes were associated with histological f­ eatures The excess influx of NEFA to the liver, and resultant excess
of NAFLD, including MBOAT7 and GCKR [41]. In addition, formation of hepatic TG are accompanied by increase in VLDL
epigenetic changes also seem to play a role in disease pathogen- and VLDL‐TG secretion. However, this increase is limited by
esis, with data suggesting that methylation of mitochondrial the ability to secrete ApoB100 and is not sufficient to balance
DNA is associated with histological severity [42]. the rate of intrahepatic TG production [51].

β‐oxidation
PATHOGENESIS OF NAFLD There are conflicting data regarding the role of hepatic mito-
chondrial β‐oxidation in accumulation of liver TG, where some
MECHANISMS OF FAT ACCUMULATION studies suggest an increase in hepatic FA oxidation in individu-
als with NAFLD [52] while other studies show a decrease [53].
The hallmark of NAFLD is the accumulation of TG in hepato- This discrepance may be due, in part, to difference in study
cytes. The building blocks of TG are fatty acids (FA), which can methodologies.
be delivered to the hepatocyte as non‐esterified fatty acids
(NEFA) from adipose tissue lipolysis or as part of chylomicron
TG from a dietary source, and can also be formed within the
hepatocyte through the de novo lipogenesis pathway. Hepatic MECHANISMS OF HEPATIC INJURY
TG and FA can be secreted from the liver in VLDL, be oxidized AND DEVELOPMENT
through β‐oxidation or metabolized to other derivatives. Fatty OF STEATOHEPATITIS
liver reflects an imbalance between these input and output path-
ways (Figure 52.1).
Two‐hit versus multi‐hit hypotheses
Steatosis is common but NASH is only seen in a subset of
Adipose tissue NEFA patients. Similarly, steatosis is relatively easy to generate in
In NAFLD, the majority of FA in hepatic TG are derived from murine models but recapitulating NASH in them has been far
circulating NEFA (59%), while newly‐formed FA contribute more difficult. Thus, the mechanisms that underly the transition
26% and dietary fat 15% [43]. Most circulating NEFA derive from steatosis to steatohepatitis need to be elucidated. Initially a
from adipocyte lipolysis (82% in the fasting state and 62% in two‐hit hypothesis was proposed [54]. The accumulation of
the fed state), suggesting the adipocyte is the main source of hepatic TG was suggested as a first hit, sensitizing the liver to
hepatic FA. Adipose tissue insulin resistance (IR) is broadly further insults; a second hit would be required to drive the injury
found in NAFLD, manifesting as failure to suppress postpran- process, mainly through induction of lipid peroxidation. Potential
dial lipolysis and leads to increased delivery of NEFA to the suggested second hits were medications, gut‐derived endotoxin,
liver [44]. Interestingly, the concentration of intrahepatic TG in iron overload, and various inducers of oxidative stress. In recent
NAFLD [45] is strongly associated with adipose tissue IR but years it has become clear that a two‐hit model is unable to explain
not with hepatic IR. Furthermore, the degree of adipose IR has the complexity of NAFLD and a more complex, “multi‐hit” or
been shown to be worse in NASH compared to NAFLD [46], “continuous‐hit” hypothesis has been formed. The major driving
and in NASH patients to be associated with the fibrosis score. process of hepatocyte injury is thought to be lipotoxicity from
Overall, this demonstrates the important role of adipocyte dys- free FA or their derivatives [55], with subsequent activation of
function and excess delivery of adipocyte‐derived NEFA to the inflammatory responses and fibrosis. Genetic and nutritional fac-
liver in the pathogenesis of hepatic TG accumulation. tors are modulators of this process.

De novo lipogenesis Lipotoxicity


De novo lipogenesis (DNL), the formation of FA from acetyl‐ As previously discussed, there is an increase in net hepatic FA
CoA, is highly regulated and key enzymes in the pathway are flux in NAFLD. Shunting of excess FA to form TG is an impor-
upregulated in NAFLD. In healthy lean individuals, DNL tant defense mechanism against lipotoxicity, which probably
52:  Non‐alcoholic Fatty Liver Disease: Mechanisms and Treatment 673

(a)
VLDL

Acetyl-CoA
Glucose

TG
β-oxidation
FA
Glucose

CM

NEFA

(b)
VLDL

Acetyl-CoA
Glucose

TG
β-oxidation
FA
Glucose

CM

NEFA

Figure 52.1  Mechanisms of hepatic fat accumulation. (a) Fatty acids in the liver reflect a balance between input and disposal. Input pathways
include delivery of non‐esterified fatty acids from adipose tissue and triglycerides from chylomicrons and generation of fatty acids from acetyl‐CoA
through de novo lipogenesis (double arrow). Disposal includes oxidation or generation of triglycerides. Triglycerides can be secreted in VLDL or
hydrolyzed to fatty acids. (b) Adipose tissue insulin resistance in NAFLD causes an increase in delivery of non‐esterified fatty acids to the liver, while
hepatic and muscle insulin resistance cause hyperglycemia and increased availability of acetyl‐CoA. De novo lipogenesis is driven by hyperinsuline-
mia generating a net increase in fatty acids and increased shunting to form triglycerides. Secretion of VLDL particles is increased but not sufficient
to compensate. CM, chylomicrons; FA, fatty acids; NEFA, non‐esterified fatty acids; TG, triglycerides; VLDL, very low‐density lipoprotein.

explains the relatively benign phenotype of NAFL. In fact, TG adaptation is impaired, possibly as a result of lipotoxic injury to
are likely inert and not causing injury by themselves as high- the mitochondria, resulting in increased formation of reactive
lighted by animal studies in which inhibition of TG formation oxygen species (ROS) and generation of oxidative stress [59].
has led to increased liver damage despite the inability to gener- Increased microsomal and peroxisomal oxidation of FA also
ate steatosis [56]. When this mechanism is overwhelmed, excess contributes to the formation of excess ROS [60]. Uncontrolled
FA, especially saturated ones, and downstream metabolites such ROS generate injury to the membranes and DNA and lead to the
as ceramides, diacylglycerols, and lysophosphatidylcholine formation of highly toxic lipid peroxidation products, especially
(LPC) accumulate and initiate a cascade of cell injury and death, malondialdehyde and 4‐hydroxynonenal [61].
inflammation, and resultant fibrosis [55]. Endoplasmic reticulum (ER) stress has been demonstrated in
NAFLD [62], likely due to lipotoxic injury from saturated FA, and
leads to activation of the unfolded protein response. As with mito-
Oxidative and endoplasmic reticulum stress chondria, NASH appears to involve an overwhelming of this adap-
Oxidative stress has been strongly implicated in the pathogenesis tive response, and chronic ER stress by itself induces increased
of NASH [57]. The increased hepatic metabolic load drives an DNL and formation of ROS, forming a vicious cycle [63]. When
adaptive process with increased capacity of hepatic mitochon- overwhelmed, ER stress activates IkB and c‐Jun N‐terminal kinases,
dria for oxidative phosphorylation [58]; however, in NASH this initiating a cascade of proinflammatory and pro‐apoptotic responses.
674 THE LIVER:  NATURAL HISTORY OF NAFLD

Extension of damage beyond the hepatocyte The gut microbiota can affect NAFLD through several mech-
anisms. Disruption or leakage across the gut barrier can lead to
The lipotoxic damage to hepatocytes activates regulated death increased hepatic exposure to lipopolysaccharide (LPS) and
pathways in NASH, predominantly apoptosis [64], with activa- other microbial products that act as PAMPs and activate innate
tion of both intrinsic and extrinsic pathways. In addition, there immune responses in the liver, as previously described [75].
appears to be a population of stressed cells which undergo suble- Metabolites derived from the interaction of the gut microbiota
thal injury [65]; the histological correlate of these cells are the with gut content such as trimethylamine (TMA), a metabolite of
ballooned hepatocytes seen in NASH livers. The apoptotic and choline, short‐chain fatty acids, amino acid metabolites, and
ballooned hepatocytes are releasing a myriad of signals includ- endogenous ethanol, can activate signaling pathways in the liver
ing sonic hedgehog [66] and, in addition, these cells are releasing and modulate NAFLD and NASH. Finally, gut bacteria metabo-
extracellular vesicles [65] containing a variety of signaling mol- lize and modify the components of the bile acid pool; this in turn
ecules including ceramides, CXCL10, TRAIL, miRNA, and can affect the activation of intestinal and hepatic farnesoid X
mitochondrial DNA [67]. These act as chemokines and damage‐ receptor (FXR) and modulate the development of steatosis and
associated molecular patterns (DAMPs), which, together with steatohepatitis [76]. Furthermore, microbial metabolism of bile
pathogen‐associated molecular patterns (PAMPs) derived from acids can modify activation of hepatic NKT cells and their anti-
gut bacteria, are recognized by pattern recognition receptors tumor activity, suggesting a link between the gut microbiome
(PRRs) including toll‐like receptors (TLR) and NOD‐like recep- and risk for liver cancer [77]. However, these mechanistic asso-
tors (NLR). Of the various TLRs and NLRs, the ones mostly ciations have to be confirmed in human studies.
implicated in NASH pathogenesis are TLR4 and the NLRP3
inflammasome, respectively. The DAMP/PAMP‐PRR cascade
leads to activation of cells of the innate immune system [68].
NATURAL HISTORY OF NAFLD
Immune cell activation
NAFLD is a slowly progressing disease, and will affect patients
Kupffer cells (KC), the liver‐resident macrophages, are impli- over a span of decades, if not through their entire lifetime. Although
cated in the initiation of the inflammatory process in NASH. often asymptomatic, NAFLD is associated with increased overall
Recognition of PAMP/DAMP by PRRs, as well as a direct effect and liver‐associated mortality.
of free fatty acids leads to KC activation and secretion of
cytokines such as tumor necrosis factor (TNF) and interleukin‐1β
and chemokines such as C‐C motif ligand 2 (CCL2) and CCL5.
Hepatic outcomes of NAFLD
These, in turn, lead to recruitment and activation of circulating Symptoms of NAFLD, if any, are subtle and nonspecific and
monocyte‐derived macrophages to the liver, and amplification may include hepatic discomfort and fatigue. More serious
of the inflammatory process [69]. Furthermore, KC themselves symptoms and complications, such as hepatic decompensation,
are major producers of ROS, further propagating the damage need for liver transplant, and liver‐related mortality, are typically
cascade. Neutrophils are also recruited to the liver by cytokines associated with progression to cirrhosis [78]. As such, the
and DAMPs/PAMPs and further contribute to the inflammatory ­presence of fibrosis is the main determinant of eventual hepatic
process through their myeloperoxidase activity [70]. Targeted outcomes of NAFLD [79–81]. Fibrosis progresses at a variable
depletion of KC in mouse models or genetic disruption of TLR4, rate and can even regress spontaneously [82, 83], likely reflect-
the CCL2‐CCR2 axis or neutrophil enzymes, all lead to amelio- ing the highly dynamic and fluctuating nature of the underlying
ration of experimental NASH, confirming the important role of disease. In paired‐biopsy series, the most important predictor of
these innate immune cells [69]. There is also emerging data progression of fibrosis is the presence of NASH or inflamma-
regarding the role of additional immune cells, including natural tion [83, 84]. Thus, the pathophysiological paradigm of injury
killer (NK) and NKT cells, regulatory T cells, and mucosal and inflammation as drivers of fibrosis is consistent with the
associated invariant T (MAIT) cells [71]. clinical data.
Beyond progression to hepatic decompensation, patients with
NAFLD can also develop hepatocellular carcinoma (HCC) [85].
Microbiome As with other liver diseases, most cases of HCC arise in the
In recent years, multiple studies have demonstrated a difference context of cirrhosis; however, there is now clear evidence that
in the composition of gut microbial populations between sub- NAFLD can lead to HCC in the absence of underlying cirrhosis
jects with NAFLD and controls, and between subjects with and that the risk of noncirrhotic HCC appears to be higher in
NASH and those with NAFL. However, findings have not been NAFLD compared to viral hepatitis [86].
consistent across studies, limiting the ability to generalize those
findings and utilize them to establish a plausible mechanism
[72]. Stronger evidence supporting the role of dysbiosis in
Extrahepatic manifestations
NAFLD comes from animal studies. In a pivotal experiment, the Although the presence of NAFLD (and especially NASH) is
development of fatty liver on a steatogenic diet in germ‐free accompanied by an increase in liver‐related mortality, cardio-
mice was modulated by the microbiome implanted into these vascular disease and cancer are still the first and second leading
mice [73] and steatosis could be transmitted by co‐housing of causes of death in these patients. Thus, there is a clear associa-
mice with different microbiota [74]. tion of NAFLD with extrahepatic manifestations.
52:  Non‐alcoholic Fatty Liver Disease: Mechanisms and Treatment 675

Cardiovascular disease Currently, guidelines recommend considering a liver biopsy in


patients who are at an increased risk of having advanced fibro-
The presence of fatty liver disease is associated with increased
sis, as well as in patients who are suspected to have competing
cardiovascular risk. This increased risk spans the spectrum of
etiologies for hepatic steatosis, or to identify the chief cause of
cardiovascular disease (CVD) from asymptomatic atherosclero-
liver damage in case of more than one liver disease [1]. Thus,
sis [87] to increased cardiovascular events [88] and mortality
there is an unmet need for non‐invasive tools to identify and
[81, 89]. Given that NAFLD and CVD share risk factors such as
stratify NAFLD.
obesity and insulin resistance, it is still not clear whether the
presence of NAFLD confers an increased risk beyond those risk
factors, and whether there is a direct mechanistic effect of Non‐invasive assessments
NAFLD on atherosclerosis development.
Assessment of liver steatosis
Extrahepatic cancers Imaging modalities such as ultrasound or computed tomography
Beyond the risk of HCC, patients with NAFLD also have higher (CT) are reasonably sensitive and specific for the detecting the
rates of extrahepatic malignancies. Several studies have shown presence of moderate–severe steatosis [98]. For example, for
an increased risk of colorectal adenomas and cancer [90], with detection of greater than 20% steatosis, ultrasound has 91% sensi-
more advanced liver disease likely portending a higher risk [91]. tivity and 99% specificity [99] but sensitivity markedly decreases
An increase in breast cancer [92] as well as in other cancers [93] for mild steatosis. The degree of hyperechogenicity on ultrasound
was also reported to be associated with NAFLD. As with cardio- or the relative or absolute hepatic attenuation on CT can be used
vascular disease risk, it is unclear whether the increased cancer for semiquantitative assessment of the degree of steatosis but with
risk associated with NAFLD reflects an association with obesity limited accuracy. Ultrasound remains the primary modality for
and diabetes, both known to increase cancer risk, or whether it diagnosis of hepatic steatosis, due to its wide availability, lack of
is independent of them. radiation exposure, and low cost. Recently, the controlled attenua-
tion parameter (CAP), quantifying the degree of ultrasound atten-
Chronic kidney disease uation in the liver was found to be a sensitive measure for detection
and quantification of steatosis [100]
Chronic kidney disease is more common in NAFLD and the risk In contrast to ultrasound and CT, magnetic resonance imag-
is higher in patients with NASH [94], even after controlling for ing (MRI)‐based imaging modalities offer superior sensitivity
metabolic risk factors. and specificity for detecting steatosis. An additional advantage
of MRI is the ability to accurately quantify liver fat using either
MR spectroscopy or MRI proton density fat fraction (MRI‐
PDFF) [101]. While MRI‐based detection methods have a high
EVALUATION OF THE PATIENT sensitivity and specificity, cost and limited availability currently
WITH NAFLD do not make them a viable option outside of clinical research.
There are several composite scores to calculate the degree of
Clinical features and diagnosis steatosis that are based on serum biomarkers, with or without
anthropometric parameters, such as the fatty liver index, the
NAFLD is diagnosed when there is evidence of hepatic steatosis
hepatic steatosis index, and the proprietary SteatoTest; these
on imaging or histology and there are no secondary causes
have reasonable accuracy but it is unclear whether they can
of  steatosis present [1]. Patients with NAFLD are typically
replace the need for an imaging study in clinical care [102].
asymptomatic, and diagnosis is often incidental. Patients may
occasionally present with fatigue or a vague, nondescript pain or
discomfort in the right upper quadrant and may have hepato- Identification of NASH
megaly due to fatty infiltration of the liver. Signs or symptoms A crucial element in managing patients with NAFLD and in
of chronic liver disease are generally absent unless patients clinical trial recruitment is the differentiation of NAFL from
­present with advanced liver disease or cirrhosis. NASH. As detailed above, a liver biopsy is the gold standard for
diagnosis, but poses significant limitations in terms of accepta-
Role of liver biopsy bility, risk, and cost. Several blood‐based markers for the diag-
nosis of NASH have been studied, based on disease pathogenesis.
The liver biopsy is an important tool in the diagnosis and man- Examples are cytokeratin‐18 fragments as markers of hepato-
agement of NAFLD. It is the gold standard and the only reliable cyte apoptosis [103, 104], oxidized lipid products [104], and
method to differentiate NASH from NAFL and also allows for several proprietary panels (reviewed in [102]). However, these
assessment of the multiple components of disease and for accu- all suffer from limited accuracy, incomplete validation, and/or
rately staging the degree of fibrosis. There are several scoring cost and are not yet ready for routine clinical application.
schemes that are utilized for semiquantitative assessment of
NAFLD histology, most commonly the NASH–CRN scoring
Assessment of fibrosis
system [95]. These scores are appropriate for use in clinical tri-
als but should not be used in lieu of an appropriate histological Since fibrosis is the main determinant of mortality and liver‐related
diagnosis [96]. Despite its usefulness, the use of a liver biopsy is complications in NAFLD, a non‐invasive marker of fibrosis would
limited by its inherent risk [97], cost and sampling error. provide risk stratification and allow for selection of patients for
676 THE LIVER:  TREATMENT OPTIONS IN NAFLD

treatment. As with assessment of steatosis, investigated biomark- Other blood‐based biomarkers are more specific and are
ers include imaging modalities and blood‐based markers [105]. based on measuring markers of the hepatic fibrotic process and
Liver fibrosis is associated with increased hepatic stiffness. deposition of extracellular matrix. Blood levels of hyaluronic
Several imaging modalities measure liver elasticity, which is acid, TIMP1, and pro‐collagen III amino terminal peptide
directly related to stiffness, by generating a shear wave in the (PIIINP and Pro‐C3) are among the promising ones. Several
liver and measuring its velocity. Vibration controlled transient compound and proprietary scores are based on these markers,
elastography (Fibroscan), shear‐wave elastography, and acoustic including the enhanced liver fibrosis (ELF), the FibroSure/
radial force imaging (ARFI) are based on ultrasound, while MRI FibroTest, and FibroMeter NAFLD. In general, these specific
is used for magnetic resonance elastography (MRE). In general, scores provide greater accuracy than the nonspecific ones, but
elastography‐based tests provide high accuracy in detecting cir- are currently limited by availability and cost.
rhosis and advanced fibrosis but are not sufficiently accurate to
diagnose the presence of earlier fibrosis (stage 2) [102, 106].
Several blood‐based biomarkers or compound scores were
developed to detect fibrosis. Some of these are nonspecific but TREATMENT OPTIONS IN NAFLD
employ readily available data such as liver enzymes, platelet
count, and clinical and anthropometric features. These include Despite the rise in prevalence and impact of NAFLD and NASH,
scores initially applied to viral hepatitis, such as the AST/ALT there are currently no approved therapies for the disease.
ratio, AST/platelet ratio index (APRI) or FIB‐4, as well as However, several medications approved for other indications
scores developed specifically for NAFLD such as the NAFLD have shown benefit in treating NASH and a plethora of new
fibrosis score (NFS) or BARD score. The accuracy of all of pharmacological agents are being studied. With the increasing
these scores is limited and despite easy availability, they are not understanding of pathogenic processes, there is a plethora of
sufficient for routine clinical use, though they are useful in epi- agents developed to target these processes (Figure 52.2) [107].
demiological studies [105]. Reviewed here are currently available treatments with proven

Figure 52.2  Mechanistic targets for NAFLD/NASH treatment. Increased availability of carbohydrates and non‐esterified fatty acids, together
with insulin resistance, drive hepatic accumulation of triglycerides and fatty acids, leading to metabolic and oxidative stress, hepatocyte injury,
activation of an inflammatory response and eventually fibrosis. Lifestyle modification and bariatric surgery decrease the excessive caloric load to
the liver. GLP‐1 receptor agonists (such as liraglutide) decrease caloric load (through reduction in appetite and weight loss) and improve insulin
sensitivity with possible direct effect on the liver. PPAR agonists (like pioglitazone or elafibranor) improve insulin resistance in adipose tissue and/
or muscle, reducing the excess load of glucose and fatty acids to the liver, while also decreasing hepatic lipogenesis. Agonists of FXR (such as
obeticholic acid) or FGF‐19 analogs (NGM‐282) decrease bile acid synthesis and hepatic lipogenesis; NGM‐282 may also act as an insulin sensi-
tizer. Inhibitors of acetyl‐coA carboxylase, the first step in hepatic de novo lipogenesis, block hepatic fatty acid synthesis and accumulation.
Downstream of hepatic lipid metabolism, vitamin E, a fat‐soluble antioxidant, ameliorates the oxidative stress that results from lipotoxicity while
ASK1 inhibitors (such selonsertib) block the ASK1‐JNK pathway through which oxidative stress leads to hepatocyte injury and apoptosis.
Cenicriviroc, a C‐C chemokine receptor type 2 (CCR2) and 5 (CCR5) antagonist, blocks the recruitment of inflammatory cells and decreases
inflammatory injury. Finally, anti‐fibrotic agents target the fibrotic process itself. ACCi, acetyl‐coA carboxylase inhibitors; ASK1i, apoptosis sig-
nal‐regulating kinase 1 inhibitors; CCR, C‐C chemokine receptor; CHO, carbohydrates; FA, fatty acids; FGF, fibroblast growth factor; FXR,
farnesoid X receptor; GLP‐1RA, glucagon‐like peptide 1 receptor agonists, NEFA, non‐esterified fatty acids; PPAR, peroxisome proliferator‐acti-
vator receptor; TG, triglycerides.
52:  Non‐alcoholic Fatty Liver Disease: Mechanisms and Treatment 677

Table 52.1  Select response rates in clinical trials of NASH


Histological Resolution of
Treatment Duration responsea (%) NASH (%) Improvement in fibrosis (%) References
Lifestyle (all) 12 months 47 25 19b [109]
Lifestyle (Subjects 12 months 82 58 26b [109]
with ≥ 5% weight loss) (Subjects with ≥ 10% weight
loss had 45% fibrosis regression)
Pioglitazone 18 months –  34b–58 47–51 44b [112, 113]
96 weeks
Vitamin E 96 weeks 43 36 41b [112]
Obeticholic acid 72 weeks 45 20b 35 [115]
Elafibranor (120 mg) 52 weeks 48 19 Not reported [114]
Liraglutide 48 weeks 43 35 26b [124]
a
 Histological response criteria varied between trials but typically include at least 2‐point reduction in the NAFLD activity score (NAS). Response percentages calculated by
intention‐to‐treat.
b
 Non‐significant compared to placebo.

benefit, and those that are in advanced clinical trial stages at the ligands. PPARγ is expressed predominantly in the adipose tissue
time of writing (Table 52.1). and controls adipogenesis, lipogenesis, and glucose metabolism.
Thiazolidinediones (TZDs) are synthetic agonists of PPARγ and
are approved for the treatment of diabetes. Pioglitazone, the TZD
Reduction of caloric intake
most commonly used, was studied for the treatment of NASH in
Treatment trials of lifestyle intervention, including diet and several trials. In the PIVENS study, 163 non‐diabetics with
physical activity, have shown not only reduction in liver TG NASH were treated with pioglitazone or placebo for 96 weeks.
content but also histological improvement, directly related to Histological improvement was seen in 34% of subjects treated
the degree of weight loss [108]. Recently Vilar‐Gomez and oth- with pioglitazone, compared to 19% in placebo, but the difference
ers [109] reported on the results of a 52‐week lifestyle interven- did not reach the pre‐specified statistical significance level [112].
tion program in 293 subjects with NASH. Only 30% of In contrast, in a study including 101 diabetics or pre‐diabetics
participants achieved greater than 5% weight loss, highlighting with NASH, 18 months of pioglitazone treatment were associated
the difficulty of achieving a sustained effect, the main limitation with resolution of NASH in 51% of subjects, which was signifi-
of lifestyle interventions. However, among those who lost cantly greater than the placebo‐treated arm [113]. Overall, piogl-
greater than 5% of their weight, resolution of NASH was seen in itazone treatment appears to be effective for NASH, especially in
58%. Furthermore, in the few subjects who lost greater than subjects with diabetes or pre‐diabetes, but its acceptability may
10% of their weight, resolution of NASH was seen in 90% and be limited by concerns about cardiovascular risks and by weight
a reduction in fibrosis stage was seen in 81%. Another evidence, gain, which is a common side‐effect.
albeit indirect, for the benefit of weight reduction is the marked Elafibrinor is a dual PPAR agonist, targeting both PPARα and
rate of histological response in the control arms of therapeutic PPARδ. The phase IIb GOLDEN‐505 trial compared two doses
trials, which is typically associated with losing weight [110]. of elafibrinor (80 mg and 120 mg) for 52 weeks to placebo.
Given the proven effects of lifestyle modification for NASH and Treatment with elafibranor was associated with histological
its known benefits for associated comorbidities such as diabetes improvement, but only in a post hoc analysis in subjects with
and metabolic syndrome, it should be advocated in any patient significant disease activity at baseline [114]. A phase III trial
with NAFLD, irrespective of pharmacological therapy options. is underway.
Due to the difficulty of sustaining weight loss with lifestyle
interventions alone, bariatric surgery is useful for patients with Bile acid signaling
obesity and has shown benefit in treating diabetes. In the largest
FXR acts as an intracellular bile acid sensor and when acti-
series to date [15], histological resolution of NASH was seen in
vated, inhibits bile acid synthesis and secretion as well as
85% of subjects one year after surgery, and 33% showed
decreasing lipogenesis. Obeticholic acid (OCA) is a synthetic
improvement in steatosis. Recently, 5‐year follow‐up data was
bile acid that acts as an FXR agonist. The phase IIb FLINT trial
reported from the same cohort [111]. The 85% NASH resolu-
compared OCA with placebo in patients with NASH; OCA was
tion rate persisted and fibrosis continued to improve between
superior to placebo in achieving histological improvement
years 1 and 5. Thus, bariatric surgery may be an attractive ther-
(46%) and NASH resolution (22%) [115] but was associated
apy for a subset of patients with NASH,
with significant pruritus and a mild increase in LDL‐­cholesterol.
A phase III trial to determine efficacy and safety is ongoing and
Pharmacological treatment other FXR agonists are in earlier stages of development.
Fibroblast growth factor 19 (FGF19) is a peptide hormone that
PPAR activation
is released in response to FXR activation in the terminal ileum
The peroxisome proliferator‐activator receptors (PPARs) are a and controls hepatocyte bile acid synthesis through its receptor
group of nuclear receptors that transcriptionally regulate a wide FGFR4. FGF19 also has insulin‐like effects on hepatic gluco-
variety of metabolic and inflammatory processes. Their activation neogenesis and glycogen synthesis. Phase II studies in patients
is typically driven by binding various FA and FA derivatives as with NAFLD using NGM282, an FGF19 analog, demonstrate
678 THE LIVER:  REFERENCES

efficacy in reducing hepatic fat content [116] and improving hypertriglyceridemia was seen in 16% of treated patients. It is
NASH histology [117]. still unknown whether GS‐0976 treatment can also lead to his-
tological improvement in NASH.
Vitamin E
Other treatments
Given the evidence for the role of oxidative stress in the patho-
genesis of NASH, antioxidants could have therapeutic benefit. There is a large number of investigative agents currently studied
Vitamin E (α‐tocopherol), a lipid‐soluble antioxidant, has been as potential treatments for NASH, targeting various aspects of
studied in several clinical trials. In the previously mentioned the disease. Among those are agonists directed at the thyroid
PIVENS trial, 43% of non‐diabetic patients with NASH treated hormone receptor β, inhibitors of apoptosis signal‐regulating
with 800 IU/day of natural vitamin E had histological improve- kinase 1 (ASK‐1), a key mediator of hepatocyte stress response,
ment and 36% had resolution of NASH [112] and similar results modulators of gut microbiome, combined agonists of receptors
were seen in the TONIC pediatric trial [118]. Furthermore, of all to gut‐derived hormones, an analogue of FGF21, and antifi-
pharmacological therapies, vitamin E is the only one shown to brotic agents. In addition, there are agents in early stages of
decrease the rate of clinically meaningful outcomes such as development targeting genes that are associated with the dis-
decompensation or need for liver transplant, albeit in a retro- ease. Finally, combination treatment with agents targeting dif-
spective analysis [119]. However, there have been reports of ferent pathways may be needed to achieve meaningful responses
increased risk of hemorrhagic stroke [120] and prostate cancer in a large proportion of subjects.
[121] in subjects receiving high‐dose vitamin E, and a recent
report from a small trial suggests an increase in gastrointestinal
bleeding in NAFLD patients, especially in diabetics [122].
Thus, using vitamin E to treat NASH can be considered but CONCLUSION
requires monitoring and assessment of potential risks.
NAFLD has risen to epidemic proportions and directly or indi-
rectly affects the health of many. Increased understanding of
GLP 1 receptor agonists
genetic and pathogenic mechanisms will hopefully lead to
GLP‐1 (glucagon‐like peptide‐1) is an incretin hormone secreted effective and safe therapeutic options.
by enteroendocrine L cells in response to ingested nutrients. It
acts predominantly in the fed‐state, modulating and augmenting
the insulin response. GLP‐1 potentiates glucose‐stimulated
insulin secretion from beta cells, mediated by activation of
REFERENCES
its  G  protein‐coupled receptor. GLP‐1 receptor agonists   1. Chalasani, N., Younossi, Z., Lavine, J.E. et al. The diagnosis and manage-
(GLP‐1RA) are in clinical use for the treatment of diabetes and ment of nonalcoholic fatty liver disease: practice guidance from the American
obesity and are thought to mediate their effects through a com- Association for the Study of Liver Diseases. Hepatology, 2018; 67:328–57.
bination of the insulinotropic effect, weight loss, increased  2. Szczepaniak, L.S., Nurenberg, P., Leonard, D. et  al. Magnetic resonance
energy expenditure, and delayed gastric emptying. spectroscopy to measure hepatic triglyceride content: prevalence of hepatic
steatosis in the general population. Am J Physiol Endocrinol Metab, 2005;
In diabetes‐associated NAFLD, GLP‐1RAs lead to reduction 288:E462–8.
in liver enzymes and liver fat content in human and animal stud-  3. Kwiterovich, P.O., Jr., Sloan, H.R., and Fredrickson, D.S. Glycolipids and
ies [123]. In the phase II randomized controlled LEAN trial other lipid constituents of normal human liver. J Lipid Res, 1970; 11:322–30.
[124], 52 non‐diabetics with NASH were treated with liraglu-   4. Younossi, Z., Anstee, Q.M., Marietti, M. et al. Global burden of NAFLD and
NASH: trends, predictions, risk factors and prevention. Nat Rev Gastroenterol
tide, a GLP‐1RA, or placebo for 48 weeks. Liraglutide provided
Hepatol, 2018; 15:11–20.
histological benefit including resolution of NASH in 39%.   5. Clark, J.M., Brancati, F.L., and Diehl, A.M. The prevalence and etiology of
Larger trials with other GLP‐1RA formulations are ongoing. elevated aminotransferase levels in the United States. Am J Gastroenterol,
Given the combined benefit on insulin resistance, weight, liver 2003; 98:960–7.
histology, and beneficial cardiovascular outcome in diabetics,   6. Ioannou, G.N., Boyko, E.J., and Lee, S.P. The prevalence and predictors of
elevated serum aminotransferase activity in the United States in 1999–2002.
GLP‐1RAs are an attractive option to explore further in the
Am J Gastroenterol, 2006; 101:76–82.
management of NAFLD.   7. Takyar, V., Nath, A., Beri, A. et  al. How healthy are the “Healthy volun-
teers”? Penetrance of NAFLD in the biomedical research volunteer pool.
Inhibition of de novo lipogenesis Hepatology, 2017; 66:825–33.
  8. Fracanzani, A.L., Valenti, L., Bugianesi, E. et al. Risk of severe liver disease
As detailed above, DNL is increased in the liver of subjects in nonalcoholic fatty liver disease with normal aminotransferase levels: a
with NAFLD and DNL‐derived fatty acids may be drivers of role for insulin resistance and diabetes. Hepatology, 2008; 48:792–8.
  9. Lazo, M., Hernaez, R., Eberhardt, M.S. et al. Prevalence of nonalcoholic fatty
hepatic lipotoxicity. GS‐0976 is a liver‐targeted inhibitor of liver disease in the United States: the Third National Health and Nutrition
acetyl‐CoA carboxylase (ACC), the rate‐limiting enzyme in Examination Survey, 1988–1994. Am J Epidemiol, 2013; 178:38–45.
DNL. In a prospective randomized phase 2 study, 100 subjects 10. Williams, C.D., Stengel, J., Asike, M.I. et  al. Prevalence of nonalcoholic
with likely NASH received GS‐0976 (5 or 20 mg/d) for fatty liver disease and nonalcoholic steatohepatitis among a largely middle‐
12  weeks and were compared with 26 subjects who received aged population utilizing ultrasound and liver biopsy: a prospective study.
Gastroenterology, 2011; 140:124–31.
­placebo [125]. GS‐0976 treatment was significantly associated 11. Browning, J.D., Szczepaniak, L.S., Dobbins, R. et al. Prevalence of hepatic
with a dose‐dependent and marked reduction in liver fat steatosis in an urban population in the United States: Impact of ethnicity.
­content, and was generally well‐tolerated, although marked Hepatology, 2004; 40:1387–95.
52:  Non‐alcoholic Fatty Liver Disease: Mechanisms and Treatment 679

12. Younossi, Z.M., Koenig, A.B., Abdelatif, D. et al. Global epidemiology of 37. Smagris, E., Gilyard, S., Basuray, S. et al. Inactivation of Tm6sf2, a gene
nonalcoholic fatty liver disease  –  meta‐analytic assessment of prevalence, defective in fatty liver disease, impairs lipidation but not secretion of very
incidence, and outcomes. Hepatology, 2016; 64:73–84. low density lipoproteins. J Biol Chem, 2016; 291:10659–76.
13. Welsh, J.A., Karpen, S., and Vos, M.B. Increasing prevalence of nonalco- 38. Abul‐Husn, N.S., Cheng, X., Li, A.H. et al. A protein‐truncating HSD17B13
holic fatty liver disease among United States adolescents, 1988–1994 to variant and protection from chronic liver disease. New Eng J Med, 2018;
2007–2010. J Pediatr, 2013; 162:496–500. 378:1096–106.
14. Schwimmer, J.B., Deutsch, R., Kahen, T. et al. Prevalence of fatty liver in 39. Ma, Y., Belyaeva, O.V., Brown, P.M. et al. HSD17B13 is a hepatic retinol
children and adolescents. Pediatrics, 2006; 118:1388–93. dehydrogenase associated with histological features of non‐alcoholic fatty
15. Lassailly, G., Caiazzo, R., Buob, D. et al. Bariatric surgery reduces features liver disease. Hepatology, 2019;69(4):1504–19.
of nonalcoholic steatohepatitis in morbidly obese patients. Gastroenterology, 40. Kozlitina, J., Stender, S., Hobbs, H.H., and Cohen, J.C. HSD17B13 and chronic
2015; 149:379–88. liver disease in Blacks and Hispanics. N Engl J Med, 2018; 379:1876–7.
16. Sasaki, A., Nitta, H., Otsuka, K. et al. Bariatric surgery and non‐alcoholic 41. Eslam, M., Valenti, L., and Romeo, S. Genetics and epigenetics of NAFLD
fatty liver disease: current and potential future treatments. Front Endocrinol, and NASH: clinical impact. J Hepatol, 2018; 68:268–79.
2014; 5:164. 42. Pirola, C.J., Gianotti, T.F., Burgueno, A.L. et al. Epigenetic modification of
17. Doycheva, I., Patel, N., Peterson, M., and Loomba, R. Prognostic implica- liver mitochondrial DNA is associated with histological severity of nonalco-
tion of liver histology in patients with nonalcoholic fatty liver disease in holic fatty liver disease. Gut, 2013; 62:1356–63.
diabetes. J Diabetes Complications, 2013; 27:293–300. 43. Donnelly, K.L., Smith, C.I., Schwarzenberg, S.J. et al. Sources of fatty acids
18. Anstee, Q.M., Targher, G., and Day, C.P. Progression of NAFLD to diabetes stored in liver and secreted via lipoproteins in patients with nonalcoholic
mellitus, cardiovascular disease or cirrhosis. Nat Rev Gastroenterol Hepatol, fatty liver disease. J Clin Invest, 2005; 115:1343–51.
2013; 10:330–44. 44. Sanyal, A.J., Campbell‐Sargent, C., Mirshahi, F. et al. Nonalcoholic steato-
19. Marchesini, G., Bugianesi, E., Forlani, G. et  al. Nonalcoholic fatty liver, hepatitis: association of insulin resistance and mitochondrial abnormalities.
steatohepatitis, and the metabolic syndrome. Hepatology, 2003; 37:917–23. Gastroenterology, 2001; 120:1183–92.
20. Zelber‐Sagi, S., Lotan, R., Shlomai, A. et  al. Predictors for incidence and 45. Bril, F., Barb, D., Portillo‐Sanchez, P. et al. Metabolic and histological impli-
remission of NAFLD in the general population during a seven‐year prospec- cations of intrahepatic triglyceride content in nonalcoholic fatty liver dis-
tive follow‐up. J Hepatol, 2012; 56:1145–51. ease. Hepatology, 2017; 65:1132–44.
21. Ma, J., Hwang, S.J., Pedley, A. et al. Bi‐directional analysis between fatty 46. Musso, G., Cassader, M., De Michieli, F. et al. Nonalcoholic steatohepatitis
liver and cardiovascular disease risk factors. J Hepatol, 2017; 66:390–7. versus steatosis: Adipose tissue insulin resistance and dysfunctional response
22. Kojima, S.‐I., Watanabe, N., Numata, M. et al. Increase in the prevalence of to fat ingestion predict liver injury and altered glucose and lipoprotein
fatty liver in Japan over the past 12 years: analysis of clinical background. J metabolism. Hepatology, 2012; 56:933–42.
Gastroenterol, 2003; 38:954–61. 47. Schwarz, J.M., Neese, R.A., Turner, S. et al. Short‐term alterations in carbo-
23. Loomba, R., Schork, N., Chen, C.H. et al. Heritability of Hepatic Fibrosis hydrate energy intake in humans. Striking effects on hepatic glucose produc-
and Steatosis Based on a Prospective Twin Study. Gastroenterology, 2015; tion, de novo lipogenesis, lipolysis, and whole‐body fuel selection. J Clin
149:1784–93. Invest, 1995; 96:2735–43.
24. Schwimmer, J.B., Celedon, M.A., Lavine, J.E. et al. Heritability of nonalco- 48. Lambert, J.E., Ramos‐Roman, M.A., Browning, J.D., and Parks, E.J.
holic fatty liver disease. Gastroenterology, 2009; 136:1585–92. Increased de novo lipogenesis is a distinct characteristic of individuals with
25. Romeo, S., Kozlitina, J., Xing, C. et al. Genetic variation in PNPLA3 confers nonalcoholic fatty liver disease. Gastroenterology, 2014; 146:726–35.
susceptibility to nonalcoholic fatty liver disease. Nature Genet, 2008; 49. Brown, M.S. and Goldstein, J.L. Selective versus total insulin resistance: a
40:1461–5. pathogenic paradox. Cell Metab, 2008; 7:95–6.
26. Yuan, X., Waterworth, D., Perry, J.R.B. et  al. Population‐based genome‐ 50. Otero, Y.F., Stafford, J.M., and Mcguinness, O.P. Pathway‐selective insulin
wide association studies reveal six loci influencing plasma levels of liver resistance and metabolic disease: the importance of nutrient flux. J Biol
enzymes. Am J Hum Genet, 2008; 83:520–8. Chem, 2014; 289:20462–9.
27. Rotman, Y., Koh, C., Zmuda, J.M. et al. The association of genetic variability 51. Fabbrini, E., Mohammed, B.S., Magkos, F. et al. Alterations in adipose tis-
in patatin‐like phospholipase domain‐containing protein 3 (PNPLA3) with sue and hepatic lipid kinetics in obese men and women with nonalcoholic
histological severity of nonalcoholic fatty liver disease. Hepatology, 2010; fatty liver disease. Gastroenterology, 2008; 134:424–31.
52:894–903. 52. Iozzo, P., Bucci, M., Roivainen, A. et al. Fatty acid metabolism in the liver,
28. Speliotes, E.K., Butler, J.L., Palmer, C.D. et al. PNPLA3 variants specifi- measured by positron emission tomography, is increased in obese individu-
cally confer increased risk for histologic nonalcoholic fatty liver disease but als. Gastroenterology, 2010; 139:846–56, 56 e1–6.
not metabolic disease. Hepatology, 2010; 52:904–12. 53. Naguib, G., Morris, N., Haynes‐Williams, V. et  al. Beta‐oxidation is
29. Hassan, M.M., Kaseb, A., Etzel, C.J. et al. Genetic variation in the PNPLA3 decreased in non‐alcoholic fatty liver disease: a non‐invasive assessment uti-
gene and hepatocellular carcinoma in USA: risk and prognosis prediction. lizing palmitate and acetate breath tests. Hepatology, 2017; 66:1043.
Mol Carcinog, 2013; 52: Suppl 1:E139–47. 54. Day, C.P. and James, O.F.W. Steatohepatitis: a tale of two “hits”?
30. Tian, C., Stokowski, R.P., Kershenobich, D. et  al. Variant in PNPLA3 is Gastroenterology, 1998; 114:842–5.
associated with alcoholic liver disease. Nature Genet, 2010; 42:21–3. 55. Neuschwander‐Tetri, B.A. Hepatic lipotoxicity and the pathogenesis of non-
31. Basuray, S., Smagris, E., Cohen, J.C., and Hobbs, H.H. The PNPLA3 variant alcoholic steatohepatitis: the central role of nontriglyceride fatty acid metab-
associated with fatty liver disease (I148M) accumulates on lipid droplets by olites. Hepatology, 2010; 52:774–88.
evading ubiquitylation. Hepatology, 2017; 66:1111–24. 56. Yamaguchi, K., Yang, L., Mccall, S. et  al. Inhibiting triglyceride synthe-
32. Mitsche, M.A., Hobbs, H.H., and Cohen, J.C. Patatin‐like phospholipase sis  improves hepatic steatosis but exacerbates liver damage and fibrosis
domain‐containing protein 3 promotes transfers of essential fatty acids from in  obese mice with nonalcoholic steatohepatitis. Hepatology, 2007;
triglycerides to phospholipids in hepatic lipid droplets. J Biol Chem, 2018; 45:1366–74.
293:6958–68. 57. Videla, L.A., Rodrigo, R., Orellana, M. et al. Oxidative stress‐related param-
33. Pirazzi, C., Valenti, L., Motta, B.M. et al. PNPLA3 has retinyl‐palmitate lipase eters in the liver of non‐alcoholic fatty liver disease patients. Clin Sci (Lond),
activity in human hepatic stellate cells. Hum Mol Genet, 2014; 23:4077–85. 2004; 106:261–8.
34. Kozlitina, J., Smagris, E., Stender, S. et al. Exome‐wide association study 58. Sunny, N.E., Parks, E.J., Browning, J.D., and Burgess, S.C. Excessive
identifies a TM6SF2 variant that confers susceptibility to nonalcoholic fatty hepatic mitochondrial TCA cycle and gluconeogenesis in humans with non-
liver disease. Nat Genet, 2014; 46:352–6. alcoholic fatty liver disease. Cell Metab, 2011; 14:804–10.
35. Holmen, O.L., Zhang, H., Fan, Y. et  al. Systematic evaluation of coding 59. Koliaki, C., Szendroedi, J., Kaul, K. et al. Adaptation of hepatic mitochon-
­variation identifies a candidate causal variant in TM6SF2 influencing total drial function in humans with non‐alcoholic fatty liver is lost in steatohepa-
cholesterol and myocardial infarction risk. Nat Genet, 2014; 46:345–51. titis. Cell Metab, 2015; 21:739–46.
36. Mahdessian, H., Taxiarchis, A., Popov, S. et al. TM6SF2 is a regulator of 60. Bellanti, F., Villani, R., Facciorusso, A. et al. Lipid oxidation products in the
liver fat metabolism influencing triglyceride secretion and hepatic lipid pathogenesis of non‐alcoholic steatohepatitis. Free Radic Biol Med, 2017;
droplet content. Proc Nat Acad Sci USA, 2014; 111:8913–8. 111:173–85.
680 THE LIVER:  REFERENCES

61. Seki, S., Kitada, T., Yamada, T. et al. In situ detection of lipid peroxidation   86. Mittal, S., El‐Serag, H.B., Sada, Y.H. et al. Hepatocellular carcinoma in the
and oxidative DNA damage in non‐alcoholic fatty liver diseases. J Hepatol, absence of cirrhosis in United States veterans is associated with nonalco-
2002; 37:56–62. holic fatty liver disease. Clin Gastroenterol Hepatol, 2016; 14:124–31 e1.
62. Puri, P., Mirshahi, F., Cheung, O. et al. Activation and dysregulation of the   87. Pais, R., Giral, P., Khan, J.F. et al. Fatty liver is an independent predictor of
unfolded protein response in nonalcoholic fatty liver disease. early carotid atherosclerosis. J Hepatol, 2016; 65:95–102.
Gastroenterology, 2008; 134:568–76.   88. Targher, G., Byrne, C.D., Lonardo, A. et al. Non‐alcoholic fatty liver dis-
63. Lebeaupin, C., Vallee, D., Hazari, Y. et al. Endoplasmic reticulum stress sig- ease and risk of incident cardiovascular disease: a meta‐analysis. J Hepatol,
nalling and the pathogenesis of non‐alcoholic fatty liver disease. J Hepatol, 2016; 65:589–600.
2018; 69:927–47.   89. Söderberg, C., Stål, P., Askling, J. et al. Decreased survival of subjects with
64. Feldstein, A.E., Canbay, A., Angulo, P. et al. Hepatocyte apoptosis and fas elevated liver function tests during a 28‐year follow‐up. Hepatology, 2010;
expression are prominent features of human nonalcoholic steatohepatitis. 51:595–602.
Gastroenterology, 2003; 125:437–43.   90. Adams, L.A., Anstee, Q.M., Tilg, H., and Targher, G. Non‐alcoholic fatty
65. Ibrahim, S.H., Hirsova, P., and Gores, G.J. Non‐alcoholic steatohepatitis liver disease and its relationship with cardiovascular disease and other
pathogenesis: sublethal hepatocyte injury as a driver of liver inflammation. extrahepatic diseases. Gut, 2017; 66:1138–53.
Gut, 2018; 67:963–72.   91. Ahn, J.S., Sinn, D.H., Min, Y.W. et al. Non‐alcoholic fatty liver diseases and
66. Rangwala, F., Guy, C.D., Lu, J. et al. Increased production of sonic hedge- risk of colorectal neoplasia. Aliment Pharmacol Ther, 2017; 45:345–53.
hog by ballooned hepatocytes. J Pathol, 2011; 224:401–10.   92. Kim, G.A., Lee, H.C., Choe, J. et al. Association between non‐alcoholic
67. Garcia‐Martinez, I., Santoro, N., Chen, Y. et al. Hepatocyte mitochondrial fatty liver disease and cancer incidence rate. J Hepatol.2017;
DNA drives nonalcoholic steatohepatitis by activation of TLR9. J Clin  93. Hicks, S.B., Mara, K., Larson, J.J. et  al. The incidence of extrahepatic
Invest, 2016; 126:859–64. malignancies in nonalcoholic fatty liver disease (NAFLD). Hepatology,
68. Arrese, M., Cabrera, D., Kalergis, A.M., and Feldstein, A.E. Innate immu- 2018; 68:20A–A.
nity and inflammation in NAFLD/NASH. Dig Dis Sci, 2016; 61:1294–303.   94. Musso, G., Gambino, R., Tabibian, J.H. et al. Association of non‐alcoholic
69. Schuster, S., Cabrera, D., Arrese, M., and Feldstein, A.E. Triggering and fatty liver disease with chronic kidney disease: a systematic review and
resolution of inflammation in NASH. Nat Rev Gastroenterol Hepatol, 2018; meta‐analysis. PLoS Med, 2014; 11:e1001680.
15:349–64.   95. Kleiner, D.E., Brunt, E.M., Van Natta, M. et al. Design and validation of a
70. Xu, R., Huang, H., Zhang, Z., and Wang, F.S. The role of neutrophils in the histological scoring system for nonalcoholic fatty liver disease. Hepatology,
development of liver diseases. Cell Mol Immunol, 2014; 11:224–31. 2005; 41:1313–21.
71. Byun, J.S. and Yi, H.S. Hepatic immune microenvironment in alcoholic and   96. Brunt, E.M., Kleiner, D.E., Wilson, L.A. et al. Nonalcoholic fatty liver dis-
nonalcoholic liver disease. Biomed Res Int, 2017; 2017:6862439. ease (NAFLD) activity score and the histopathologic diagnosis in NAFLD:
72. Sharpton, S.R., Ajmera, V., and Loomba, R. Emerging role of the gut micro- distinct clinicopathologic meanings. Hepatology, 2011; 53:810–20.
biome in nonalcoholic fatty liver disease: from composition to function. Clin   97. Takyar, V., Etzion, O., Heller, T. et al. Complications of percutaneous liver
Gastroenterol Hepatol.2019;17(2):296–306. biopsy with Klatskin needles: a 36‐year single‐centre experience. Aliment
73. Le Roy, T., Llopis, M., Lepage, P. et  al. Intestinal microbiota determines Pharmacol Ther, 2017; 45:744–53.
development of non‐alcoholic fatty liver disease in mice. Gut, 2013;  98. Zhang, Y.N., Fowler, K.J., Hamilton, G. et  al. Liver fat imaging‐a clinical
62:1787–94. overview of ultrasound, CT, and MR imaging. Br J Radiol, 2018; 91:20170959.
74. Henao‐Mejia, J., Elinav, E., Jin, C. et al. Inflammasome‐mediated dysbiosis   99. Hernaez, R., Lazo, M., Bonekamp, S. et al. Diagnostic accuracy and relia-
regulates progression of NAFLD and obesity. Nature, 2012; 482:179–85. bility of ultrasonography for the detection of fatty liver: a meta‐analysis.
75. Chu, H., Duan, Y., Yang, L., and Schnabl, B. Small metabolites, possible big Hepatology, 2011; 54:1082–90.
changes: a microbiota‐centered view of non‐alcoholic fatty liver disease. 100. Karlas, T., Petroff, D., Sasso, M. et al. Individual patient data meta‐analysis
Gut, 2019;68(2):359–70. of controlled attenuation parameter (CAP) technology for assessing steato-
76. Jiang, C., Xie, C., Li, F. et al. Intestinal farnesoid X receptor signaling pro- sis. J Hepatol, 2017; 66:1022–30.
motes nonalcoholic fatty liver disease. J Clin Invest, 2015; 125:386–402. 101. Noureddin, M., Lam, J., Peterson, M.R. et al. Utility of magnetic resonance
77. Ma, C., Han, M., Heinrich, B. et  al. Gut microbiome‐mediated bile acid imaging versus histology for quantifying changes in liver fat in nonalcoholic
metabolism regulates liver cancer via NKT cells. Science, 2018; 360:(6391). fatty liver disease trials. Hepatology (Baltimore, Md), 2013; 58:1930–40.
78. Vilar‐Gomez, E., Calzadilla‐Bertot, L., Wai‐Sun Wong, V. et al. Fibrosis sever- 102. Wong, V.W., Adams, L.A., De Ledinghen, V. et al. Noninvasive biomarkers
ity as a determinant of cause‐specific mortality in patients with advanced non- in NAFLD and NASH ‐ current progress and future promise. Nat Rev
alcoholic fatty liver disease: a multi‐national cohort study. Gastroenterology, Gastroenterol Hepatol, 2018; 15:461–78.
2018; 155:443–57 e17. 103. Kwok, R., Tse, Y.K., Wong, G.L. et al. Systematic review with meta‐analy-
79. Angulo, P., Kleiner, D.E., Dam‐Larsen, S. et al. Liver fibrosis, but no other sis: non‐invasive assessment of non‐alcoholic fatty liver disease‐‐the role of
histologic features, is associated with long‐term outcomes of patients with transient elastography and plasma cytokeratin‐18 fragments. Aliment
nonalcoholic fatty liver disease. Gastroenterology, 2015; 149:389–97. Pharmacol Ther, 2014; 39:254–69.
80. Dulai, P.S., Singh, S., Patel, J. et al. Increased risk of mortality by fibrosis 104. Feldstein, A.E., Lopez, R., Tamimi, T.A. et al. Mass spectrometric profiling
stage in nonalcoholic fatty liver disease: systematic review and meta‐analy- of oxidized lipid products in human nonalcoholic fatty liver disease and
sis. Hepatology, 2017; 65:1557–65. nonalcoholic steatohepatitis. J Lipid Res, 2010; 51:3046–54.
81. Ekstedt, M., Hagström, H., Nasr, P. et al. Fibrosis stage is the strongest pre- 105. Younossi, Z.M., Loomba, R., Anstee, Q.M. et al. Diagnostic modalities for
dictor for disease‐specific mortality in NAFLD after up to 33 years of fol- nonalcoholic fatty liver disease, nonalcoholic steatohepatitis, and associ-
low‐up. Hepatology, 2015; 61:1547–54. ated fibrosis. Hepatology, 2018; 68:349–60.
82. Ekstedt, M., Franzén, L.E., Mathiesen, U.L. et al. Long‐term follow‐up of 106. Hannah, W.N., Jr. and Harrison, S.A. Noninvasive imaging methods to
patients with NAFLD and elevated liver enzymes. Hepatology, 2006; determine severity of nonalcoholic fatty liver disease and nonalcoholic
44:865–73. steatohepatitis. Hepatology, 2016; 64:2234–43.
83. Singh, S., Allen, A.M., Wang, Z. et al. Fibrosis progression in nonalcoholic 107. Rotman, Y. and Sanyal, A.J. Current and upcoming pharmacotherapy for
fatty liver vs nonalcoholic steatohepatitis: a systematic review and meta‐­ non‐alcoholic fatty liver disease. Gut, 2017; 66:180–90.
analysis of paired‐biopsy studies. Clin Gastroenterol Hepatol, 2015; 13: 108. Promrat, K., Kleiner, D.E., Niemeier, H.M. et al. Randomized controlled
643–54.e1–9; quiz e39–40. trial testing the effects of weight loss on nonalcoholic steatohepatitis.
84. Argo, C.K., Northup, P.G., Al‐Osaimi, A.M.S., and Caldwell, S.H. Hepatology, 2010; 51:121–9.
Systematic review of risk factors for fibrosis progression in non‐alcoholic 109. Vilar‐Gomez, E., Martinez‐Perez, Y., Calzadilla‐Bertot, L. et  al. Weight
steatohepatitis. J Hepatol, 2009; 51:371–9. loss through lifestyle modification significantly reduces features of nonal-
85. Younossi, Z.M., Otgonsuren, M., Henry, L. et  al. Association of nonalco- coholic steatohepatitis. Gastroenterology, 2015; 149:367–78.e5.
holic fatty liver disease (NAFLD) with hepatocellular carcinoma (HCC) in 110. Han, M.a.T., Altayar, O., Hamdeh, S. et al. Rates of and factors associated
the United States from 2004 to 2009. Hepatology, 2015; 62:1723–30. with placebo response in trials of pharmacotherapies for nonalcoholic
52:  Non‐alcoholic Fatty Liver Disease: Mechanisms and Treatment 681

s­teatohepatitis: systematic review and meta‐analysis. Clin Gastroenterol 118. Lavine, J.E., Schwimmer, J.B., Van Natta, M.L. et al. Effect of vitamin E or
Hepatol, 2010;17(4):616–29. metformin for treatment of nonalcoholic fatty liver disease in children and
111. Lassailly, G., Caiazzo, R., Gnemmi, V. et al. Regression of fibrosis after adolescents. JAMA, 2011; 305:1659.
disappearance of nash in morbidly obese patients: a prospective bariatric 119. Vilar‐Gomez, E., Vuppalanchi, R., Gawrieh, S. et al. Vitamin E improves
surgery cohort with sequential liver biopsies. Hepatology, 2018; 68:44a‐5a. transplant‐free survival and hepatic decompensation among patients with
112. Sanyal, A.J., Chalasani, N., Kowdley, K.V. et al. Pioglitazone, vitamin E, or NASH and advanced fibrosis. Hepatology.2018.
placebo for nonalcoholic steatohepatitis. New Eng J Med, 2010; 362:1675–85. 120. Schürks, M., Glynn, R.J., Rist, P.M. et al. Effects of vitamin E on stroke
113. Cusi, K., Orsak, B., Bril, F. et  al. Long‐term pioglitazone treatment for subtypes: meta‐analysis of randomised controlled trials. BMJ, 2010;
patients with nonalcoholic steatohepatitis and prediabetes or type 2 diabe- 341:c5702.
tes mellitus. Ann Int Med, 2016; 165:305. 121. Klein, E.A., Thompson, I.M., Jr., Tangen, C.M. et al. Vitamin E and the risk
114. Ratziu, V., Harrison, S.A., Francque, S. et al. Elafibranor, an agonist of the of prostate cancer: the selenium and vitamin E cancer prevention trial
peroxisome proliferator—activated receptor—α and −δ, induces resolution (SELECT). JAMA, 2011; 306:1549–56.
of nonalcoholic steatohepatitis without fibrosis worsening. Gastroenterology, 122. Rotman, Y., Lingala, S., and Morris, N. Subtle iron deficiency is common
2016; 150:1147–59.s during vitamin E treatment for NAFLD. Hepatology, 2016; 64:571a‐2a.
115. Neuschwander‐Tetri, B.A., Loomba, R., Sanyal, A.J. et  al. Farnesoid X 123. Shao, N., Kuang, H.Y., Hao, M. et al. Benefits of exenatide on obesity and
nuclear receptor ligand obeticholic acid for non‐cirrhotic, non‐alcoholic non‐alcoholic fatty liver disease with elevated liver enzymes in patients
steatohepatitis (FLINT): a multicentre, randomised, placebo‐controlled with type 2 diabetes. Diabetes Metab Res Rev, 2014; 30:521–9.
trial. Lancet, 2015; 385:956–65. 124. Armstrong, M.J., Gaunt, P., Aithal, G.P. et al. Liraglutide safety and effi-
116. Harrison, S.A., Rinella, M.E., Abdelmalek, M.F. et al. NGM282 for treat- cacy in patients with non‐alcoholic steatohepatitis (LEAN): a multicentre,
ment of non‐alcoholic steatohepatitis: a multicentre, randomised, double‐ double‐blind, randomised, placebo‐controlled phase 2 study. Lancet, 2016;
blind, placebo‐controlled, phase 2 trial. Lancet, 2018; 391:1174–85. 387:679–90.
117. Harrison, S.A., Trotter, J.F., Paredes, A.H. et al. NGM282 rapidly improves 125. Loomba, R., Kayali, Z., Noureddin, M. et al. GS‐0976 reduces hepatic stea-
NAFLD activity score (NAS) and fibrosis in 12 weeks in patients with tosis and fibrosis markers in patients with nonalcoholic fatty liver disease.
biopsy‐confirmed nonalcoholic steatohepatitis (NASH): results of a phase Gastroenterology, 2018; 155:1463–73 e6.
2 multi‐center dose finding study. Hepatology, 2018; 68:66a‐a.
Alcoholic Liver Disease
53 Bin Gao1, Xiaogang Xiang1,2, Lorenzo Leggio3, and George F. Koob4
1
Laboratory of Liver Diseases, National Institute on Alcohol Abuse and Alcoholism, National Institutes of
Health, Bethesda, MD, USA
2
Department of Infectious Diseases, Ruijin Hospital, School of Medicine, Shanghai Jiao Tong University,
Shanghai, China
3
Section on Clinical Psychoneuroendocrinology and Neuropsychopharmacology, National Institute on
Alcohol Abuse and Alcoholism and National Institute on Drug Abuse, National Institutes of Health,
Bethesda, MD, USA
4
National Institute on Alcohol Abuse and Alcoholism and National Institute on Drug Abuse, National
Institutes of Health, Bethesda, MD, USA

Excessive alcohol consumption leads a spectrum of liver disor- oxygen species (ROS) and free radicals that induce lipid peroxi-
ders from simple steatosis (fatty liver) to severe forms of liver dation to promote adduct formation with several macromole-
injury including steatohepatitis, cirrhosis, and hepatocellular cules. Finally, another enzyme that can oxidize ethanol is
carcinoma [1, 2]. Many proinflammatory mediators, metabolic catalase, which is located in the peroxisome. However, this
pathways, transcriptional factors, and epigenetic factors have pathway is considered a minor pathway of ethanol oxidation
been identified to contribute to the development and progression in the liver.
of alcoholic liver disease (ALD), which were discussed in detail The second step of ethanol metabolism is to convert acetal-
in the previous edition of this book. In this chapter, we briefly dehyde into acetate by aldehyde dehydrogenase (ALDH) [3].
discuss ethanol metabolism, ALD pathogenesis, clinical diag- ALDH also exists in multiple forms with ALDH2 being a key
nosis and treatment of ALD, and in more detail recent advances isoform to metabolize acetaldehyde. ALDH2 is expressed at
in the pathogenesis and treatment of ALD. We also discuss the the highest levels in the liver and is located in the mitochon-
neurobiology of alcohol use disorder (AUD) and treatment of dria. Acetate generated from acetaldehyde metabolism in
this disorder in patients with ALD. hepatocytes is rapidly secreted into the circulation and is fur-
ther converted into acetyl coenzyme A (acetyl‐CoA) by acetyl‐
CoA synthetases. Acetyl‐CoA then enters into the citric acid
cycle (Krebs cycle) and is finally converted to carbon dioxide
ETHANOL METABOLISM and water.
Human ADH and ALDH2 genes exist as polymorphisms with
The liver is the major organ for converting ethanol into acetalde- the most relevant ALDH2 variants including the ALDH2*1 and
hyde via the three enzymatic pathways [3]. The liver expresses ALDH2*2 alleles. The ALDH2*1 allele encodes active acetalde-
high levels of alcohol dehydrogenase (ADH), which is the most hyde metabolizing enzyme with G at nucleotide position 42421
important enzyme for converting ethanol to acetaldehyde. with corresponding glutamate at amino acid position 487;
Human ADH genes encode at least seven isozymes including whereas the ALDH2*2 allele encodes inactive enzyme with A at
ADH1–7 with the major forms of ADH1, ADH2, and ADH3 in position 42421 and lysine at position 487. People with homozy-
human liver. ADH exists in cytosol and oxidizes ethanol into gous ALDH2*1/1 have high levels of acetaldehyde metaboliz-
acetaldehyde and reduces NAD+ to NADH, which generates a ing activity in the liver; whereas individuals with homozygous
highly reduced cytosolic environment in cells. Normal liver also ALDH2*2/2 have very low or undetectable enzyme activities.
expresses high levels of the cytochrome P450 isozyme CYP2E1, Individuals with heterozygous ALDH2*2/1 have approximately
which is located in the endoplasmic reticulum and is  highly 80–90% lower enzyme activity because ALDH2 is the isote-
induced after chronic ethanol consumption. CYP2E1 can oxi- tramer enzyme and all four subunits of ALDH2 are needed to be
dize ethanol into acetaldehyde with conversion of normal for keeping its full activity. Individuals with the inactive
NADPH+H++O2 to NADP++2H2O, thereby generating reactive variants of ALDH2 exhibit acetaldehyde accumulation after

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
53:  Alcoholic Liver Disease 683

alcohol consumption, and present an acetaldehyde‐mediated a high-fat diet and coadministration of ­ethanol synergistically
“flushing syndrome,” which includes facial flushing, palpita- induced an elevation of serum alanine aminotransferase (ALT)
tions, drowsiness, and other unpleasant symptoms. levels, hepatic steatosis, liver inflammation, and fibrosis via the
induction of nitrosative, endoplasmic reticulum, and mitochon-
drial stress and upregulation of hepatic Toll‐like receptor 4 [6].
The type of fat (saturated versus unsaturated) may also affect
SPECTRUM OF ALD the outcome of ALD. Saturated fat has been shown to be protec-
tive against ALD, while unsaturated fat is deleterious for ALD
ALD has a broad spectrum of disorders from simple fatty liver in animal models [7]. However, how dietary factors affect
(steatosis) to severe forms of steatohepatitis, cirrhosis, and hepa- ALD  in humans is probably complex and remains poorly
tocellular carcinoma (HCC) [1, 2]. Fatty liver occurs in almost characterized.
all chronic heavy drinkers with early fat accumulation in zone 3
(perivenular) hepatocytes, which can extend to zone 2 and even
zone 1 (periportal) hepatocytes when liver damage becomes Drinking patterns
severe. Steatohepatitis (steatosis and inflammation) and cirrho- During the last five years, one of the major advances in the field
sis are detected in approximately 20–40% and 8–20% of chronic of ALD research was the discovery that acute ethanol gavage
heavy drinkers, respectively. Chronic excessive drinking can caused significant neutrophilia, liver injury, and inflammation
also cause HCC, which occurs mostly after cirrhosis in approxi- in chronically ethanol‐fed mice [8–11] and in high‐fat diet‐fed
mately 3–10% of excessive drinkers. Many patients with chronic mice [5]. In agreement with these findings from animal models,
ALD are asymptomatic with fully compensated and preserved clinical studies show that individuals with AUD with recent
liver function, but some of them may develop episodes of super- excessive alcohol consumption have higher levels of circulating
imposed alcoholic hepatitis (AH) with obvious jaundice, fever, neutrophils, serum alanine transaminase (ALT), and aspartate
abdominal pain, anorexia, and weight loss [4]. AH has high transaminase (AST) compared with individuals with AUD with-
short‐term ­mortality especially in those with cirrhosis [4]. out recent drinking [12]. The levels of circulating neutrophils
correlate positively with serum ALT and AST in individuals
with AUD with recent excessive drinking, suggesting that recent
excessive drinking elevates circulating neutrophils and subse-
GENETIC FACTORS, COMORBIDITY, quently induces liver injury [12]. In addition, early studies sug-
AND ALD gest that frequent heavy drinking that begins at an early age is a
risk factor for the development of severe forms of ALD com-
Although almost all heavy drinkers develop fatty liver, only pared with episodic or binge drinking [13]. However, a recent
about 35% of them develop advanced ALD, which is probably prospective cohort study suggests that recent alcohol drinking
because other risk factors are involved in the pathogenesis of rather than earlier in life is associated with risk of alcoholic
ALD. Indeed, gender, obesity, dietary factors, drinking patterns, ­cirrhosis [14].
non‐sex‐linked genetic factors, and cigarette smoking have been
suggested to affect susceptibility to ALD.
Non‐sex‐linked genetic factors
Gender Many non‐sex‐linked genetic factors have been implicated in
affecting the susceptibility to advanced ALD. These include
Female sex is a well‐documented risk factor for susceptibility to variations in genes that encode antioxidant enzymes, cytokines
ALD in rodent models. It is believed that the increased risk and other inflammatory mediators, and alcohol‐metabolizing
among women is probably due to lower levels of gastric ADH, enzymes. However, the evidence that links most of these genetic
a higher proportion of body fat, and the presence of estrogens. factors to ALD is weak, only a few of these factors seem to be
strongly associated with ALD. For example, a single‐nucleotide
variant (I148M) in human patatin‐like phospholipase domain‐
Obesity and dietary factors
containing 3 gene (PNPLA3), known as I148M variant, is one of
Obesity is another important risk factor that accelerates ALD the best characterized variants and has strong correlations with
development and progression in alcohol use disorder (AUD). increased risk of steatosis, fibrosis, and cirrhosis in various
Experimental studies indicate that acute or chronic ethanol feed- types of liver diseases including ALD [15].
ing and obesity synergistically induced liver injury and inflam-
mation via the activation of endoplasmic reticulum stress, hepatic
macrophage and neutrophilic infiltration, and lipotoxicity. For
Viral hepatitis
example, acute ethanol gavage induces acute steatohepatitis in Viral hepatitis infection is a leading cause of chronic liver dis-
obese mice induced by high‐fat diet feeding but not in chow‐fed ease worldwide. The synergistic effects of excessive alcohol
mice [5]. Acute ethanol gavage upregulates the expression of drinking and viral hepatitis B or C infection on liver disease
chemokine (C‐X‐C motif) ligand 1 (CXCL1) mRNA in the liver progression have been well documented. It has been reported
but not in other organs by 30‐fold in obese mice, which attracts that excessive alcohol drinking markedly accelerates liver fibro-
neutrophil accumulation in the liver and subsequently causes sis and HCC development and progression in patients with viral
liver injury [5]. In addition, intragastric overfeeding of mice with hepatitis C infection.
684 THE LIVER:  ALCOHOLIC STEATOHEPATITIS (ASH) AND ALCOHOLIC HEPATITIS (AH)

Non‐alcoholic fatty liver disease (NAFLD) at high levels in adipose tissues, playing a critical role in promot-
ing lipid droplet formation. In normal liver, FSP27 is expressed
Excessive alcohol consumption likely worsens NAFLD pro- at very low levels but highly upregulated post alcohol consump-
gression; while the reports regarding the effects of moderate tion, especially acute alcohol consumption [9]. Upregulation of
alcohol drinking on NAFLD have been controversial [16]. hepatic FSP27 by acute ethanol consumption is attributable to
Current clinical data are not sufficient to support a recommen- activation of hepatic endoplasmic reticulum stress, which acti-
dation for or against moderate alcohol consumption in vates the cyclic AMP‐responsive element‐binding protein H
NAFLD [16]. (CREBH) and subsequent nuclear translocation of nCREBH [9].
The important role of FSP27 in inducing alcoholic fatty liver is
Other comorbidities supported by the fact that disruption of the Fsp27 gene prevents
the development of alcoholic fatty liver [9]. SREBP‐1c is a mas-
Long‐term alcohol consumption may also accelerate liver dis- ter transcription factor for controlling hepatic expression of
ease progression in patients with human immunodeficiency many lipogenic genes that encode proteins and enzymes to
virus infection, hemochromatosis, and so on. A greater under- induce FA biosynthesis. Disruption of the Srebp‐1c gene mark-
standing of the interaction between alcohol and these comorbid edly prevents alcoholic fatty liver in mice, suggesting a critical
factors could help us design better therapies for the treatment of role of SREBP‐1c in inducing alcoholic fatty liver [21]. Alcohol
chronic liver disease. consumption can directly upregulate hepatic SREBP‐1c expres-
sion via its metabolite acetaldehyde [22], or indirectly stimu-
late its expression by modulating multiple factors and pathways
that control hepatic expression of SREBP‐1 (see references
ALCOHOLIC FATTY LIVER therein [1, 17]). In contrast to promotion of FA biosynthesis,
alcohol attenuates FA β‐oxidation, which is another critical
Fat accumulation in the liver (fatty liver) is the earliest response mechanism responsible for the formation of alcoholic fatty
of the liver to acute and chronic alcohol drinking [17]. Fatty liver. Early studies revealed that alcohol treatment increases the
liver is diagnosed when the liver fat weight is greater than 5 to ratio of reduced nicotinamide adenine dinucleotide/oxidized
10% of the liver’s weight and is characterized by the accumula- nicotinamide adenine dinucleotide in the liver and conse-
tion of triglycerides, phospholipids, cholesterol esters, and other quently disrupts the mitochondrial β‐oxidation of FAs and
types of fat in hepatocytes. Although alcohol drinking causes fat results in fat accumulation in hepatocytes [23]. Recent studies
accumulation in the liver, there is no evidence that ethanol and unveiled an additional mechanism by which ethanol inhibits
its metabolite acetaldehyde directly participate in biosynthesis FA β‐oxidation, which is inactivation of peroxisome prolifera-
of fatty acids (FAs). Acetate, the metabolite of acetaldehyde, tor‐activated receptor (PPAR)‐α, a nuclear hormone receptor
can be converted to acetyl‐CoA by acetyl‐CoA synthetases with that upregulates expression of a range of genes involved in FA
acetyl‐CoA synthetase 2. Once it is formed, acetyl‐CoA enters transport and oxidation [24].
the citric acid cycle (Krebs cycle) and contributes to fatty acid Alcohol consumption can also promote hepatic fat accumula-
biosynthesis. However, how much this acetate–acetyl‐CoA tion by inhibiting 5’ AMP‐activated protein kinase (AMPK) and
pathway directly contributes to the formation of alcoholic fatty deregulating autophagy. AMPK plays a critical role in control-
liver is not clear because acetate derived from ethanol metabo- ling the activity of several enzymes and transcriptional factors
lism in hepatocytes is rapidly secreted into the circulation and that regulate fat metabolism [25]. For example, AMPK inhibits
does not maintain high concentrations in hepatocytes, and the acetyl‐CoA carboxylase (ACC) and SREBP activity and conse-
liver does not express the mitochondrial acetyl‐CoA synthetase quently attenuates FA biosynthesis but enhances its oxidation
2 that is a major form for converting acetate into acetyl‐CoA. [25]. Alcohol exposure can suppress AMPK activity and subse-
Emerging evidence suggests that alcohol consumption pro- quently enhance FA synthesis and reduces FA oxidation, result-
motes fat mobilization to the liver and de novo FA biosynthesis ing in alcoholic fatty liver [26]. Autophagy plays an important
but inhibits fat clearance and FA β‐oxidation in the liver, result- role in clearing lipid droplets in hepatocytes, and chronic alco-
ing in fat accumulation in the liver [1, 17]. Early studies demon- hol consumption inhibits autophagy, thereby reducing lipid
strated that acute alcohol administration can increase the supply clearance and causing hepatic steatosis [27]. However, acute
of dietary lipids to the liver from the small intestine by increasing alcohol intake seems to activate autophagy, which may play a
the intestinal lymph flow and output of both dietary and nondie- compensatory role in preventing the development of fatty liver
tary lipids, but this effect was less evident after chronic alcohol during the early stages of alcoholic liver injury [28].
treatment. Recent studies focused on how alcohol consumption
increases the supply of lipids from adipose tissues to the liver
[18]. It has been shown that alcohol consumption induces lipoly-
sis [19] and adipocyte death [20] and subsequently elevates cir- ALCOHOLIC STEATOHEPATITIS (ASH)
culating FAs and their accumulation in the liver. In addition to AND ALCOHOLIC HEPATITIS (AH)
augmenting FA supply, alcohol can also increase fat accumula-
tion by promoting lipid formation and stimulating de novo FA Accumulation of fat in hepatocytes is the first response of the
biosynthesis in hepatocytes via the upregulation of fat‐specific liver to alcohol consumption [17]; however, alcohol consump-
protein 27 (FSP27) and sterol  ­regulatory element‐binding pro- tion may also cause hepatocyte damage and inflammation,
tein 1c (SREBP‐1c), respectively [1, 9, 17]. FSP27 is expressed which is defined as ASH [4]. ASH is a histological concept that
53:  Alcoholic Liver Disease 685

is characterized by steatosis, hepatocyte death, ballooning are associated with severe cases of AH. Most patients with AH
degeneration, neutrophilic infiltration, Mallory–Denk bodies, have pre‐existing alcoholic cirrhosis, while some patients with
and chicken‐wire fibrosis [4]. Most patients with ASH are mild underlying liver disease or silent ALD can also develop
asymptomatic for prolonged periods of time and are defined as AH. Diagnosis of AH is based on clinical syndromes with an
walking ASH or subclinical ASH, but some patients with ASH episode of jaundice with serum bilirubin levels greater than 3
develop a significant clinical syndrome, which is defined as AH mg dL−1 and AST levels greater than 50 IU mL−1 in chronic
(see New therapeutic targets for AH) [4]. Over the last four dec- AUD with heavy drinking greater than 5 years [33]. Jaundice is
ades, multiple mechanisms have been identified to contribute to an essential syndrome for the diagnosis of AH, but not all epi-
ASH caused by alcohol drinking (Figure  53.1). First, alcohol sodes of jaundice in AUD with chronic ALD are attributed to
and its metabolite acetaldehyde can directly cause hepatotoxic- AH [4]. For example, severe sepsis, biliary obstruction, diffuse
ity by generating ROS and inducing endoplasmic reticulum liver cancer, drug‐induced liver injury, and ischemic hepatitis
stress and mitochondrial dysfunction [29]. Second, damaged (i.e. due to massive bleeding or cocaine use) can also cause an
hepatocytes caused by alcohol release damage‐associated episode of jaundice and liver decompensation in AUD, which
molecular patterns (DAMPs) induce liver inflammation, such as should not be diagnosed as AH [4]. Patients with severe AH
neutrophil infiltration [30]. Third, chronic alcohol consumption especially those with cirrhosis usually have high short‐term
induces gut bacterial overgrowth and dysbiosis and increases mortality with a mortality rate of 20–50% at three months [32].
gut permeability, resulting in elevation of bacteria and their Liver histology analysis of AH revealed infiltration of a large
related products in the liver and circulation and subsequent number of inflammatory cells in the liver with predominant
inflammation [31]. neutrophils and macrophages [30]. Microarray analysis has
AH is a severe form of ALD that presents a form of acute‐on‐ identified that a large number of inflammatory mediators are
chronic liver failure in individuals with AUD with underlying upregulated in the liver from AH patients [34]. In addition to
ALD including ASH and cirrhosis [4, 32, 33]. AH is character- massive liver inflammation, a systemic inflammatory response
ized by an abrupt rise in jaundice (serum bilirubin levels) and is also a key feature of AH and correlates with disease severity
liver‐related complications and is often accompanied by fever, [35]. The inflammation in most AH patients is likely triggered
abdominal pain, anorexia, and weight loss [33]. In addition, por- by recent excessive binge drinking, bacterial infection, or both
tal hypertension, ascites, encephalopathy, and variceal bleeding [4]. Recent excessive binge drinking was reported in many

Figure 53.1  Pathogenesis and molecular mechanisms that trigger liver failure, systemic inflammatory response, and multiorgan failure in alco-
holic hepatitis. Excessive binge drinking in AUD with underlying ALD induces hepatocellular damage, induces systemic inflammation, and impairs
liver regeneration, resulting in liver failure and other complications in alcoholic hepatitis. Modified from [4] and reproduced with permission
from Elsevier.
686 THE LIVER:  EMERGING MECHANISMS

patients with AH, and can cause elevation of serum ALT and promoting HSC activation and liver fibrosis, but ­activation of
AST levels as well as circulating neutrophils in AUD [12]. It is some components of innate immunity negatively regulates HSC
possible that excessive binge drinking causes massive hepato- activation and suppresses liver fibrosis. For example, activation
cyte damage, which releases a large amount of DAMPs that trig- of natural killer (NK) cells can directly kill activated HSCs and
ger systemic inflammation in patients with AUD. For example, produce IFN‐γ that blocks HSC proliferation and induces HSC
damaged hepatocytes can release mitochondrial DNA, which death. Such an antifibrotic effect of NK cells is markedly sup-
can activate various types of inflammatory cells via Toll‐like pressed after chronic alcohol consumption, thereby accelerating
receptors [11]. Bacterial infection is detected in up to 50% of liver fibrosis [38].
AH patients and is associated with high mortality in these
patients [36]. Pathogen-associated molecular patterns (PAMPs)
produced during bacterial infection including endotoxin, can
activate macrophages/Kupffer cells to produce a wide variety of
ALCOHOLIC LIVER CANCER
inflammatory mediators in AH [35].
Liver cirrhosis derived from all etiologies of liver pathology
Normal liver has the great ability to regenerate after injury or
including alcohol consumption is a major risk factor for liver
loss of tissue; however, hepatocytes in the injured liver in AH
cancer, mainly HCC. The mechanisms by which cirrhosis pro-
patients seem unable to efficiently regenerate. Instead, liver
motes HCC are not fully understood. Several hypotheses have
­progenitor cells are markedly expanded and accumulated in AH
been proposed. First, cirrhosis is associated with telomere
patients, and the number of these cells positively correlates with
shortening, which can induce chromosomal instability and
disease severity. It is likely that these liver progenitor cells are
mutations of tumor suppressors and oncogenes. Second, cirrho-
unable to fully differentiate into functional hepatocytes and do
sis is associated with chronic inflammation and elevation of
not contribute to liver regeneration. Thus, impaired liver regen-
many growth factors and cytokines that stimulate HCC growth.
eration is probably an important mechanism contributing to
In addition to these common mechanisms, there are some
liver failure in patients with severe AH (Figure 53.1).
exceptional mechanisms that are believed to specifically con-
tribute to alcoholic liver cancer development. Acetaldehyde, an
ethanol metabolite, is considered a carcinogen with mutagenic
properties. It is well documented that ALDH2‐deficient indi-
ALCOHOLIC LIVER FIBROSIS
viduals have high levels of acetaldehyde after alcohol drinking
and are associated with an increased risk of esophageal cancers;
Fibrosis is a scarring response of the liver to chronic liver dam-
however, the ­association between HCC and individuals with
age caused by virtually all etiologies of liver pathology includ-
AUD with ALDH2 deficiency is less clear. It is known that acet-
ing excessive alcohol drinking. Although all types of chronic
aldehyde is electrophilic and forms an adduct with DNA and
liver injury cause liver fibrosis, alcoholic liver fibrosis has a
interstrand crosslinks, causing DNA damage. In addition, acet-
characteristic pattern of pericellular and sinusoidal chicken‐
aldehyde suppresses DNA repair enzyme activity and subse-
wire fibrosis, which could be extended to panlobular fibrosis
quently blocks DNA repair. Thus, acetaldehyde not only causes
when this characteristic pattern becomes diffused [37]. Liver
DNA damage but also prevents DNA repair, contributing to
fibrosis is characterized by excessive accumulation of collagen
hepatocarcinogenesis. Moreover, heavy drinking can induce
and other extracellular matrix (ECM) proteins. Most of these
aberrant DNA methylation including genome‐wide hypometh-
ECM proteins are produced by activated hepatic stellate cells
ylation, leading to chromosomal instability and subsequently
(HSCs). Several other types of cells can also produce ECM
promoting the development of liver cancer.
­protein, but to a much lesser extent. These cells include portal
Excessive alcohol drinking is known to cause broad immune
fibroblasts and bone marrow‐derived myofibroblasts. Activation
suppression including attenuation of antitumor immunity, which
of HSCs, a key step in developing liver fibrosis, is controlled by
is probably another important mechanism that can accelerate
many inflammatory mediators and growth factors that are com-
HCC development and progression.
monly elevated in ALD and other types of liver diseases [37].
For example, transforming growth factor (TGF)‐β is one of the
most important mediators that induces HSC transformation;
while platelet‐derived growth factor (PDGF) is a critical growth EMERGING MECHANISMS
factor that stimulates HSC proliferation. Both TGF‐β and PDGF
are elevated in ALD in patients and animal models of alcoholic
Gut microbiome
liver fibrosis [37]. In addition, there are some unique mecha-
nisms that promote alcoholic liver fibrosis [37]. First, acetalde- An important role of gut bacteria‐derived LPS in the develop-
hyde, the first ethanol metabolite, can directly stimulate and ment of ALD in animal models was reported more than two
maintain HSC activation by stimulating multiple signaling path- ­decades ago. Recent studies suggest that the gut microbiome
ways and transcription factors. Second, ALD patients exhibit plays a complex role in the pathogenesis of ALD [31]. Chronic
elevated lipopolysaccharides (LPSs) due to increased gut bacte- alcohol consumption can cause bacterial overgrowth in the large
rial growth and gut permeability. LPS is a well‐known mediator and small intestines and dysbiosis in humans and animals [31].
that can directly induce HSC activation, thereby playing an In general, changes in gut microbiome post chronic alcohol
important role in promoting alcoholic liver fibrosis. Third, acti- consumption are complex but are often associated with a
­
vation of innate immunity is known to play a critical role in decrease in “good” bacteria (such as Lactobacillus spp.) but an
53:  Alcoholic Liver Disease 687

increase in “bad” bacteria (such as Enterobacteriaceae). These Extracellular vesicles (EVs)


changes could cause pathological bacterial translocation, altera-
tions of intestinal metabolites and bile acid metabolism, result- EVs are nanometer‐sized, membrane‐bound, extracellular milieu
ing in elevation of systemic levels of gut‐derived microbial vesicles derived from cells, playing an important role in connect-
production and subsequent hepatocellular damage and liver ing cell to cell communication and in promoting inflammation in
inflammation [31]. In addition, a recent study suggests that gut a variety of diseases, including liver diseases. EVs can be
fungi may also play a role in promoting ALD [39]. Chronic grouped as three types based on their cellular biogenesis and
alcohol consumption increases microbiota populations in gut sizes: exosomes (40–150 nm diameter in size), microvesicles/
and treatment with antifungal drugs attenuated intestinal fungal microparticles (MPs; 50–1000 nm diameter), and apoptotic bod-
overgrowth and ameliorated ALD in mice. Patients with ALD ies (greater than 500 nm in diameter). EVs exert their functions
have reduced intestinal fungal diversity and Candida overgrowth by delivering their contents from one cell to another. For instance,
and increased systemic exposure and immune response to myc- stressed hepatocytes can activate macrophages and neutrophils
obiota. The levels of these responses correlate with m ­ ortality, via the release of EVs that contain lipids, proteins, chemokines,
suggesting that altered gut mycobiomes contribute to ALD and nucleic acids (e.g. mitochondrial DNA [mtDNA]), playing a
pathogenesis [39]. Moreover, transplantation of intestinal role in the pathogenesis of ALD [43]. Patients with AUD or ALD
microbiota from a patient with severe AH induced greater liver had an elevated number of EVs in the circulation, which are
inflammation and injury in mice than transplantation of these enriched in mtDNA and correlate with elevated circulating neu-
from an individual with AUD without AH [40]. Thus, manipu- trophils and serum levels of ALT [11]. Compared with pair‐fed
lating the intestinal microbiome (both bacteria and fungi) might control mice, mice post chronic or chronic‐plus‐binge ethanol
be an effective strategy for preventing and treating ALD, and feeding also have elevated circulating EVs, which are enriched
fecal microbiota transplant might be a potential therapeutic in mRNAs, proteins, mtDNAs, and likely promote liver inflam-
option for ALD. mation and injury in ALD [11, 44]. In addition to being an
important player in ALD, EVs may be used as potential biomark-
ers for diagnosis because EVs from ALD contain specific signa-
Epigenetics tures of proteins, RNAs, and DNAs [43].

Epigenetics is the investigation of heritable alterations of phe-


notype (appearance) or gene expression that are caused by Adaptive immunity
mechanisms other than alterations in DNA coding sequences. The role of innate immunity in the pathogenesis of ALD has
Epigenetic modifications include DNA methylation, histone been well documented; however, the involvement of adaptive
modifications, and RNA‐based mechanisms. Alcohol consump- immunity in ALD remains obscure. It is generally believed that
tion is known to affect metabolism of methionine and thereby excessive alcohol consumption induces oxidative stress, which
DNA methylation [41]. In the liver, homocysteine is methylated subsequently generates lipid peroxidation products, such as
to methionine and then S‐adenosylmethionine (SAMe) in a malondialdehyde and 4‐hydroxynonenal. These products can
transmethylation reaction that is catalyzed by methionine aden- modify many proteins and induce formation of protein adducts
osyltransferase. SAMe is a principal methyl donor in methyla- that can act as neoantigens to induce the activation of adaptive
tion and plays a critical role in inducing DNA and histone immunity [45]. Activation of adaptive immunity has not been
methylation. Excessive alcohol consumption reduces hepatic characterized or detected in animal models of ALD, but it was
levels of SAMe and DNA and histone methylation, which sub- reported in patients with ALD [45]. For example, patients with
sequently upregulates the expression of genes that control the ALD are associated with increased levels of circulating antibod-
endoplasmic reticulum stress response and ALD [41]. ies against lipid peroxidation adducts and increased numbers of
intrahepatic T cells [45]. The infiltration of intrahepatic T cells
in ALD is not just bystander activation, and these T cells present
miRNAs a pronounced oligoclonal nature along with ALD‐associated
miRNAs (small non‐coding RNA molecules with 19–25 nucle- clonotypes as demonstrated in a recent high‐throughput T‐cell
otides) play a critical role in controlling gene expression by receptor sequencing study [46]. If activation of adaptive immu-
inducing RNA silencing and by altering post‐transcriptional nity is a major driver of disease progression in some ALD
regulation of gene expression. Many miRNAs have been shown patients, immunosuppressive therapy may be required to
to play important roles in the development of ALD. For exam- ­effectively treat these patients.
ple, hepatocyte‐specific miR122 seems to play a role in protect-
ing against ALD by reducing hypoxia‐inducible factor 1 alpha
Innate‐like T cells
(HIF1α) mRNA. Hepatic miR122 is downregulated in liver
samples from patients with ALD and ethanol‐fed mice, thereby The roles of Kupffer cells/macrophages in the pathogenesis of
exacerbating liver injury and inflammation [42]. In contrast, ALD have been extensively studied over the last two decades.
neutrophil‐specific miR‐223 is upregulated in the liver and Recent studies suggest that nature killer T (NKT) cells and
serum from ethanol‐fed mice and individuals with AUD com- mucosa‐associated invariant T cells (MAIT) also play a role in
pared to their controls [12]. Such elevations likely play a com- the pathogenesis of ALD. Mouse liver lymphocytes contain
pensatory role in limiting neutrophil activation via inhibition of approximately 30–40% natural killer T (NKT) cells, which are
the IL‐6–p47phox pathway [12]. a heterogeneous group of T cells that share properties of both
688 THE LIVER:  CLINICAL DIAGNOSIS AND TREATMENT OF ALD

T  cells and NK cells and rapidly produce a large number of alcohol drinking and evidence of liver disease are the key diag-
cytokines such as IFN‐γ, IL‐4, and so on. Several recent studies nostic factors of ALD.
reported that hepatic NKT cells are activated and increased
post chronic‐plus‐binge ethanol feeding. These activated NKT
cells contribute to Kupffer cell activation and alcoholic liver
Medical history
injury [47, 48]. Of note, human liver lymphocytes contain a Denial of excessive alcohol use or underreported alcohol intake
very low number of NKT cells but are enriched with MAIT in the medical history is a frequent occurrence in AUD patients,
cells, representing 20–50% of intrahepatic T cells in normal which may make diagnosis of ALD challenging in these
human livers. MAIT cells are detected at low levels in com- patients. Therefore, indirect evidence of excessive alcohol use
monly used laboratory mouse strains, accounting for less than is always considered, such as questionnaires, information from
1% of intrahepatic T cells in mice. MAIT cells play a key role family members, and laboratory tests to reinforce or confirm
in host defense against bacterial infection through invariant the suspicion of AUD. Various questionnaires for screening
T‐cell receptors (TCRs) that recognize microbial riboflavin/ ­excessive alcohol use are used by clinicians, such as the CAGE
vitamin B2 metabolites presented by the major histocompatibil- (cut‐annoyed‐guilty‐eye), MAST (Michigan Alcoholism
ity complex class I‐related protein 1. Patients with severe ALD Screening Test), and AUDIT (Alcohol use disorders identifica-
are associated with a marked reduction of MAIT cells [49], tion test) [2]. The CAGE and MAST questionnaires are the
which may contribute to an increased risk of bacterial infection most commonly used.
in these patients.
Clinical symptoms and signs
Chronic‐plus‐binge ethanol
Anorexia, nausea and vomiting, upper abdominal pain, malaise,
Over the last 5 years, one important advance in the field of ALD fatigue, dark urine, fever, confusion, and weight loss are the most
study is the introduction of binge ethanol in chronically ethanol‐ commonly presenting nonspecific symptoms of ALD. The most
fed mice [50] or in high‐fat diet‐fed mice [5]. This chronic‐plus‐ common sign of AH is rapid development of jaundice; others
binge ethanol feeding model was initially called the NAAA include fever, ascites, proximal muscle wasting, tender hepato-
model [50] and was later also called the Gao‐binge model [51]. megaly, and hepatic bruit. Severe cases often develop hepatic
Such an ethanol challenge induces more severe alcoholic steato- encephalopathy and gastrointestinal bleeding. Patients with mild
hepatitis with marked elevation of serum ALT and AST and sig- to moderate ALD are often asymptomatic. In some patients, bilat-
nificant hepatic neutrophil infiltration compared with binge eral parotid gland hypertrophy, muscle wasting, malnutrition,
ethanol alone, chronic ethanol feeding alone, and high‐fat diet Dupuytren’s contracture sign, and signs of symmetric peripheral
feeding alone [5, 8, 52]. Acute ethanol gavage was also intro- neuropathy are suggestive of harmful alcohol consumption.
duced in the chronic intragastric ethanol infusion model, which Splenomegaly, gynecomastia, spider angiomas, asterixis, palmar
causes marked hepatic neutrophil infiltration [53]. By using erythema, and digital clubbing are frequently observed in ALD
these new animal models, researchers have identified many patients with cirrhosis [2]. In decompensated cirrhosis, jaundice,
novel mechanisms that contribute to acute‐on‐chronic liver ascites, peripheral edema, and hepatic encephalopathy are always
injury in ALD. These include activation of endoplasmic reticu- seen in addition to the cirrhotic physical ­findings [2].
lum stress, activation of inflammatory cells (e.g. neutrophils,
Kupffer cells, and NKT cells), upregulation of several factors
that promote steatosis (e.g. FSP27), and dysregulation of cell
Laboratory tests
survival/death signaling pathways (e.g. pyroptosis, apoptosis, There are no specific laboratory biomarkers for ALD. Elevations
etc.) [10, 53]. For example, several studies suggest that neutro- of serum enzymes such as AST, ALT, alkaline phosphatase
phils contribute to hepatocellular damage and hepatic injury in (ALP), and GGT can only provide clues for the diagnosis of
this early stage of ALD that is induced by chronic‐plus‐binge ALD. The NIAAA Alcoholic Hepatitis Consortia proposed a
ethanol feeding or high‐fat diet‐plus‐binge ethanol feeding working definition of acute AH to include an AST to ALT ratio
[5, 8]. However, neutrophils can also promote liver repair and above 1.5 with AST and ALT less than 400 U L−1 [33]. Clinically,
are vital in the control of bacterial infection. Therefore, neutro- GGT is the most frequently used biomarker to detect previous
phils likely play both beneficial and detrimental roles in the alcohol consumption, though it can also arise from other etiolo-
­pathogenesis of ALD. gies [55]. Routine blood tests and biochemical detection, such
as elevations of red blood cell mean corpuscular volume (RBC
MCV), white blood cell (WBC) count, GGT, AST, and immu-
noglobulin A (IgA) may indicate early ALD, while a decrease in
CLINICAL DIAGNOSIS AND TREATMENT albumin, prolonged prothrombin time (PT), elevated total bili-
OF ALD rubin (TBIL) level, and low platelet (PLT) count are indications
of advanced ALD [56]. Recent studies suggest that a new bio-
The diagnosis of ALD is based on the combination of clinical marker of apoptotic and necrotic cell death, the M30/M65 frag-
and laboratory findings, including a medical history of signifi- ment of cytokeratin 18 (CK18), is sensitive to liver injury in
cant alcohol intake, physical signs of liver disease, supporting ALD patients [57]. In some ALD patients, metabolic abnormali-
laboratory tests of liver disease, non‐invasive liver imaging, and ties such as elevated levels of triglycerides and uric acid,
invasive liver biopsy [2, 54]. Documentation of excessive hypokalemia, and hyponatremia are frequently observed.
53:  Alcoholic Liver Disease 689

Liver imaging end‐stage liver disease (MELD), the Glasgow alcoholic hepati-
tis score (GAHS), the Age, bilirubin, INR, creatinine (ABIC)
Ultrasound, computerized tomography, and magnetic resonance score, the Lille model, and the Child–Turcotte–Pugh (CTP)
imaging can be used to detect the presence of underlying liver score [2]. MDF (4.6 × [patient prothrombin time – control pro-
disease, but they cannot provide specific information for diag- thrombin time] + serum total bilirubin) is the most widely used
nosing ALD. Ultrasound is routinely used to check hepatocel- [2]. ALD patients with a MDF value greater than or equal to 32
lular carcinoma, biliary obstruction, ascites, splenomegaly, is indicative of a high risk of short‐term mortality (30–50% at
portal hypertension, and portal vein thrombosis. Non‐contrast one month) with improved short‐term clinical outcomes after
computerized tomography more easily detects macroscopic fat receiving corticosteroid [60].
in the liver in fatty liver disease [58]. Magnetic resonance imag-
ing is the most sensitive and specific imaging modality for
detecting steatosis (95% sensitivity, 98% specificity) except for
patients with iron overload [59]. Recently, the clinical applica- TREATMENT OF ALD
tions of newer imaging techniques such as transient elastogra-
phy (TE, Fibroscan®) and magnetic resonance elastography Alcohol abstinence
(MRE) improve diagnostic accuracy in quantifying steatosis
and fibrosis [60]. Among all imaging modalities, ultrasound is The backbone therapeutic intervention for patients with ALD is
the most widely used due to its low cost. By detecting values of alcohol abstinence. Following abstinence, relapse is a major risk
controlled attenuation parameter (CAP value) and liver stiff- in these patients. Continued alcohol use after diagnosis of ALD
ness, transient elastography (TE, Fibroscan) has demonstrated has been shown to promote disease progression [2, 54].
good accuracy in quantifying steatosis and fibrosis, and it is a
significantly cheaper alternative [54]. Nutritional support
Malnutrition is very common in patients with severe ALD, such
Liver biopsy as deficiencies of vitamin A, vitamin D, vitamin E, thiamine,
Liver biopsy is not routinely recommended for the diagnosis of folate, niacin, pyridoxine, zinc, magnesium, and selenium. A
ALD in most patients in the United States. However, it is dietary intake of 1.2–1.5 g of protein kg−1 and 35–40 kcal kg−1
believed to be the gold standard for diagnosing and assessing body weight is recommended in these patients [60].
the severity of steatosis and staging of fibrosis, and it is the only
modality that is available to distinguish between simple steato- Corticosteroids
sis and steatohepatitis [2]. Liver biopsy is invasive and mostly
Corticosteroids are recommended for patients with severe AH in
­performed percutaneously. A transjugular route of biopsy is
the absence of sepsis and infection, which can improve short‐
required in patients with coagulopathy or ascites. The histologic
term survival rates in these patients [2]. However, long‐term fol-
features of ALD on liver biopsy vary based on the extent and
low‐up does not reveal a significant survival difference between
stage of hepatic injury. The typical histological lesions of ALD
those treated with corticosteroids and controls [2].
are large fat droplets and ballooning of hepatocytes including
Mallory–Denk bodies and mega‐mitochondria as well as neu-
trophil infiltration and intrasinusoidal chicken‐like fibrosis [60]. Pentoxifylline (PTX)
Macrovesicular steatosis is the earliest and most frequently seen
PTX is a competitive, nonselective, phosphodiesterase inhibitor
pattern of alcohol‐induced liver injury. Alcoholic steatohepatitis
that can inhibit tumor necrosis factor and leukotriene synthesis,
is defined by the coexistence of steatosis, hepatocyte balloon-
inflammation, and innate immunity. PTX can also reduce the
ing, and neutrophil infiltration [54].
development of hepatorenal syndrome (HRS) in ASH patients,
resulting in improvements in short‐term survival [61].
Complications in ALD
Cholestasis, fat embolism, and portal hypertension are occa- N‐acetyl cysteine (NAC)
sionally observed in patients with severe fatty liver caused by NAC has antioxidant activity and is ineffective when used alone
alcohol. Alcoholic ketoacidosis is often observed in patients in ALD patients. However, a randomized trial showed that the
with chronic alcohol consumption and concurrent malnutrition. combination of NAC with prednisolone reduced one‐month
Patients with alcoholic cirrhosis or severe AH always have mortality and incidence of HRS/infection [62].
­complications such as ascites, spontaneous bacterial peritonitis
(SBP), variceal bleeding, electrolyte disturbance, hepatorenal
syndrome (HRS), hepatic encephalopathy, and HCC [2]. Artificial liver support system
Artificial liver support systems are a therapy option for liver
failure associated with severe ALD, which has been controver-
Assessment of ALD sial for decades. Recently, there was a study that showed that the
There are several prognostic scoring models for assessing application of an artificial liver support system in patients with
­disease severity and short‐term mortality of ALD, including the severe liver dysfunction secondary to ALD showed some bene-
Maddrey discriminant function (MDF), the Mayo model for fits in a subset of the patients [63].
690 THE LIVER:  NEW THERAPEUTIC TARGETS FOR ALCHOLIC HEPATITIS

Liver transplantation hyperimmune bovine colostrum enriched with anti‐LPS anti-


bodies, and so on. Collectively, many inflammatory mediators
Patients with end‐stage liver disease secondary to alcoholic have been implicated in liver injury and inflammation in patients
­cirrhosis should be considered for liver transplantation. Early with severe AH. Many of these ­mediators probably synergisti-
liver transplantation, which has been performed for acute severe cally or additively promote liver inflammation in AH. Clinical
AH, is discussed in the following section. trials are needed to identify the inflammatory mediators that
play critical roles in the pathogenesis of ALD in patients and
can be used as targets for the ­treatment of ALD, which will take
years to complete. The immunosuppressive drug prednisolone
NEW THERAPEUTIC TARGETS or other steroids will likely continue to be used for the treatment
FOR ALCHOLIC HEPATITIS of severe AH until more specific immunosuppressive drugs are
identified.
Although no new drugs for ALD have successfully been devel- AH is associated not only with hepatocellular injury but
oped, many therapeutic targets were recently identified from also with impairments of liver regeneration. The application of
studies of human ALD samples and animal models, and some of hepatoprotective agents may provide some benefits in ALD
them are currently in clinical trials for the treatment of AH, therapy to protect against hepatocellular damage and promote
which are discussed below. liver regeneration. Indeed, the hepatoprotective cytokine IL‐22
Inflammation is considered a critical factor for causing liver is currently being tested in clinical trials for the treatment of
damage in AH and has been actively investigated as a therapeu- AH. IL‐22, which induces STAT3 activation in hepatocytes, has
tic target for the treatment of AH [64]. For instance, immuno- many ­beneficial effects in the liver [67] but likely has minimal
suppressive drug steroids, such as prednisolone, have been used side‐effects due to the restricted expression of IL‐22 receptors,
for the treatment of AH for more than five decades, but accumu- which are mainly expressed in epithelial cells but not in immune
lating data suggest that prednisolone treatment is beneficial for cells. By targeting hepatocytes, IL‐22 plays an important role in
short‐term survival but has no benefits for long‐term survival in ameliorating hepatocellular damage, promoting liver regenera-
AH patients [65]. Steroid drugs have broad immunosuppressive tion, and alleviating liver fibrosis [67]. In addition, IL‐22 treat-
functions but are ineffective for a variety of neutrophil‐mediated ment may effectively impede bacterial infection and ameliorate
disorders (e.g. asthma, septic shock), which is, at least in part, kidney injury, two deleterious conditions that often contribute to
mediated by promoting neutrophil survival. AH is associated death in AH patients. IL‐22 therapy is currently being tested in
with hepatocyte necrosis and infiltration of neutrophils, which clinical trials for the treatment of patients with severe AH [68].
may contribute to the ineffectiveness of prednisolone treatment Because severe ALD and AH are associated with multiple
in some AH patients. In addition, steroid treatment is associated problems, including inflammation, hepatocellular damage, poor
with an increased risk of bacterial infection in patients with AH liver regeneration, and many complications, combination thera-
due to its broad immunosuppressive function. More specific tar- pies are likely to be needed to treat these severe disorders [68].
gets for inflammation are urgently needed for the treatment of These combination therapies include the inhibition of inflamma-
AH. Indeed, a large number of therapeutic targets for inflamma- tion, hepatoprotection and stimulation of liver regeneration, and
tion have been recently identified and include inflammatory organ‐specific support in multiple organ failure for critically ill
cytokines and mediators, chemokines and their receptors, gut patients. Liver transplantation is probably the only potentially
microbiota and bacterial products, and so on [66]. Some of them life‐saving treatment for many patients with end‐stage ALD and
are currently being tested in clinical trials for the treatment of multiple organ failure who do not respond to medical treatment.
AH, including IL‐1 inhibitors, apoptosis signal‐regulating Indeed, ALD now replaces hepatitis C (HCV) infection as the
kinase 1 (ASK1) inhibitors, LPS blockers, and probiotics [66]. leading indication for liver transplantation in the United States.
Preclinical studies demonstrated that IL‐1 plays an important Current treatment guidelines require a six‐month period of alco-
role in promoting ALD in animal models. Thus, IL‐1 inhibitors, hol abstinence; however, most patients with severe AH do not
which have been approved for the treatment of several types of survive six months. Thus, several centers in Europe and the
inflammatory disease due to their good safety profiles and low United States have started early liver transplantation with less
tendency for adverse side‐effects, are currently being tested in than six‐month abstinence, achieving excellent clinical out-
clinical trials for the treatment of AH. ASK1, a serine/threonine comes with great survival rates and low rates of alcohol relapse
signaling kinase, is activated by oxidative stress, and its activa- in highly selected patients [69]. Based on these data, the
tion leads to activation of p38 mitogen‐activated kinase (p38 American Gastroenterological Association Clinical Practices
MAPK) and c‐Jun N‐terminal kinase (JNK), playing a role in Update Committee gave best practice advice: “Patients with
promoting hepatic inflammation, apoptosis, and fibrosis. ASK1 severe AH, particularly those with a MELD score > 26 with good
inhibitors are currently being tested in clinical trials for several insight into their AUD and good social support should be referred
types of liver diseases including non‐alcoholic steatohepatitis for evaluation for liver transplantation, as the 90‐day mortality
and AH. In addition, a large number of studies have demon- rate is very high” [70]. However, this recommendation has not
strated that excessive alcohol consumption induces gut bacterial been broadly adopted yet by the majority of transplantation cent-
overgrowth, dysbiosis, and elevation of circulating LPS, which ers and is still challenged by several issues, such as a shortage of
contribute to liver inflammation and injury in AH. Several ongo- donor livers, competition for donor livers by other types of liver
ing clinical trials are targeting these factors to treat AH using diseases, alcohol relapse post transplantation, and potentially
probiotics, antibiotics, fecal microbiota transplantation, spontaneous recovery in these patients after abstinence.
53:  Alcoholic Liver Disease 691

NEUROBIOLOGY OF ALCOHOL USE of Mental Disorders, 5th edition (DSM‐5; American Psychiatric
DISORDER (AUD) Association, 2013). AUD and addiction in general have been
­heuristically framed as a three‐stage cycle: binge/intoxication,
withdrawal/negative affect, and preoccupation/anticipation
Definitions and conceptual framework (“craving”). These three stages r­epresent dysregulation in three
for the neurocircuitry of AUD functional domains (incentive salience/pathological habits, nega-
AUD can be defined as a chronically relapsing disorder that is tive emotional states, and executive function, respectively) and are
characterized by a compulsion to seek and take the drug (alcohol), mediated by three major neurocircuitry elements (basal ganglia,
loss of control in limiting drug (alcohol) intake, and the emergence extended amygdala, and prefrontal cortex, respectively;
of a negative emotional state (e.g. dysphoria, anxiety, irritability Figure  53.2). The three stages are conceptualized as interacting
[hyperkatifeia]), reflecting a motivational withdrawal syndrome, with each other, becoming more intense, and ultimately leading to
when access to the drug (alcohol) is prevented [71]. These key ele- the pathological state known as addiction [71].
ments embrace most of the symptoms of AUD as expressed in In AUD, a pattern of oral drug taking evolves that is often
moderate to severe AUD in the Diagnostic and Statistical Manual characterized by binges of alcohol intake that can be daily

Figure 53.2  Conceptual framework for the neurobiological basis of addiction. In the binge/intoxication stage, reinforcing effects of drugs may
engage reward neurotransmitters and associative mechanisms in the nucleus accumbens shell and core and then engage stimulus‐response habits
that depend on the dorsal striatum. In the withdrawal/negative affect stage, the negative emotional state of withdrawal may engage activation of the
extended amygdala. The extended amygdala is composed of several basal forebrain structures, including the bed nucleus of the stria terminalis,
central nucleus of the amygdala, and possibly a transition zone in the medial portion (or shell) of the nucleus accumbens. There are major projec-
tions from the extended amygdala to the hypothalamus and brainstem. The preoccupation/anticipation (craving) stage involves the processing of
conditioned reinforcement in the basolateral amygdala and the processing of contextual information by the hippocampus. Executive control depends
on the prefrontal cortex and includes the representation of contingencies, the representation of outcomes, and their value and subjective states (i.e.
craving and, presumably, feelings) associated with drugs. The subjective effects, termed “drug craving” in humans, involve activation of the orbito-
frontal and anterior cingulate cortices and temporal lobe, including the amygdala. ACC, anterior cingulate cortex; BNST, bed nucleus of the stria
terminalis; CeA, central nucleus of the amygdala; DS, dorsal striatum; dlPFC, dorsolateral prefrontal cortex; GP, globus pallidus; HPC, hippocam-
pus; NAC, nucleus accumbens; OFC, orbitofrontal cortex; Thal, thalamus; vlPFC, ventrolateral prefrontal cortex; vmPFC, ventromedial prefrontal
cortex. Modified from [125] and reproduced with permission of Springer Nature.
692 THE LIVER:  NEUROBIOLOGY OF ALCOHOL USE DISORDER (AUD)

episodes or prolonged days of heavy drinking and is character- individual moves from impulsivity to compulsivity, a shift occurs
ized by a severe emotional and somatic withdrawal syndrome. from positive reinforcement that drives the motivated behavior to
Many individuals with AUD continue with such a binge/with- negative reinforcement that drives the motivated behavior [72].
drawal pattern for extended periods of time, but some individu- Negative reinforcement can be defined as the process by which
als can evolve into an opioid addiction‐like situation in which the removal of an aversive stimulus (e.g. negative emotional state
they must have alcohol available at all times to avoid the nega- of  drug withdrawal) increases the probability of a response.
tive consequences of abstinence. Here, intense preoccupation Importantly, negative reinforcement is not punishment, although
with obtaining alcohol (craving) develops that is linked not only both involve an aversive stimulus. In punishment, the aversive stim-
to stimuli that are associated with obtaining the drug but also to ulus suppresses behavior, including drug taking (e.g. disulfiram
stimuli that are associated with the aversive motivational state of [Antabuse]). Negative reinforcement can perhaps be described in
withdrawal. A pattern ultimately develops in moderate to severe lay terms as reward via relief (i.e. relief reward), such as the removal
AUD where the drug often is taken to avoid the severe dysphoria of pain or in the case of AUD removal of the negative emotional
and discomfort of abstinence. state of acute withdrawal or protracted abstinence. Driving negative
The thesis argued here, derived largely from fundamental work reinforcement is a negative emotional state that is a common
­
in neurobiology, is that AUD is a brain neurocircuitry disorder ­presentation in most individuals with AUD during withdrawal and
and that neuroadaptations within specific motivational circuits protracted abstinence. The neurobiological substrates that ­underlie
play an important role in defining and perpetuating the disorder. the motivation to seek alcohol will be reviewed herein using the
The argument herein is that the excessive engagement of reward heuristic framework that is outlined above.
circuitry engages high incentive salience for cues and contexts that
are conditioned to drug seeking (binge/intoxication stage) and sets Neural substrates for the binge/intoxication
up a core deficit of lower reward function and greater activation of
brain stress systems (withdrawal/negative affect stage) and sig-
stage associated with AUD
nificant impairment in executive function, all of which contribute Alcohol at intoxicating doses has a wide but selective action on
to the compulsive drinking that is associated with AUD. neurotransmitter systems in the brain reward systems, based on
Drug addiction has generally been conceptualized as a ­disorder animal models of the acute reinforcing effects of alcohol, and
that involves elements of both impulsivity and ­ compulsivity. animal studies that use selective receptor antagonists for spe-
Collapsing the cycles of impulsivity and compulsivity yields a cific neurochemical systems, and human positron emission
composite addiction cycle that consists of three stages – preoccu­ tomography imaging studies. Multiple neurochemical systems
pation/anticipation, binge/intoxication, and withdrawal/negative are implicated in the acute reinforcing effects of alcohol, includ-
affect – in which impulsivity often dominates at the early stages and ing γ‐aminobutyric acid (GABA), opioid peptides, dopamine,
compulsivity dominates at the terminal stages (Figure 53.2). As an serotonin, and glutamate (Figure 53.3).

Figure 53.3  Schematic diagram describing the etiology of addiction based on an early drawing by Dr Loren Parsons. Notice that as positive
reinforcement fades as a motivating factor, negative reinforcement is conceptualized to gain importance as driving compulsive drug seeking.
Relapse is hypothesized to return one to the addictive process at any point in the cycle, and repeated binges–withdrawals–relapses often speed the
trajectory back to compulsive use, presumably through residual changes in neurocircuitry.
53:  Alcoholic Liver Disease 693

There has been significant work that shows, at least at the Within‐system neuroadaptations that
pharmacological level, a role for GABA in the intoxicating contribute to the compulsivity associated
effects of alcohol [73]. Systemic injections of GABAA receptor
antagonists or inverse agonists reverse the motor‐impairing
with the dark side of AUD
effects of alcohol, the anxiolytic‐like effects of alcohol, and One prominent hypothesis is that dopamine and opioid peptide
alcohol drinking [73]. Endogenous opioid peptide systems have reward/incentive motivational systems are compromised in cru-
long been hypothesized to play a role in the reinforcing effects cial phases of the addiction cycle, such as withdrawal and pro-
of alcohol. Naltrexone decreases alcohol drinking and self‐ tracted abstinence. The argument is that a decrease in dopamine
administration in a variety of animal models [73], and these and opioid peptide function is hypothesized to lead to lower
results led to the clinical use of naltrexone in reducing alcohol motivation for non‐drug‐related stimuli and greater sensitivity
consumption and preventing relapse. to cues that are associated with the abused drug (i.e. increase in
Significant evidence also supports a role for the mesolimbic incentive salience) [81]. Supporting this hypothesis, decreases
dopamine system in alcohol reinforcement. Alcohol self‐admin- in activity of the mesolimbic dopamine system and decreases in
istration increases extracellular levels of dopamine in the serotonergic neurotransmission in the nucleus accumbens have
nucleus accumbens in nondependent rats [74]. Such increases been observed during alcohol withdrawal in animal studies [73].
occur not only during the actual self‐administration session but Parallel to these studies is strong evidence that the firing of ven-
also precede the self‐administration session, possibly reflecting tral tegmental area dopamine neurons dramatically decreases
the incentive motivational properties of cues that are associated during acute withdrawal from alcohol [82].
with alcohol [74]. Incentive motivation (i.e. incentive‐salience) Imaging studies of humans with addiction have consistently
is anchored within the construct of conditioned reinforcement shown long‐lasting decreases in the numbers of dopamine D2
and is argued to be a phenomenon by which a previously neutral receptors in alcohol‐dependent subjects compared with controls
stimulus acquires incentive value through pairings with a drug [83]. Additionally, alcohol‐dependent subjects had dramatically
of abuse [75]. Systemic injections of dopamine receptor antago- lower dopamine release in the striatum in response to a pharma-
nists also decrease responses to alcohol. However, mesolimbic cological challenge with the stimulant drug methylphenidate
dopamine does not appear to be essential for the acute reinforc- [83]. These findings suggest an overall reduction of the sensitiv-
ing effects of alcohol, since lesions of the mesolimbic dopamine ity of the dopamine component of reward circuitry to natural
system failed to block operant self‐administration of alcohol reinforcers and other drugs in individuals with addiction.
[73], suggesting multiple redundant sources of alcohol’s actions Other within‐system neuroadaptations under this conceptual
that converge on the nucleus accumbens and amygdala. framework could include greater sensitivity of receptor trans-
A prominent hypothesis is that when drug addiction progresses duction mechanisms in the nucleus accumbens. Drugs of abuse
from occasional recreational use to compulsive use, drug‐seeking have acute receptor actions that are linked to intracellular sign-
behavior shifts from reward‐driven to habit‐driven behavior to aling pathways that may undergo adaptations with chronic treat-
compulsive‐like responding. Three cortico‐basal ganglia circuits ment. In the context of chronic alcohol administration, multiple
form what have been described as cortico‐striatal‐pallidal‐­thalamic molecular mechanisms have been hypothesized to counteract
loops that process associative, sensorimotor, and emotional infor- the acute effects of alcohol that could be considered within‐­
mation [76]. During this progression to habit‐driven behavior, the system neuroadaptations. For example, chronic alcohol
control over drug‐seeking behavior also appears to shift from decreases GABA receptor function, possibly through downreg-
the nucleus accumbens to the dorsal striatum. Furthermore, within ulation of the α1 subunit, and also decreases the acute inhibition
the dorsal striatum, there is a shift in function [77]. of adenosine reuptake [73]. Whereas acute alcohol activates
adenylate cyclase, withdrawal from chronic alcohol decreases
cyclic adenosine monophosphate response element binding pro-
Neural substrates for the withdrawal/negative tein phosphorylation in the amygdala and is linked to decreases
affect stage associated with AUD in the function of neuropeptide Y (NPY) and increases in anxi-
ety‐like responses during acute alcohol withdrawal [84].
The neurocircuitry and neuropharmacology of the withdrawal/
negative affect stage of the addiction cycle support a conceptual
framework that builds on opponent process theory [78] and Between‐system neuroadaptations that
extends to an allostatic model of brain motivational systems that contribute to the compulsivity associated
has been proposed to explain persistent changes in motivation
that are associated with dependence in addiction [79, 80]. In this
with the dark side of AUD
formulation, addiction is conceptualized as a cycle of increasing Another major neuroadaptation that can contribute to the nega-
dysregulation of brain reward/anti‐reward mechanisms that tive emotional state that drives negative reinforcement in the
results in a negative emotional state that contributes to the withdrawal/negative affect stage is brain neurocircuits and
­compulsive use of drugs. Counteradaptive processes that are ­neurochemical systems that are involved in arousal–stress mod-
part of the normal homeostatic limitation of reward function ulation that are engaged within the neurocircuitry of the brain
fail to return within the normal homeostatic range. These coun- stress systems in an attempt to overcome the chronic presence of
teradaptive processes are hypothesized to be mediated by the perturbing drug (alcohol) and to restore normal function
two mechanisms: within‐system neuroadaptations and between‐­ despite the presence of drug. The neuroanatomical entity that is
system neuroadaptations [80]. termed the extended amygdala [85] may represent a common
694 THE LIVER:  NEUROBIOLOGY OF ALCOHOL USE DISORDER (AUD)

anatomical substrate that integrates brain arousal–stress systems Similar effects have been observed in animal models, with a
with hedonic processing systems to produce the between‐­ blunted corticosterone response in rats that are made dependent
system opponent process that is elaborated above. The extended using the chronic intermittent alcohol vapor model [93]. One
amygdala forms a separate entity within the basal forebrain [86] hypothesis is that activation of the HPA axis can drive neuroad-
and has been conceptualized to be composed of several basal aptive changes in extrahypothalamic CRF systems in the
forebrain structures, including the bed nucleus of the stria termi- extended amygdala. High corticosterone increases CRF mRNA
nalis, central nucleus of the amygdala, sublenticular substantia in the central nucleus of the amygdala and lateral bed nucleus of
innominata, and a transition zone in the medial part of the the stria terminalis and decreases CRF mRNA in the paraven-
nucleus accumbens (e.g. shell) [85]. The extended amygdala tricular nucleus of the hypothalamus. Thus, an initial exposure
receives numerous afferents from limbic structures, such as the to high corticosterone, stimulated by moderate to heavy drink-
basolateral amygdala and hippocampus, and sends efferents to ing, may stimulate CRF expression in the central nucleus of the
the medial part of the ventral pallidum and a large projection to amygdala and lateral bed nucleus of the stria terminalis, eventu-
the lateral hypothalamus, thus further defining the specific brain ally leading to neuroadaptive changes, including the further sen-
areas that interface classical limbic (emotional) structures with sitization of CRF activation in the extended amygdala and lower
the extrapyramidal motor system. The extended amygdala HPA function [93]. Consistent with this hypothesis, chronic
has long been hypothesized to play a key role not only in fear ­glucocorticoid receptor blockade with mifepristone during the
conditioning [87] but also in the emotional component of pain course of alcohol vapor exposure prevented the escalation of
processing [88]. alcohol intake and blocked the increase in progressive‐ratio
The brain stress system that is mediated by corticotropin‐ responding for alcohol in dependent animals [94]. Chronic glu-
releasing factor (CRF) systems in both the extended amygdala cocorticoid receptor antagonism also blocked escalated and
and hypothalamic–pituitary–adrenal axis (HPA) are dysregu- compulsive alcohol drinking during protracted abstinence in
lated by the chronic administration of all major drugs with rats with a history of alcohol dependence. These results suggest
dependence or abuse potential, with a common response of a critical role for glucocorticoid receptors in the development
­elevated adrenocorticotropic hormone, corticosterone, and and maintenance of alcohol dependence.
amygdala CRF during acute withdrawal from chronic drug Other brain neurotransmitter or neuromodulatory systems
administration [80] (Figure 53.3). In animal models of alcohol that have pro‐stress actions also converge on the extended
dependence, extrahypothalamic CRF systems become hyperac- amygdala, all of which may contribute to negative emotional
tive during alcohol withdrawal, with an increase in extracellular states that are associated with drug withdrawal or protracted
CRF in the central nucleus of the amygdala and bed nucleus of abstinence [80]. Chronic administration of psychostimulants
the stria terminalis in dependent rats [80]. For example, alcohol and opioids has long been hypothesized to activate cyclic aden-
withdrawal reliably produces anxiety‐like responses in animal osine monophosphate response element binding protein, which
models that can be reversed by CRF receptor antagonists in turn activates dynorphin in medium spiny neurons that in turn
[80].  Indeed, using models of repeated alcohol exposure and feedback and decrease the activity of ventral tegmental dopa-
­withdrawal, intermittent alcohol exposure facilitates withdrawal‐ mine neurons [95]. Although κ‐opioid receptor agonists sup-
associated anxiety‐like responses, and a CRF1 receptor antago- press nondependent drinking, presumably via aversive stimulus
nist prevented the sensitization of withdrawal‐induced anxiety effects, κ‐opioid receptor antagonists block the excessive drink-
[89]. These results are consistent with a prolonged history of ing associated with alcohol withdrawal and dependence, and
alcohol exposure that produces persistent upregulation of both this effect may be mediated by the shell of the nucleus accum-
CRF and CRF1 receptors in the brain [89]. bens [73].  Other neurotransmitter/neuromodulatory systems
Perhaps even more compelling, a peptide CRF1/CRF2 antago- that comprise the brain stress system in the extended amyg-
nist, when administered directly in the central nucleus of the dala include ­vasopressin, hypocretin (orexin), substance P, and
amygdala, blocked alcohol self‐administration in alcohol‐ ­neuroimmune factors. In addition to CRF and dynorphin, there
dependent rats. Cellular studies have identified the actions of is evidence that norepinephrine, vasopressin, substance P, and
CRF on GABAergic interneurons in the central nucleus of the hypocretin (orexin) may all contribute to negative emotional
amygdala [90]. Systemic injections of small‐molecule CRF1 states of drug withdrawal, particularly alcohol withdrawal
receptor antagonists also blocked the increase in alcohol intake (Figure  53.3). Thus, activation of this pro‐stress, pro‐negative
that was associated with acute withdrawal and protracted absti- emotional state system is multi‐determined and comprises the
nence. A CRF receptor antagonist that was administered chroni- neurochemical bases for hedonic opponent processes [96].
cally during the development of dependence blocked the Neurotransmitter systems that are implicated in anti‐stress
development of compulsive‐like responding for alcohol (for actions include NPY, nociceptin, and endocannabinoids. NPY
review, see reference [89]). has powerful orexigenic and anxiolytic effects and has been
Alcohol addiction has long been associated with dysregula- hypothesized to act in opposition to the actions of CRF in
tion of the HPA axis [91]. Clinical studies have reported impair- addiction [84]. Nociceptin and synthetic ­nociception receptor
ments in HPA axis responsivity in AUD [91], even to the point agonists have effects on GABA synaptic activity in the central
of pseudo‐Cushing’s syndrome, manifested by high levels of nucleus of the amygdala that are similar to NPY and can block
cortisol, in individuals with alcohol addiction [92]. However, a high alcohol consumption in a genetically selected line of rats
more common observation is a blunted cortisol response in indi- that is known to be hypersensitive to s­ tressors [97]. Evidence
viduals with alcohol dependence, again possibly reflecting also implicates endocannabinoids in the regulation of affective
adaptations to an initial hyperactive cortisol response [91]. states, in which reductions of cannabinoid CB1 receptor
53:  Alcoholic Liver Disease 695

signaling produce anxiogenic‐like behavioral effects [98]. hypothesized to reflect the neural representations of reward and
Thus, vulnerability to AUD may involve not only a sensitized incentive salience constructs [104]. One parsimonious view of
stress system but also a hypoactive stress buffer system, and the human imaging data that is consistent with animal model
behavioral and pharmacological interventions that block pro‐ data is that there is a “go” system in the dorsal prefrontal/­
stress and stimulate anti‐stress (described in Neural substrates cingulate cortex that drives impulsivity and craving and a “stop”
for the preoccupation/anticipation stage associated with AUD) system in ventromedial prefrontal cortex that inhibits impulsiv-
systems may be particularly interesting targets for the treatment ity and craving [105]. In humans, stress and stressors have long
of AUD. been associated with relapse and the vulnerability to relapse
In summary, two processes are hypothesized to form the neu- [106]. Individuals with addiction are hypersensitive to pain dur-
robiological basis for the withdrawal/negative affect stage: loss ing withdrawal, particularly in the face of negative affect [107].
of function in the reward systems (within‐system neuroadapta- Indeed, the leading precipitant of relapse is negative emotion/
tion) and recruitment of the brain stress systems (between‐­ affect including elements of anger, frustration, sadness, anxiety,
system neuroadaptation) [71]. As dependence and withdrawal and guilt [108].
develop, brain stress systems, such as CRF, norepinephrine, and Studies of “craving” in animal models can be divided into
dynorphin, are recruited, producing aversive or stress‐like states three domains: alcohol seeking that is induced by the drug itself,
[99]. The combination of decreases in reward neurotransmitter alcohol priming‐induced reinstatement, alcohol seeking that is
function and recruitment of brain stress systems provides a induced by stimuli that are paired with drug taking, cue‐ and
powerful motivation for reengaging in drug taking and drug context‐induced reinstatement, and alcohol seeking that is
seeking. induced by an acute stressor or a state of stress (i.e. stress‐
induced reinstatement) [109]. Most evidence from animal stud-
Neural substrates for the preoccupation/ ies suggests that drug‐induced reinstatement is localized to the
medial prefrontal cortex/ventral striatum circuit and mediated
anticipation stage associated with AUD
by the neurotransmitter glutamate [110]. Neurotransmitter sys-
The preoccupation/anticipation or “craving” stage of the addic- tems that are involved in drug‐induced reinstatement involve a
tion cycle has long been hypothesized to be a key part of the glutamatergic projection from the frontal cortex to the nucleus
neurocircuitry that mediates relapse in humans. Dysregulation accumbens that is modulated by dopamine activity in the frontal
of the frontal cortex mediates not only elements of impulsivity cortex (Figure  53.3). In contrast, neuropharmacological and
and compulsivity but also protracted abstinence and craving. neurobiological studies that use animal models of cue‐induced
Human imaging studies reveal neurocircuitry dysregulation reinstatement involve the basolateral amygdala as a critical sub-
during the preoccupation/anticipation stage in AUD that strate, with a possible feed‐forward mechanism possibly through
includes not only compromises in frontal cortical executive the same prefrontal cortex system that is involved in drug‐
function but also dysregulated substrates that mediate craving. induced reinstatement [109, 111]. Cue‐induced reinstatement
Lower frontal cortex activity parallels deficits in executive func- involves dopamine modulation in the basolateral amygdala and
tion in neuropsychologically challenging tasks in individuals a glutamatergic projection to the nucleus accumbens from both
with AUD with and without Wernicke–Korsakoff’s syndrome the basolateral amygdala and ventral subiculum [111]. Such
(for review, see references [100, 101]). In individuals with AUD, reinstatement in alcohol drinking can be blocked by the sys-
there are impairments in the maintenance of spatial information, temic administration of naltrexone and a selective μ and δ
the disruption of decision making, and impairments in behavio- ­opioid receptor antagonist [109]. These results are consistent
ral inhibition. Such frontal cortex‐derived executive function with human studies that showed that opioid receptor antagonists
disorders in AUD have been linked to deficits in the ability of may blunt the urge to drink that is elicited by the presentation of
behavioral treatments to effect recovery from AUD post‐­ alcohol‐related cues in individuals with AUD [112]. In contrast,
detoxification [100]. Thus, deficits in the prefrontal cortical the stress‐induced reinstatement of drug‐related responding in
control of incentive salience may represent a key mechanism to animal models appears to depend on the activation of both CRF
explain individual differences in the vulnerability to AUD, and and norepinephrine in elements of the extended amygdala
the excessive attribution of incentive salience to drug‐related (­central nucleus of the amygdala and bed nucleus of the stria
cues and residual hypersensitivity of the brain stress systems terminalis) [113].
may perpetuate excessive drug intake, compulsive behavior,
and relapse. Compulsivity in AUD: a negative
Craving responses to cues in human imaging studies activate
an overarching cognitive control network in the brain, involving
reinforcement perspective
dorsolateral prefrontal, anterior cingulate, and parietal cortices, Compulsivity in AUD can derive from neurocircuitry changes at
all of which support a broad range of executive functions [102]. all three stages of the addiction cycle (Figure 53.3). In the binge/
Such studies have led to the hypothesis that a frontal–cingulate– intoxication stage, such changes may involve neurocircuits that
parietal–subcortical cognitive control network is consistently include enhanced incentive salience [75] and the engagement of
recruited across a range of traditional executive function tasks, pathological habit function [114]. In the withdrawal/negative
many of which show deficits in humans with AUDs. Indeed, affect stage, such changes involve the recruitment of neurocir-
the most prominent activation by alcohol‐related cues involves cuits that are involved in negative emotional states. In the preoc­
the dorsolateral prefrontal cortex, cingulate cortex, and orbito- cupation/anticipation stage, such changes involve impairments
frontal cortex [103]. Such drug cue‐evoked responses are in executive function and the dysregulation of inhibitory control
696 THE LIVER:  TREATMENT OF AUD PATIENTS WITH ALD

that regulates impulsivity [115]. Many of the cue‐ and stress‐ patients with ALD are very limited. Lieber et  al. [118] con-
related neuroadaptations persist into protracted abstinence, ducted a clinical pharmacology study of liver fibrosis and
largely described as dysregulation that persists beyond acute showed that the addition of a brief intervention resulted in a
withdrawal [73]. significant reduction in alcohol consumption. Other studies
However, negative emotional states of acute and protracted have investigated behavioral platforms, like CBT and MET, as
abstinence that are induced by chronic high‐dose alcohol have ways to facilitate alcohol abstinence in patients with ALD. A
been largely neglected in the clinical domain. Negative emo- systematic review of these studies was conducted by Khan et al.
tional states may strongly impact compulsivity not only by [119]. A total of 13 papers were selected for a whole sample of
directly driving negative reinforcement but also by enhancing 1945 patients. This systematic review indicated that combining
the value of incentive salience, enhancing the value of habit medical care and behavioral approaches, like CBT and MET,
expression, or exacerbating impairments in executive function. facilitated alcohol abstinence, a conclusion that supports the
Thus, the overall conceptual theme that is argued herein is that importance of multidisciplinary approaches to treat patients
moderate to severe AUD represents a break with homeostatic with AUD and ALD [119].
brain regulatory mechanisms that regulate the emotional state of
the individual. One hypothesis is that the dysregulation of
­emotion begins with the binge and subsequent acute withdrawal
Medications
but leaves a residual neuroadaptive trace that allows rapid Pharmacological approaches to treat AUD patients include med-
“re‐addiction” even months and years after detoxification and ications to treat alcohol withdrawal syndrome (AWS) and those
abstinence [79]. Thus, the emotional dysregulation of alcohol used to help patients reduce their craving, cut their drinking,
addiction represents more than the simple homeostatic dysregu- facilitate abstinence, and prevent relapse. As for the treatment of
lation of hedonic function; it also represents a dynamic break AWS, it is important to keep in mind that while up to 50% of
with the homeostasis of this system that has been termed individuals with AUD present with alcohol withdrawal symp-
allostasis [79]. A further argument is that this hypernegative toms after they stop drinking, only a small percentage requires
emotional state, termed hyperkatifeia, sensitizes over time, pharmacological treatment. For these patients, benzodiazepines
extends into protracted abstinence, and provides the driving are the gold standard because they represent the only class of
force for another source of motivation for prolonging and main- medications that reduces the risk of withdrawal seizures and/or
taining addiction, that of negative reinforcement (Figure 53.3). delirium tremens [120]. In those individuals with AUD and
ALD who develop AWS, a symptom‐triggered schedule with
lorazepam or oxazepam is preferred. In fact, these benzodiaz-
epines do not undergo phase I biotransformation; rather, they
TREATMENT OF AUD PATIENTS only undergo glucuronidation, which is preserved even if liver
WITH ALD function is compromised [116].
A better understanding of the neurobiology of addiction,
While reductions below harmful drinking levels is beneficial, reviewed above in this chapter, has played an important role in
the cornerstone in the treatment of patients with ALD remains the development of medications for AUD, as summarized in
total alcohol abstinence, as abstinence can improve outcome at Table 53.1. In the United States, acamprosate, disulfiram, and
all stages of liver disease [116]. However, there are only a few naltrexone (oral and intramuscular) are approved by the Food
­randomized studies that have investigated behavioral and/or and Drug Administration (FDA) for treatment of AUD. A recent
pharmacological treatments in patients with AUD and ALD. meta‐analysis supports the efficacy of naltrexone and acampro-
sate, but not disulfiram, for AUD [121]. Furthermore, nalmefene
was recently approved in Europe for the treatment of AUD, but
Behavioral treatments
it is not approved in the United States. Not only is the number of
Brief interventions are aimed at educating the patient about approved medications for AUD very limited, but also their use
harmful drinking, increasing motivation to change behavior, and in patients with ALD is even more narrow. In fact, disulfiram
reinforcing skills to address problematic drinking. In primary may cause hepatotoxicity and is not recommended in patients
care settings, studies support the use of brief interventions to with ALD. Naltrexone’s potential for causing hepatotoxicity
reduce drinking [117]. Furthermore, these interventions may also exists, even if it seems to be rare. Acamprosate has not for-
synergize with other areas of treatment, such as, compliance to mally been tested in AUD patients with liver disease, but it does
medications and referral to treatment programs, an approach not undergo hepatic metabolism and there are no reports of
referred to as screening, brief intervention and referral to treat­ hepatotoxicity [116].
ment (SBIRT). Specific psychosocial and behavioral treatments In the past decades, preclinical and clinical studies have
for AUD include 12‐step facilitation, cognitive behavioral ther- ­provided support for some medications as potential novel treat-
apy (CBT), and motivational enhancement therapy (MET) ments for AUD. While none of these medications are FDA‐
[116]. Twelve‐step facilitation focuses on abstinence and approved, some of them have shown efficacy for AUD in phase
involves Alcoholics Anonymous meetings. The goal of CBT is 2/3 trials. Among them, the most promising are baclofen,
to develop coping mechanisms where alcohol is replaced by gabapentin, ondansetron, topiramate, and varenicline, as sum-
alcohol‐free circumstances. MET helps patients work through marized in Table  53.1. However, formal clinical trials testing
the dilemma by “rolling with the resistance” to change [116]. these medications in AUD patients with ALD are lacking, except
Specific studies investigating behavioral treatments for AUD in for baclofen. In fact, while general clinical trials testing baclofen
53:  Alcoholic Liver Disease 697

Table 53.1  FDA‐approved medications and others tested in alcohol use disorder patientsa
Dosage Pharmacological target Possible use in alcohol use disorder patients with alcoholic liver disease?

FDA‐approved medications for alcohol use disorder


Acamprosate 666 mg TID Possibly NMDA receptor agonist Yes (no hepatic metabolism)
Disulfiram 250–500 mg QD Inhibition of acetaldehyde No (hepatic metabolism; cases of liver toxicity have been reported)
dehydrogenase
Naltrexoneb PO: 50 mg QD μ opioid receptor antagonist With caution (perceptions of liver toxicity limit use in advanced
PO or IM IM: 380 mg monthly alcoholic liver disease)
Not FDA‐approved medications tested for alcohol use disorder
Baclofen 10 mg TID; GABAB receptor agonist Yes (minimal hepatic metabolism)
80 mg QD max Baclofen has been formally tested in clinical studies with alcohol use
disorder patients with liver cirrhosis
Gabapentin 900–1800 mg QD Unclear – the most likely Yes (no hepatic metabolism)
mechanism is blockade of
voltage‐dependent Ca2+ channels.
Ondansetron 1–16 mcg /kg−1 BID 5HT3 antagonist Yes, but with caution because liver toxicity has been reported, albeit
relationship to ondansetron administration is not determined
Topiramate 300 mg QD Anticonvulsant multiple targets: Yes (partial hepatic metabolism mostly by glucuronidation)
−Glutamate/+GABA In patients with hepatic encephalopathy, use with caution: topiramate‐
related cognitive side‐effects may confound the clinical course and
treatment of hepatic encephalopathy
Varenicline 2 mg QD Nicotinic acetylcholine receptor Yes (minimal hepatic metabolism)
partial agonist
a
 Reproduced from [116] with permission of Elsevier.
b
 Nalmefene is not‐FDA approved but was recently approved in Europe for alcohol use disorder. Compared to naltrexone, nalmefene has a longer half‐life and no evidence of
hepatotoxicity.
FDA, Food and Drug Administration; TID, three times a day; NMDA, N‐methyl‐D‐aspartate; QD, once a day; PO, per os (oral); IM, intramuscular; GABA, gamma‐­
aminobutyric acid; BID, twice a day; HT, serotonin.

have generated conflicting results [122], two independent clini- 4. Mandrekar, P., Bataller, R., Tsukamoto, H. et al. Alcoholic hepatitis: transla-
cal trials testing baclofen in AUD patients with ALD both tional approaches to develop targeted therapies. Hepatology, 2016;64:
1343–55.
­support its efficacy in treating this specific subpopulation [123, 5. Chang, B., Xu, M.J., Zhou, Z. et al. Short‐ or long‐term high‐fat diet feeding
124]. However, the mechanisms by which baclofen may be effi- plus acute ethanol binge synergistically induce acute liver injury in mice: an
cacious, specifically in AUD patients with ALD, are unknown. important role for CXCL1. Hepatology, 2015;62:1070–85.
Furthermore, although not formally tested in patients with ALD, 6. Xu, J., Lai, K.K., Verlinsky, A. et al. Synergistic steatohepatitis by moderate
obesity and alcohol in mice despite increased adiponectin and p‐AMPK.
the other medications mentioned above (especially gabapentin
J Hepatol, 2011;55:673–82.
and varenicline), might also be useful in AUD individuals with 7. You, M., Considine, R.V., Leone, T.C. et  al. Role of adiponectin in the
ALD because they have no evidence of liver toxicity (for addi- ­protective action of dietary saturated fat against alcoholic fatty liver in mice.
tional details, please see Table 53.1). Hepatology, 2005;42:568–77.
In conclusion, the crucial goal in treating patients with AUD 8. Bertola, A., Park, O., and Gao, B. Chronic plus binge ethanol feeding syner-
gistically induces neutrophil infiltration and liver injury in mice: a critical
and ALD is to help them to achieve and maintain alcohol absti-
role for E‐selectin. Hepatology, 2013;58:1814–23.
nence. As briefly summarized above, both behavioral and phar- 9. Xu, M.J., Cai, Y., Wang, H. et al. Fat‐specific protein 27/CIDEC promotes
macological treatments are available and may be effective in development of alcoholic steatohepatitis in mice and humans.
helping AUD patients to quit drinking. These treatments remain Gastroenterology, 2015;149:1030–41.
underutilized, and it is therefore very important that their use is 10. Gao, B., Xu, M.J., Bertola, A. et al. Animal models of alcoholic liver disease:
pathogenesis and clinical relevance. Gene Expr, 2017;17:173–86.
expanded in clinical care, including in hepatology settings 11. Cai, Y., Xu, M.J., Koritzinsky, E.H. et  al. Mitochondrial DNA‐enriched
where many patients are referred for their ALD. Therefore, it is microparticles promote acute‐on‐chronic alcoholic neutrophilia and hepato-
critical that addiction medicine is integrated into treatment and toxicity. JCI Insight, 2017;2.
prescribed by the multidisciplinary teams that care for patients 12. Li, M., He, Y., Zhou, Z. et  al. MicroRNA‐223 ameliorates alcoholic liver
with AUD and ALD. injury by inhibiting the IL‐6‐p47phox‐oxidative stress pathway in neutro-
phils. Gut, 2017;66:705–15.
13. Hatton, J., Burton, A., Nash, H. et  al. Drinking patterns, dependency and
life‐time drinking history in alcohol‐related liver disease. Addiction,
2009;104:587–92.
14. Askgaard, G., Gronbaek, M., Kjaer, M.S. et al. Alcohol drinking pattern and
REFERENCES risk of  alcoholic liver cirrhosis: a prospective cohort study. J Hepatol,
2015;62:1061–7.
1. Gao, B. and Bataller, R. Alcoholic liver disease: pathogenesis and new thera- 15. Trepo, E., Romeo, S., Zucman‐Rossi, J. et al. PNPLA3 gene in liver d­ iseases.
peutic targets. Gastroenterology, 2011;141:1572–85. J Hepatol, 2016;65:399–412.
2. O’Shea, R.S., Dasarathy, S., McCullough, A.J. et al. Alcoholic liver disease. 16. Ajmera, V.H., Terrault, N.A., and Harrison, S.A. Is moderate alcohol use in
Hepatology, 2010;51:307–28. nonalcoholic fatty liver disease good or bad? A critical review. Hepatology,
3. Zakhari, S. and Li, T.K. Determinants of alcohol use and abuse: impact 2017;65:2090–9.
of  quantity and frequency patterns on liver disease. Hepatology, 2007; 17. Purohit, V., Gao, B., Song, B.J. Molecular mechanisms of alcoholic fatty
46;2032–9. liver. Alcohol Clin Exp Res, 2009;33:191–205.
698 THE LIVER:  REFERENCES

18. Parker, R., Kim, S.J., and Gao, B. Alcohol, adipose tissue and liver disease: 43. Szabo, G. and Momen‐Heravi, F. Extracellular vesicles in liver disease and
mechanistic links and clinical considerations. Nat Rev Gastroenterol potential as biomarkers and therapeutic targets. Nat Rev Gastroenterol
Hepatol, 2018;15:50–9. Hepatol, 2017;14:455–66.
19. Zhong, W., Zhao, Y., Tang, Y. et  al. Chronic alcohol exposure stimulates 44. Saha, B., Momen‐Heravi, F., Furi, I. et al. Extracellular vesicles from mice
­adipose tissue lipolysis in mice: role of reverse triglyceride transport in the with alcoholic liver disease carry a distinct protein cargo and induce
pathogenesis of alcoholic steatosis. Am J Pathol, 2012;180:998–1007. ­macrophage activation via heat shock protein 90. Hepatology, 2018;67(5):
20. Sebastian, B.M., Roychowdhury, S., Tang, H. et  al. Identification of a 1986–2000.
cytochrome P4502E1/Bid/C1q‐dependent axis mediating inflammation in 45. Sutti, S., Bruzzi, S., and Albano, E. The role of immune mechanisms in
adipose tissue after chronic ethanol feeding to mice. J Biol Chem, alcoholic and nonalcoholic steatohepatitis: a 2015 update. Expert Rev
­
2011;286:35989–97. Gastroenterol Hepatol, 2016;10:243–53.
21. Ji, C., Chan, C., and Kaplowitz, N. Predominant role of sterol response ele- 46. Liaskou, E., Klemsdal Henriksen, E.K., Holm, K. et  al. High‐throughput
ment binding proteins (SREBP) lipogenic pathways in hepatic steatosis in T‐cell receptor sequencing across chronic liver diseases reveals distinct
the murine intragastric ethanol feeding model. J Hepatol, 2006;45:717–24. ­disease‐associated repertoires. Hepatology, 2016;63:1608–19.
22. You, M., Fischer, M., Deeg, M.A. et al. Ethanol induces fatty acid synthesis 47. Mathews, S., Feng, D., Maricic, I. et al. Invariant natural killer T cells con-
pathways by activation of sterol regulatory element‐binding protein tribute to chronic‐plus‐binge ethanol‐mediated liver injury by promoting
(SREBP). J Biol Chem, 2002;277:29342–7. hepatic neutrophil infiltration. Cell Mol Immunol, 2016;13:206–16.
23. Baraona, E. and Lieber, C.S. Effects of ethanol on lipid metabolism. J Lipid 48. Cui, K., Yan, G., Xu, C. et al. Invariant NKT cells promote alcohol‐induced
Res, 1979;20:289–315. steatohepatitis through interleukin‐1beta in mice. J Hepatol, 2015;
24. Galli, A., Pinaire, J., Fischer, M. et al. The transcriptional and DNA binding 62:1311–8.
activity of peroxisome proliferator‐activated receptor alpha is inhibited by 49. Riva, A., Patel, V., Kurioka, A. et al. Mucosa‐associated invariant T cells link
ethanol metabolism. A novel mechanism for the development of ethanol‐ intestinal immunity with antibacterial immune defects in alcoholic liver
induced fatty liver. J Biol Chem, 2001;276:68–75. ­disease. Gut, 2017;67(5):918–30.
25. O’Neill, H.M., Holloway, G.P., and Steinberg, G.R. AMPK regulation of 50. Bertola, A., Mathews, S., Ki, S.H. et al. Mouse model of chronic and binge
fatty acid metabolism and mitochondrial biogenesis: implications for ethanol feeding (the NIAAA model). Nat Protoc, 2013;8:627–37.
­obesity. Mol Cell Endocrinol, 2013;366:135–51. 51. Chao, X., Wang, S., Zhao, K. et  al. Impaired TFEB‐mediated lysosome
26. You, M., Matsumoto, M., Pacold, C.M. et  al. The role of AMP‐activated ­biogenesis and autophagy promote chronic ethanol‐induced liver injury and
protein kinase in the action of ethanol in the liver. Gastroenterology, steatosis in mice. Gastroenterology, 2018;155(3):865–79.
2004;127:1798–1808. 52. Wang, W., Xu, M.J., Cai, Y. et al. Inflammation is independent of steatosis in
27. Dolganiuc, A., Thomes, P.G., Ding, W.X. et  al. Autophagy in alcohol‐ a murine model of steatohepatitis. Hepatology, 2017;66:108–23.
induced liver diseases. Alcohol Clin Exp Res, 2012;36:1301–8. 53. Khanova, E., Wu, R., Wang, W. et al. Pyroptosis by caspase11/4‐gasdermin‐
28. Ding, W.X., Li, M., Chen, X. et al. Autophagy reduces acute ethanol‐induced D pathway in alcoholic hepatitis. Hepatology, 2017;67(5):1737–53.
hepatotoxicity and steatosis in mice. Gastroenterology, 2010;139:1740–52. 54. European Association for the Study of Liver. EASL clinical practical guide-
29. Nagy, L.E., Ding, W.X., Cresci, G. et al. Linking pathogenic mechanisms lines: management of alcoholic liver disease. J Hepatol, 2012;57:399–420.
of  alcoholic liver disease with clinical phenotypes. Gastroenterology, 55. Alatalo, P., Koivisto, H., Puukka, K. et  al. Biomarkers of liver status in
2016;150:1756–68. heavy  drinkers, moderate drinkers and abstainers. Alcohol Alcohol,
30. Gao, B. and Tsukamoto, H. Inflammation in alcoholic and nonalcoholic fatty 2009;44:199–203.
liver disease: friend or foe? Gastroenterology, 2016;150:1704–9. 56. Aday, A.W., Mitchell, M.C., Casey, L.C. Alcoholic hepatitis: current trends
31. Hartmann, P., Seebauer, C.T., and Schnabl, B. Alcoholic liver disease: the in management. Curr Opin Gastroenterol, 2017;33:142–8.
gut microbiome and liver cross talk. Alcohol Clin Exp Res, 2015;39: 57. Ku, N.O., Strnad, P., Bantel, H. et al. Keratins: Biomarkers and modulators
763–75. of apoptotic and necrotic cell death in the liver. Hepatology, 2016;
32. Lucey, M.R., Mathurin, P., and Morgan, T.R. Alcoholic hepatitis. N Engl J 64:966–76.
Med, 2009;360:2758–69. 58. Mortele, K.J. and Ros, P.R. Imaging of diffuse liver disease. Semin Liver Dis,
33. Crabb, D.W., Bataller, R., Chalasani, N.P. et  al. Standard definitions and 2001;21:195–212.
common data elements for clinical trials in patients with alcoholic hepatitis: 59. Borra, R.J., Salo, S., Dean, K. et al. Nonalcoholic fatty liver disease: rapid
recommendation from the NIAAA alcoholic hepatitis consortia. evaluation of liver fat content with in‐phase and out‐of‐phase MR imaging.
Gastroenterology, 2016;150:785–90. Radiology, 2009;250:130–6.
34. Affo, S., Dominguez, M., Lozano, J.J. et al. Transcriptome analysis identi- 60. Stickel, F., Datz, C., Hampe, J. et al. Pathophysiology and management of
fies TNF superfamily receptors as potential therapeutic targets in alcoholic alcoholic liver disease: update 2016. Gut Liver, 2017;11:173–88.
hepatitis. Gut, 2013;62:452–60. 61. Parker, R., Armstrong, M.J., Corbett, C. et al. Systematic review: pentoxifyl-
35. Michelena, J., Altamirano, J., Abraldes, J.G. et al. Systemic inflammatory line for the treatment of severe alcoholic hepatitis. Aliment Pharmacol Ther,
response and serum lipopolysaccharide levels predict multiple organ failure 2013;37:845–54.
and death in alcoholic hepatitis. Hepatology, 2015;62:762–72. 62. Nguyen‐Khac, E., Thevenot, T., Piquet, M.A. et  al. Glucocorticoids plus
36. Louvet, A., Wartel, F., Castel, H. et al. Infection in patients with severe alco- N‐acetylcysteine in severe alcoholic hepatitis. N Engl J Med, 2011
holic hepatitis treated with steroids: early response to therapy is the key ;365:1781–9.
­factor. Gastroenterology, 2009;137:541–8. 63. Piechota, M. and Piechota, A. An evaluation of the usefulness of extracorpor-
37. Bataller, R. and Gao, B. Liver fibrosis in alcoholic liver disease. Semin Liver eal liver support techniques in patients hospitalized in the ICU for severe
Dis, 2015;35:146–56. liver dysfunction secondary to alcoholic liver disease. Hepat Mon,
38. Jeong, W.I., Park, O., and Gao, B. Abrogation of the antifibrotic effects of 2016;16:e34127.
natural killer cells/interferon‐gamma contributes to alcohol acceleration 64. Wang, H.J., Gao, B., Zakhari, S. et  al. Inflammation in alcoholic liver
of liver fibrosis. Gastroenterology, 2008;134:248–58. ­disease. Annu Rev Nutr, 2012;32:343–68.
39. Yang, A.M., Inamine, T., Hochrath, K. et  al. Intestinal fungi contribute to 65. Lieber, S.R., Rice, J.P., Lucey, M.R. et al. Controversies in clinical trials for
development of alcoholic liver disease. J Clin Invest, 2017;127:2829–41. alcoholic hepatitis. J Hepatol, 2018;68(3):586–92.
40. Llopis, M., Cassard, A.M., Wrzosek, L. et  al. Intestinal microbiota 66. Xu, M.J., Zhou, Z., Parker, R. et al. Targeting inflammation for the treatment
contributes to individual susceptibility to alcoholic liver disease. Gut,
­ of alcoholic liver disease. Pharmacol Ther, 2017;180:77–89.
2016;65:830–9. 67. Kong, X., Feng, D., Mathews, S. et  al. Hepatoprotective and anti‐fibrotic
41. Esfandiari, F., Medici, V., Wong, D.H. et al. Epigenetic regulation of hepatic functions of interleukin‐22:therapeutic potential for the treatment of
endoplasmic reticulum stress pathways in the ethanol‐fed cystathionine beta ­alcoholic liver disease. J Gastroenterol Hepatol, 2013;28(1):56–60.
synthase‐deficient mouse. Hepatology, 2010;51:932–41. 68. Gao, B. and Shah, V.H. Combination therapy: new hope for alcoholic hepa-
42. Satishchandran, A., Ambade, A., Rao, S. et al. MicroRNA 122, Regulated by titis? Clin Res Hepatol Gastroenterol, 2015;39(1):S7–11.
GRLH2, protects livers of mice and patients from ethanol‐induced liver 69. Kubiliun, M., Patel, S.J., Hur, C. et al. Early liver transplantation for alco-
­disease. Gastroenterology, 2018;154:238–52. holic hepatitis: ready for primetime? J Hepatol, 2018;68(3):380–2.
53:  Alcoholic Liver Disease 699

70. Mitchell, M.C., Friedman, L.S., and McClain, C.J. Medical management 95. Carlezon, W.A., Jr., Nestler, E.J., and Neve, R.L. Herpes simplex virus‐
of  severe alcoholic hepatitis: expert review from the clinical practice mediated gene transfer as a tool for neuropsychiatric research. Crit Rev
updates  committee of the AGA Institute. Clin Gastroenterol Hepatol, Neurobiol, 2000;14:47–67.
2017;15:5–12. 96. Koob, G.F. Neurobiology of addiction, in Textbook of Substance Abuse
71. Koob, G.F. and Le Moal, M. Drug abuse: hedonic homeostatic dysregula- Treatment, 5th edn, (eds. M. Galanter et  al.), American Psychiatric
tion. Science, 1997;278:52–8. Publishing, Washington, DC, 2015, pp. 3–24.
72. Koob, G.F. Allostatic view of motivation: implications for psychopathology, 97. Economidou, D., Hansson, A.C., Weiss, F. et al. Dysregulation of nocicep-
in Motivational Factors in the Etiology of Drug Abuse, (eds. R.A. Bevins and tin/orphanin FQ activity in the amygdala is linked to excessive alcohol
M.T. Bardo), (series title: Nebraska Symposium on Motivation, Vol 50). drinking in the rat. Biol Psychiatry, 2008;64:211–8.
University of Nebraska Press, Lincoln NE, 2004, pp. 1–18. 98. Serrano, A. and Parsons, L.H. Endocannabinoid influence in drug
73. Koob, G.F. Neurocircuitry of alcohol addiction: synthesis from animal ­reinforcement, dependence and addiction‐related behaviors. Pharmacol
­models, in Alcohol and the Nervous System (eds. E.V. Sullivan and Ther, 2011;132:215–41.
A.  Pfefferbaum), (series title: Handbook of Clinical Neurology, vol 125). 99. Koob, G.F. Alcoholism: allostasis and beyond. Alcohol Clin Exp Res,
Elsevier, Amsterdam, 2014, pp. 33–54. 2003;27:232–43.
74. Weiss, F., Lorang, M.T., Bloom, F.E. et al. Oral alcohol self‐administration 100. Sullivan, E.V. and Pfefferbaum, A. Neurocircuitry in alcoholism: a sub-
stimulates dopamine release in the rat nucleus accumbens: genetic and moti- strate of disruption and repair. Psychopharmacology, 2005;180:583–94.
vational determinants. J Pharmacol Exp Ther, 1993;267:250–8. 101. Oscar‐Berman, M. Function and dysfunction of prefrontal brain circuitry in
75. Robinson, T.E. and Berridge, K.C. The neural basis of drug craving: an alcoholic Korsakoff’s syndrome. Neuropsychol Rev, 2012;22:154–69.
incentive‐sensitization theory of addiction. Brain Res Brain Res Rev, 102. Niendam, T.A., Laird, A.R., Ray, K.L. et al. Meta‐analytic evidence for a
1993;18:247–91. superordinate cognitive control network subserving diverse executive func-
76. Haber, S.N., Fudge, J.L., and McFarland, N.R. Striatonigrostriatal pathways tions. Cogn Affect Behav Neurosci, 2012;12:241–68.
in primates form an ascending spiral from the shell to the dorsolateral stria- 103. Olbrich, H.M., Valerius, G., Paris, C. et al. Brain activation during craving
tum. J Neurosci, 2000;20:2369–82. for alcohol measured by positron emission tomography. Aust N Z J
77. Lovinger, D.M. and Kash, T.L. Mechanisms of neuroplasticity and ethanol’s Psychiatry, 2006;40:171–8.
effects on plasticity in the striatum and bed nucleus of the stria terminalis. 104. Jasinska, A.J., Stein, E.A., Kaiser, J. et al. Factors modulating neural reac-
Alcohol Res, 2015;37:109–24. tivity to drug cues in addiction: a survey of human neuroimaging studies.
78. Solomon, R.L. and Corbit, J.D. An opponent‐process theory of motivation. Neurosci Biobehav Rev, 2014;38:1–16.
I. Temporal dynamics of affect. Psychol Rev, 1974;81:119–45. 105. Koob, G.F. and Volkow, N.D. Neurobiology of addiction: a neurocircuitry
79. Koob, G.F. and Le Moal, M. Drug addiction, dysregulation of reward, and analysis. Lancet Psychiatry, 2016;3:760–73.
allostasis. Neuropsychopharmacology, 2001;24:97–129. 106. Marlatt, G.J. Determinants of relapse: implications for the maintenance of
80. Koob, G.F. and Le Moal, M. Addiction and the brain antireward system. behavioral change, in Behavioral Medicine: Changing Health Lifestyles,
Annu Rev Psychol, 2008;59:29–53. (eds. P. Davidson and S. Davidson), Brynner/Mazel, New York, 1980,
81. Melis, M., Spiga, S., and Diana, M. The dopamine hypothesis of drug addic- pp. 410–52.
tion: hypodopaminergic state. Int Rev Neurobiol, 2005;63:101–54. 107. Jochum, T., Boettger, M.K., Burkhardt, C. et al. Increased pain sensitivity
82. Diana, M., Pistis, M., Carboni, S. et al. Profound decrement of mesolimbic in alcohol withdrawal syndrome. Eur J Pain, 2010;14:713–8.
dopaminergic neuronal activity during ethanol withdrawal syndrome in rats: 108. Zywiak, W.H., Connors, G.J., Maisto, S.A. et al. Relapse research and the
electrophysiological and biochemical evidence. Proc Natl Acad Sci USA, reasons for drinking questionnaire: a factor analysis of Marlatt’s relapse
1993;90:7966–9. taxonomy. Addiction, 1996;91: S121–30.
83. Volkow, N.D., Fowler, J.S., Wang, G.J. et  al. Imaging dopamine’s role in 109. Martin‐Fardon, R. and Weiss, F. Modeling relapse in animals. Curr Top
drug abuse and addiction. Neuropharmacology, 2009;56(1):3–8. Behav Neurosci, 2013;13:403–32.
84. Kyzar, E.J. and Pandey, S.C. Molecular mechanisms of synaptic remodeling 110. McFarland, K. and Kalivas, P.W. The circuitry mediating cocaine‐induced
in alcoholism. Neurosci Lett, 2015;601:11–9. reinstatement of drug‐seeking behavior. J Neurosci, 2001;21:8655–63.
85. Heimer, L. and Alheid, G.F. Piecing together the puzzle of basal forebrain 111. Everitt, B.J. and Wolf, M.E. Psychomotor stimulant addiction: a neural sys-
anatomy. Adv Exp Med Biol, 1991;295:1–42. tems perspective. J Neurosci, 2002;22:3312–20.
86. Alheid, G.F. and Heimer, L. New perspectives in basal forebrain organiza- 112. Monti, P.M., Rohsenow, D.J., Hutchison, K.E. et al. Naltrexone’s effect on
tion of special relevance for neuropsychiatric disorders: the striatopallidal, cue‐elicited craving among alcoholics in treatment. Alcohol Clin Exp Res,
amygdaloid, and corticopetal components of substantia innominata. 1999;23:1386–94.
Neuroscience, 1988;27:1–39. 113. Shaham, Y., Shalev, U., Lu, L. et  al. The reinstatement model of drug
87. LeDoux, J.E. Emotion circuits in the brain. Annu Rev Neurosci, relapse: history, methodology and major findings. Psychopharmacology,
2000;23:155–84. 2003;168:3–20.
88. Neugebauer, V., Li, W., Bird, G.C. et al. The amygdala and persistent pain. 114. Everitt, B.J., Belin, D., Economidou, D. et al. Review. Neural mechanisms
Neuroscientist, 2004;10:221–34. underlying the vulnerability to develop compulsive drug‐seeking habits and
89. Zorrilla, E.P., Logrip, M.L., and Koob, G.F. Corticotropin releasing factor: a addiction. Philos Trans R Soc Lond B Biol Sci, 2008;363:3125–35.
key role in the neurobiology of addiction. Front Neuroendocrinol, 115. Jentsch, J.D. and Taylor, J.R. Impulsivity resulting from frontostriatal dys-
2014;35:234–44. function in drug abuse: implications for the control of behavior by reward‐
90. Roberto, M., Cruz, M.T., Gilpin, N.W. et al. Corticotropin releasing factor‐ related stimuli. Psychopharmacology, 1999;146:373–90.
induced amygdala gamma‐aminobutyric acid release plays a key role in 116. Leggio, L. and Lee, M.R. Treatment of alcohol use disorder in patients with
­alcohol dependence. Biol Psychiatry, 2010;67:831–9. alcoholic liver disease. Am J Med, 2017;130:124–34.
91. Vendruscolo, L.F. and Koob, G.F. Alcohol dependence conceptualized as a 117. Bertholet, N., Daeppen, J.B., Wietlisbach, V. et  al. Reduction of alcohol
stress disorder, in: Oxford Handbook of Stress and Mental Health, (eds. consumption by brief alcohol intervention in primary care: systematic
K.  Harkness and E.P. Hayden), Oxford University Press, New York, 2018; review and meta‐analysis. Arch Intern Med, 2005;165:986–95.
in press. 118. Lieber, C.S., Weiss, D.G., Groszmann, R. et al. II. Veterans affairs coopera-
92. Kirkman, S. and Nelson, D.H. Alcohol‐induced pseudo‐Cushing’s disease: tive study of polyenylphosphatidylcholine in alcoholic liver disease.
a  study of prevalence with review of the literature. Metabolism, Alcohol Clin Exp Res, 2003;27:1765–72.
1988;37:390–4. 119. Khan, A., Tansel, A., White, D.L. et al. Efficacy of psychosocial interven-
93. Richardson, H.N., Lee, S.Y., O’Dell, L.E. et al. Alcohol self‐administration tions in inducing and maintaining alcohol abstinence in patients with
acutely stimulates the hypothalamic‐pituitary‐adrenal axis, but alcohol chronic liver disease: a systematic review. Clin Gastroenterol Hepatol,
dependence leads to a dampened neuroendocrine state. Eur J Neurosci, 2016;14:191–202.
2008;28:1641–53. 120. Leggio, L., Kenna, G.A., and Swift, R.M. New developments for the
94. Vendruscolo, L.F., Barbier, E., Schlosburg, J.E. et al. Corticosteroid‐depend- ­pharmacological treatment of alcohol withdrawal syndrome. A focus on
ent plasticity mediates compulsive alcohol drinking in rats. J  Neurosci, non‐benzodiazepine GABAergic medications. Prog Neuropsychopharmacol
2012;32:7563–71. Biol Psychiatry, 2008;32:1106–17.
700 THE LIVER:  REFERENCES

121. Jonas, D.E., Amick, H.R., Feltner, C. et  al. Pharmacotherapy for adults patients with liver cirrhosis: randomised, double‐blind controlled study.
with alcohol use disorders in outpatient settings: a systematic review and Lancet, 2007;370:1915–22.
meta‐analysis. JAMA, 2014;311:1889–900. 124. Morley, K.C., Baillie, A., Fraser, I. et  al. Baclofen in the treatment of
122. Litten, R.Z., Falk, D.E., Ryan, M.L. et al. Advances in pharmacotherapy ­alcohol dependence with or without liver disease: multisite, randomised,
development: human clinical studies. Handb Exp Pharmacol, 2018; double‐blind, placebo‐controlled trial. Br J Psychiatry, 2018;
248:579–613. 212:362–9.
123. Addolorato, G., Leggio, L., Ferrulli, A. et al. Effectiveness and safety of 125. Koob, G.F. and Volkow, N.D. Neurocircuitry of addiction. Neuropsycho­
baclofen for maintenance of alcohol abstinence in alcohol‐dependent pharmacology, 2010;35:217–38.
Drug‐Induced Liver Injury
54 Lily Dara and Neil Kaplowitz
Research Center for Liver Disease, Department of Medicine, Division of Gastrointestinal and Liver
Diseases, Keck School of Medicine, University of Southern California, Los Angeles, CA, USA

INTRODUCTION value. Drug‐induced hepatitis, when accompanied by jaundice,


results in 10% mortality without transplant. This phenomenon,
Drug‐induced liver injury (DILI) is an increasingly recognized first observed and reported by Hyman Zimmerman, is often
clinical problem encompassing over 50% of cases of acute liver referred to as “Hy’s law”. Drug‐induced cholestasis is a more
failure (ALF) in the United States [1]. First described as a dis- indolent disease, associated with elevated ALP often due to
tinct clinical entity by Hyman Zimmerman in the 1950s, DILI is cholangiocyte damage and may be accompanied by pruritus.
defined as liver injury with alanine aminotransferase (ALT) Cholestatic DILI is a signature of certain drugs, is more com-
more than 3–5 times the upper limit of normal (ULN) or an mon in the elderly and often resolves more slowly [4].
alkaline phosphatase (ALP) more than twice the ULN with or In the past few decades, hundreds of different drugs and
without elevated bilirubin. DILI remains a diagnostic challenge herbal compounds have been reported to cause DILI, with vari-
due to a lack of pathognomonic findings and objective biomark- able latency, patterns of injury, and phenotypic presentation.
ers and a wide range of presentation. DILI has a wide spectrum The underlying mechanism of liver injury and cell death has
of clinical manifestations from acute hepatitis with jaundice to also been extensively explored in vivo and in vitro using differ-
cholestatic DILI to rarer forms of DILI such as nodular regen- ent agents, most notably with the most widely used hepatotoxin,
erative hyperplasia (e.g. azathioprine), sinusoidal obstruction APAP. In this chapter we will focus on the molecular mecha-
syndrome (e.g. cyclophosphamide), steatohepatitis (e.g. tamox- nisms and signaling pathways activated in DILI. We will not
ifen), among others (Table 54.1). DILI can have a rapid or suba- cover herbal and dietary supplement‐induced liver injury and
cute presentation with elevated liver enzymes or occur more although these compounds are not studied as systematically, it is
gradually and in rare instances present as a more chronic pro- likely that the same principles and mechanisms apply.
gressive disease [2]. As such, DILI is a diagnosis of exclusion
and is suspected when there is temporal association between
taking a drug and the onset of liver injury. Despite the various
potential presentations of drug hepatotoxicity, most often DILI DIRECT ACTING VERSUS
presents as drug‐induced hepatitis or cholestatic injury or a IDIOSYNCRATIC TOXICITY
combination of the two. Most drugs have a unique signature
which can be determined by calculating the R‐value. The R‐ The liver is an important target for drugs because lipophilic
value is defined as the ratio of ALT/ULN divided by the ALP/ drugs are metabolized in the liver by conversion to more aque-
ULN. It is commonly used as an index to determine the pheno- ous soluble forms which can be excreted. Drug metabolism
type of liver injury. By convention, an R‐value ≥ 5 suggests a involves the participation of phase I cytochrome P (Cyp)‐450
drug‐induced hepatitis, an R‐value ≤ 2 suggests cholestatic system, phase II conjugation, and phase III transporter proteins,
DILI, and a value in between (2 < R‐value < 5) suggests a mixed which modulate transformation and excretion. Although parent
picture  [3]. The phenotype of liver injury also has prognostic drug accumulation may participate, usually toxic intermediates

The Liver: Biology and Pathobiology, Sixth Edition. Edited by Irwin M. Arias, Harvey J. Alter, James L. Boyer, David E. Cohen, David A. Shafritz,
Snorri S. Thorgeirsson, and Allan W. Wolkoff.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
702 THE LIVER:  ACETAMINOPHEN

Table 54.1  Phenotype and histopathological presentations of DILI


Clinical phenotype Drugs
Hepatocellular injury with non‐specific acute hepatitis on biopsy, zonal Acetaminophen, bromfenac, bupropion, carbon tetrachloride, cocaine,
necrosis, or spotty necrosis (mild to severe injury) diclofenac, halothane, heparin, INH, ipilimumab, ketoconazole,
methotrexate, methyldopa, pembrolizumab, phenytoin, propylthiouracil,
nivolumab, rifampin, statins, troglitazone
Hepatocellular injury with features of autoimmune hepatitis on biopsy Methyldopa, minocycline, nitrofurantoin, statins
Hepatocellular injury with, steatohepatitis +/− fibrosis on biopsy Amiodarone, methotrexate, tamoxifen
Hepatocellular injury with macrovesicular steatosis Anti‐psychotics, MTP (microsomal triglyceride transfer protein) inhibitors,
protease inhibitors
Hepatocellular injury with microvesicular steatosis Aspirin (Reye’s syndrome), NRTI, tetracycline, valproic acid
Cholestasis and inflammation Allopurinol, amoxicillin/clavulanic acid, carbamazepine, chlorpromazine,
hydralazine, methyldopa, penicillamine, phenylbutazone, phenytoin,
procainamide, quinidine, sulfasalazine, sulfonamides, sulindac
Bland cholestasis OCP, cyclosporin A, anabolic steroids
Chronic cholestasis with ductopenia Chlorpromazine, chlorpropamide, chlorothiazide, rarely others
Vascular lesions Anabolic steroids, azathioprine, OCPs (Budd–Chiari syndrome), HIV
Portal hypertension and peliosis hepatitis, NRH, or SOS drugs, busulfan, cyclophosphamide, pyrrolizidine alkaloids,
6‐mercaptopurine
Vascular lesions Busulfan, cyclophosphamide, pyrrolizidine alkaloids
Portal hypertension and sinusoidal obstruction syndrome on biopsy
Adenoma Anabolic steroids, oral contraceptives
Angiosarcoma Anabolic steroids, arsenic, copper, polyvinyl chloride, thorotrast
Hepatocellular carcinoma Aflatoxin, anabolic steroids, danazol

participate in mediating toxicity. Certain drugs such as APAP, ACETAMINOPHEN


aspirin, niacin, and chemotherapeutic agents can cause pre-
dictable, dose‐dependent injury directly to the liver. This direct APAP is the single most common cause of ALF in the United
and metabolic toxicity is unique to certain drugs as most drugs States. [9]. Hepatotoxicity from APAP occurs in a dose‐dependent,
cause an idiosyncratic, unpredictable, and dose‐independent predictable manner, much like the human injury, in rodents
form of hepatotoxicity often abbreviated IDILI (idiosyncratic treated with large doses of the drug. Rats are more resistant to
DILI). Among the direct hepatotoxins, APAP is the mostly APAP; however mice are susceptible and exhibit characteristic
widely studied drug. First introduced by von Mering in 1893, histologic findings, congestion around the central vein, fol-
APAP was not widely used until the 1960s [5]. The first hepa- lowed by vacuolization, and necrosis by 6 hours [5]. Mouse
totoxicity cases were reported in the same decade by Davidson models of APAP toxicity especially C57BL6 mice are com-
and Eastham who described two patients with fulminant liver monly used to study APAP DILI and hepatocyte death. However,
failure and centrilobular hepatic necrosis who subsequently marked mouse strain differences in susceptibility to APAP have
succumbed to APAP toxicity within three days [6]. The toxic- been reported [10]. APAP is the precursor to the toxic metabo-
ity from APAP was shown to be dose dependent and the drug lite, NAPQI (N‐acetyl‐p‐benzoquinone imine), formed by the
was initially considered safe up to 4 g per 24 hours. More direct two electron oxidation of the parent drug by the Cyps
recent studies on healthy volunteers receiving a daily dose of [11]. Multiple Cyps can generate NAPQI but the most common
4 g of APAP have called the concept of a universal safe thresh- enzyme is Cyp2E1. However, Cyp1A2, Cyp3A4, and Cyp2D6
old into question [7]. In addition to dose, fasting, acute illness, may also participate. NAPQI is highly reactive and covalently
concomitant drugs, and metabolic stress can result in liver binds to intracellular proteins resulting in organelle stress. At
injury from lower doses of APAP, due to depletion of glu- nontoxic doses, NAPQI is efficiently detoxified by GSH form-
tathione (GSH) stores or modulation of biotransformation to ing an APAP‐GSH conjugate [12]. GSH synthesis is limited by
toxic intermediary. cysteine availability as the other two amino acids of GSH, glu-
In contrast to direct toxins, drugs that cause IDILI do so by tamate and glycine, are abundant. When GSH is depleted and
activating the immune system. The metabolism of drugs can cysteine supply is limited, the NAPQI is free to attack protein‐
lead to the formation of intermediates which generate hap- SH groups throughout the cell, with the mitochondria being the
tenic‐peptide antigens. A number of studies have now shown key target [13]. This led to the development of the antidote,
that patients who develop IDILI harbor specific single nucle- N‐acetyl‐cysteine (Mucomyst®), which provides cysteine to the
otide polymorphisms (SNPs) in genes which encode for liver and is highly effective if administered within 10 hours of
human leukocyte antigen (HLA) regions. This association overdose in humans and 1.5–2 hours in mice [14]. In the absence
strongly suggests a genetically determined immune predispo- of GSH, NAPQI covalently binds intracellular proteins, and
sition in certain individuals leading to an adaptive immune APAP–protein adducts can be detected in the liver and serum.
response directed at drug hapten or modified autoantigens. The appearance of APAP–protein adducts in serum correlates
However, most of the identified HLA polymorphisms are rel- with hepatotoxicity and serum AST, ALT levels. These adducts
atively common in the general population and thus carry a have been shown by immunoblot analysis to originate from the
small hazard ratio for the development of IDILI. Therefore, liver [15]. This has led to the development of a highly sensitive
other contributing factors must be influencing susceptibility and specific HPLC‐EC assay for detection of acetaminophen
to IDILI [8]. protein adducts (3‐cysteine‐acetaminophen in proteins) in the
54:  Drug‐Induced Liver Injury 703

serum as specific biomarkers of lysis of hepatocytes in APAP further consolidating the importance of mitochondria in this
DILI [16, 17]. form of DILI [34]. Following MPT a combination of collapse of
Despite the excellent correlation between NAPQI‐protein mitochondrial ATP production and release of DNA damaging
adduct formation and toxicity, no direct causality between proteins from the mitochondria result in cell death. In addition
adduct formation and hepatocyte necrosis has been demon- to the MAPKs, other kinases and signaling molecules have been
strated. The removal of APAP–protein adducts though selective implicated in hepatocyte death from APAP. Knockout, knock-
autophagy occurs within 24 hours [18]. NAPQI–protein adducts down or inhibition of glycogen synthase kinase 3 beta (GSK3β),
in the presence of depleted GSH stores result in mitochondrial receptor interacting protein kinase‐1 (RIPK1), and dynamin
damage and increase of intracellular peroxides and reactive related protein‐1 (DRP1) have all been shown to abrogate or
oxygen species (ROS) and via an iron‐mediated mechanism prevent hepatocyte death from APAP [35–37]. While receptor
generate hydroxyl radical (Fenton reaction), as well as, perox- interacting protein kinase 1 (RIPK1) is thought to participate in
ynitrite through the interaction of superoxide with mitochon- APAP toxicity upstream of JNK, the role of receptor interacting
drial nitric oxide (NO) [19]. Hinson and colleagues have protein kinase‐3 (RIPK3) is more controversial [38]. Although
demonstrated the appearance of 3‐nitro‐tyrosine (a marker of APAP‐DILI results in necrotic cell death which is clearly regu-
nitrogen stress) in mouse models of APAP in the centrilobular lated through signaling mechanisms, it is not a form of necrop-
zone, mainly in the mitochondria [20]. Furthermore, the devel- tosis, as mixed lineage kinase domain like (MLKL) knockout
opment of the nitrated proteins in hepatocytes correlated with does not protect against APAP toxicity [36].
the necrotic cell death [20]. The importance of mitochondrial
damage in the pathogenesis of APAP‐induced necrosis has been
known since the 1980s when electron microscopic examination
of livers from acetaminophen treated mice indicated mitochon- IDIOSYNCRATIC DILI
drial damage [21]. Functionally, APAP has been shown to alter
mitochondrial respiration and effect the electron transport chain Idiosyncratic DILI (IDILI) remains a major diagnostic chal-
(ETC) both in vitro and in isolated hepatocytes from in vivo lenge. It is a diagnosis of exclusion and common liver disorders
treated animals [22, 23]. Generation of ROS subsequently exac- such as viral hepatitis, autoimmune hepatitis (AIH), biliary tract
erbates mitochondrial stress and may induce endoplasmic retic- disease, ischemic hepatitis, heart failure, and circulatory dys-
ulum (ER) stress leading to the activation of intracellular function as well as sepsis and alcoholic liver disease must be
signaling pathways, most importantly the mitogen activated ruled out. If a high index of suspicion is present for drug toxic-
protein kinase (MAPK) pathway [24]. Interference with various ity, the more common causes of liver disease are ruled out, and
MAPK proteins, including mixed lineage kinase protein 3 the drug is taken within the correct time frame (latency), and the
(MLK3), apoptosis signal inducing kinase (ASK1), mitogen injury phenotypically resembles the drug’s signature (if known),
activated protein kinase 4 (MKK4), and C Jun‐N‐terminal the probability of an IDILI event is high. A liver biopsy can be
kinase (JNK) as well as the JNK binding partner, SH3BP (Sab) helpful if eosinophilic infiltrates, granulomas, or centrilobular
markedly protects against APAP induced hepatocyte death [23, necrosis is noted. Since there is no specific test to determine
25–29]. Xenobiotic stress, nutrient deprivation, or organelle causality and even the liver biopsy is usually nonspecific, the
stress leads to the activation of upstream higher order MAPKs diagnosis of IDILI remains challenging [8].
which through cascading phosphorylation events ultimately The exact pathophysiology of IDILI and why it occurs is
phosphorylate JNK. Under normal conditions, JNK activation likely multifactorial involving both drug and pharmacologic
by default is transient, as sustained p‐JNK leads to cell death. factors such as dose and lipophilicity as well as host and indi-
When stress signals exceed a certain threshold (such as toxic vidual factors such as genetics, the immune response, and
doses of APAP), p‐JNK interacts with mitochondria by binding defective adaptive responses. Evidence for the involvement of
to the kinase interacting motif of Sab, a mitochondrial outer the immune system in IDILI is strong. For example, certain
membrane protein [30]. The interaction of JNK with Sab leads drugs such as halothane, dihydralazine, and anticonvulsants are
to the generation of ROS and sustained JNK activation, result- associated with an allergic type hypersensitivity and present
ing in a feedforward mechanism. JNK does not enter the mito- with rash and eosinophilia. Other notable drugs that have been
chondria. However, an intra‐mitochondrial signaling pathway reported to present with rash and eosinophilia include: trimetho-
involving the tyrosine phosphatase SH2 phosphatase 1 (SHP1), prim/sulfamethoxazole, cefazolin, and ciprofloxacin [39].
resulting in dephosphorylation of activated‐Src (proto‐onco- Severe allergic skin reactions such as toxic epidermal necrosis
gene c‐Src; Src short for sarcoma), which occurs on and requires (TEN) and Stevens–Johnson syndrome (SJS) have been seen
the platform, DOK4 (an inner membrane protein) has been with carbamazepine and phenytoin IDILI and carry a poor prog-
demonstrated [31]. The deactivation of Src is thought to increase nosis [39]. Some drugs, such as nitrofurantoin and minocycline
mitochondrial ROS generation by dampening electron transport can cause DILI which mimics autoimmune hepatitis with a
leading to the buildup of reducing equivalents. The JNK‐ positive anti‐nuclear antibody (ANA) and plasma cell infiltrates
dependent increased mitochondrial ROS then amplifies mito- which are often indistinguishable from AIH on liver biopsy
chondrial stress and leads to ROS sensitive mitochondrial [39]. This has led to the hypothesis that these drugs may be acti-
permeability transition (MPT) (Figure  54.1). MPT inhibitors vating AIH in patients with a genetic predisposition [40].
such as cyclosporine A inhibit APAP toxicity in vitro and in vivo However, the frequencies of AIH‐associated HLA alleles,
in mice [32, 33]. Cyclophilin D deficient mice are also been DRB1*0301 and DRB1*0401, were not increased in patients
reported to be protected from APAP toxicity and cell death, with AIH type DILI presentation compared to controls [41].
704 THE LIVER:  IDIOSYNCRATIC DILI

Figure 54.1  Signaling pathway in acetaminophen (APAP) induced liver injury. APAP is metabolized by cytochrome P450 2E1 (CYP2E1) to a
reactive metabolite called NAPQI N‐acetyl‐p‐benzoquinone imine which depletes glutathione (GSH) followed by covalent binding to intracellular
proteins which induces cellular and organelle stress. NAPQI targets mitochondria resulting in the generation of reactive oxygen species (ROS).
ROS subsequently activate the mitogen activated protein kinase (MAPK) cascade. Mixed‐lineage protein kinase 3 (MLK3) is activated in the early
phase and apoptosis signal‐inducing kinase‐1 (ASK1) in the later phase of injury. MLK3 and ASK1 phosphorylate mitogen activated protein kinase
4 (MKK4) which goes on to phospho‐activate c‐Jun N‐terminal kinase (JNK). Other kinases such as receptor interacting protein kinase 1 (RIPK1),
glycogen synthase kinase‐3β (GSK3β) and protein kinase C‐α (PKCα) have also been reported to activate JNK. Activated JNK (p‐JNK) binds to
SH3BP (Sab) on the outer mitochondrial membrane. This results in the activation of an intra‐mitochondrial pathway involving a SH2 phosphatase
(SHP1) and the docking protein DOK4 (located on the inner membrane of mitochondria) which ultimately results in the dephosphorylation of
intra‐mitochondrial Src. The deactivation of Src impairs mitochondrial electron transport, increasing mitochondrial ROS generation. The JNK‐
dependent increased mitochondrial ROS then amplifies mitochondrial stress and sustains p‐JNK in an active form. This eventually leads to mito-
chondrial permeability transition (MPT) and release of inter‐mitochondrial proteins which cleave nuclear DNA.

Large databases of IDILI patients have looked for evidence Further supporting the contribution of immunity to DILI are
of a genetic predisposition to IDILI and interestingly many lymphocyte stimulation tests demonstrating drugs or drug–­
large genome wide association studies (GWAS) have identified protein adducts can activate peripheral blood lymphocytes from
polymorphism in HLA regions in patients with IDILI due to affected individuals [42, 43]. Early studies using the lympho-
various agents. Since the HLA complex codes for the major his- cyte transformation test, a simple in vitro assay based on assess-
tocompatibility complex (MHC) proteins in humans, variants in ment of lymphocyte proliferative responses in drug‐treated and
HLA molecules can result in aberrant antigen presentation and vehicle control cultures, detected drug‐specific lymphocyte
inappropriate activation of an immune response. The HLA‐A, responses in approximately 50% of patients with DILI [44].
HLA‐B, and HLA‐C genes govern the structure of MHC I mol- Drug‐specific T cells from IDILI patients have been shown to
ecules, while HLA‐DP, HLA‐DQ, and HLA‐DR control the be activated in an HLA restricted manner, also suggesting an
structure of MHC II. Most identified SNPs are in the region of adaptive immune pathogenesis [42]. Using healthy donors with
MHC II molecules. This can affect antigen presentation to known HLA risk alleles for certain drugs (HLA‐B*15:02,
T helper cells (CD4+) and thus partly explain why certain indi- HLA‐B*57:01, and HLA‐B*58:01) and donors without the risk
viduals develop IDILI. mutations, priming of naïve T cells was demonstrated to skew
54:  Drug‐Induced Liver Injury 705

toward donors expressing the previously identified specific read‐outs have any bearing to the clinical scenario in idiosyn-
HLA alleles in GWAS studies [45]. T cells w

You might also like