Download as pdf or txt
Download as pdf or txt
You are on page 1of 494

SERIES EDITORS

STEPHEN G. WAXMAN
Bridget Marie Flaherty Professor of Neurology
Neurobiology, and Pharmacology;
Director, Center for Neuroscience & Regeneration/Neurorehabilitation Research
Yale University School of Medicine
New Haven, Connecticut
USA

DONALD G. STEIN
Asa G. Candler Professor
Department of Emergency Medicine
Emory University
Atlanta, Georgia
USA

DICK F. SWAAB
Professor of Neurobiology
Medical Faculty, University of Amsterdam;
Leader Research Team Neuropsychiatric Disorders
Netherlands Institute for Neuroscience
Amsterdam
The Netherlands

HOWARD L. FIELDS
Professor of Neurology
Endowed Chair in Pharmacology of Addiction
Director, Wheeler Center for the Neurobiology of Addiction
University of California
San Francisco, California
USA
Elsevier
Radarweg 29, PO Box 211, 1000 AE Amsterdam, The Netherlands
The Boulevard, Langford Lane, Kidlington, Oxford, OX5 1GB, UK

First edition 2012

Copyright Ó 2012 Elsevier B.V. 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 without the prior written permission of the publisher

Permissions may be sought directly from Elsevier’s Science & Technology Rights
Department in Oxford, UK: phone (þ44) (0) 1865 843830; fax (þ44) (0) 1865 853333;
email: permissions@elsevier.com. Alternatively you can submit your request online by
visiting the Elsevier web site at http://elsevier.com/locate/permissions, and selecting
Obtaining permission to use Elsevier material

Notice
No responsibility is assumed by the publisher for any injury and/or damage to persons
or property as a matter of products liability, negligence or otherwise, or from any use
or operation of any methods, products, instructions or ideas contained in the material
herein. Because of rapid advances in the medical sciences, in particular, independent
verification of diagnoses and drug dosages should be made

Library of Congress Cataloging-in-Publication Data


A catalog record for this book is available from the Library of Congress

British Library Cataloguing in Publication Data


A catalogue record for this book is available from the British Library

ISBN: 978-0-444-59427-3
ISSN: 0079-6123

For information on all Elsevier publications


visit our website at store.elsevier.com

Printed and bound in Great Britain

12 13 14 15 11 10 9 8 7 6 5 4 3 2 1
List of Contributors

U. Albrecht, Department of Biology, Unit of Biochemistry, University of Fribourg, Fribourg, Switzerland


A.E. Allen, Faculty of Life Sciences, University of Manchester, Manchester, United Kingdom
S.N. Anand, Neurobehavioural Genetics, MRC Harwell, Harwell Science and Innovation Campus,
Oxfordshire, United Kingdom
C. Ancel, Département de Neurobiologie des Rythmes, Institut des Neurosciences Cellulaires et
Integratives, Strasbourg, France
L. Ansel, Département de Neurobiologie des Rythmes, Institut des Neurosciences Cellulaires et
Integratives, Strasbourg, France
J.L. Barclay, Circadian Rhythms Group, Max Planck Institute for Biophysical Chemistry, Göttingen,
Germany
P. Battley, Ecology Group, Institute of Natural Resources, Massey University, Private Bag, Palmerston
North, New Zealand
C. Bertolucci, Department of Biology and Evolution, University of Ferrara, Ferrara, Italy
P.H.L.T. Bisschop, Department of Endocrinology and Metabolism, Academic Medical Center (AMC),
University of Amsterdam, Amsterdam, The Netherlands
T.M. Brown, Faculty of Life Sciences, University of Manchester, Manchester, United Kingdom
I. Bur, Département de Neurobiologie des Rythmes, Institut des Neurosciences Cellulaires et
Integratives, Strasbourg, France
A. Castillo-Ruiz, Department of Neurology, University of Massachusetts Medical School, Worcester,
MA, USA
E. Christ, Dr. Senckenbergische Anatomie II, Fachbereich Medizin, Goethe-Universität Frankfurt,
Frankfurt am Main, Germany
C.S. Colwell, Laboratory for Neurophysiology, Department of Molecular Cell Biology, Leiden Univer-
sity Medical Center, Leiden, The Netherlands, and Laboratory of Circadian and Sleep Medicine,
Department of Psychiatry and Biobehavioral Sciences, David Geffen School of Medicine, University
of California, Los Angeles, CA, USA
F. De Carli, Institute of Bioimaging and Molecular Physiology, Genoa Unit, National Research Council,
Genoa, Italy
L. De Gennaro, Department of Psychology, University of Rome “Sapienza,” Roma, Italy
T. Deboer, Laboratory for Neurophysiology, Department of Molecular Cell Biology, Leiden University
Medical Center, Leiden, The Netherlands
A. Dekinga, Department of Marine Ecology, Royal Netherlands Institute for Sea Research (NIOZ), Den
Burg, Texel, The Netherlands
H. De la Iglesia, Department of Biology and Program of Neurobiology and Behavior, University of
Washington, Seattle, USA
D.F. Dinges, Division of Sleep and Chronobiology, Department of Psychiatry, Perelman School of Med-
icine, University of Pennsylvania, Philadelphia, PA, USA

v
vi

J.K. Edwards, Neurobehavioural Genetics, MRC Harwell, Harwell Science and Innovation Campus,
Oxfordshire, United Kingdom
M. Ferrara, Department of Health Sciences, University of L’Aquila, L’Aquila, Italy
E. Fliers, Department of Endocrinology and Metabolism, Academic Medical Center (AMC), University
of Amsterdam (UvA), Amsterdam, The Netherlands
R.G. Foster, Nuffield Department of Clinical Neuroscience, Nuffield Laboratory of Ophthalmology,
John Radcliffe Hospital, Headley Way, Oxford, United Kingdom
N.S. Foulkes, Karlsruhe Institute of Technology, Institute of Toxicology and Genetics, Hermann-von-
Helmholtz-Platz, Eggenstein-Leopoldshafen, Germany
N. Goel, Division of Sleep and Chronobiology, Department of Psychiatry, Perelman School of Medicine,
University of Pennsylvania, Philadelphia, PA, USA
Y. Gothilf, Department of Neurobiology, George S. Wise Faculty of Life Sciences, Tel Aviv University,
Tel Aviv, Israel
E. Gwinnerw, Max-Planck-Institut für Ornithologie, Andechs, Germany
M.W. Hankins, Nuffield Department of Clinical Neuroscience, Nuffield Laboratory of Ophthalmology,
John Radcliffe Hospital, Headley Way, Oxford, United Kingdom
D. Hazlerigg, Institute of Biological & Environmental Sciences, University of Aberdeen, Aberdeen,
Scotland, United Kingdom
C. Helfrich-Förster, Biocenter, University of Würzburg, Am Hubland, Würzburg, Germany
B. Helm, Max-Planck-Institut für Ornithologie, Andechs; Fachbereich Biologie, Universität Konstanz,
Konstanz, Germany, and Institute of Biodiversity, Animal Health and Comparative Medicine, Univer-
sity of Glasgow, Glasgow, United Kingdom
K. Honma, Department of Chronomedicine, Hokkaido University Graduate School of Medicine,
Sapporo, Japan
S. Honma, Department of Physiology, and Department of Chronomedicine, Hokkaido University Grad-
uate School of Medicine, Sapporo, Japan
T. Houben, Laboratory for Neurophysiology, Department of Molecular Cell Biology, Leiden University
Medical Center, Leiden, The Netherlands
P. Houdek, Institute of Physiology, Academy of Sciences of the Czech Republic, Prague, Czech Republic
W.-P. Hu, Department of Pharmacology, Hubei University of Science and Technology, Xianning, Hubei,
PR China
A.T.L. Hughes, Faculty of Life Sciences, University of Manchester, Manchester, United Kingdom
S. Hughes, Nuffield Department of Clinical Neuroscience, Nuffield Laboratory of Ophthalmology, John
Radcliffe Hospital, Headley Way, Oxford, United Kingdom
R.A. Hut, Chronobiology unit, Center for Behaviour and Neurosciences, University of Groningen,
Groningen, The Netherlands
M.L. Idda, Karlsruhe Institute of Technology, Institute of Toxicology and Genetics, Hermann-von-
Helmholtz-Platz, Eggenstein-Leopoldshafen, Germany
N. Inagaki, Department of Physiology, Hokkaido University Graduate School of Medicine, Sapporo,
Japan
C.F. Jonkers, Department of Nutrition, Academic Medical Center (AMC), University of Amsterdam,
Amsterdam, The Netherlands

w
Deceased
vii

A. Kalsbeek, Department of Endocrinology and Metabolism, Academic Medical Center (AMC), Univer-
sity of Amsterdam, The Netherlands and Hypothalamic Integration Mechanisms, Netherlands Institute
for Neuroscience (NIN), An Institute of the Royal Netherlands Academy of Arts and Sciences
(KNAW), Amsterdam, The Netherlands
T. Kantermann, Centre for Chronobiology, Faculty of Health and Medical Sciences, University of Sur-
rey, Guildford, Surrey, United Kingdom
P. Klosen, Département de Neurobiologie des Rythmes, Institut des Neurosciences Cellulaires et
Integratives, Strasbourg, France
A. Koolhaas, Department of Marine Ecology, Royal Netherlands Institute for Sea Research (NIOZ),
Den Burg, Texel, The Netherlands
H.-W. Korf, Dr. Senckenbergische Anatomie II, Fachbereich Medizin, and Dr. Senckenbergisches
Chronomedizinisches Institut, Goethe-Universität Frankfurt, Frankfurt am Main, Germany
F. Kreier, Hypothalamic Integration Mechanisms, Netherlands Institute for Neuroscience, an Institute of
the Royal Netherlands Academy of Arts and Sciences, Amsterdam, The Netherlands
N. Kronfeld-Schor, Department of Zoology, Tel Aviv University, Tel Aviv, Israel
G.S. Lall, Medway School of Pharmacy, University of Kent, Chatham, United Kingdom
J.-D. Li, The State Key Laboratory of Medical Genetics, Central South University, Changsha, Hunan,
PR China
R.J. Lucas, Faculty of Life Sciences, University of Manchester, Manchester, United Kingdom
J.H. Meijer, Laboratory for Neurophysiology, Department of Molecular Cell Biology, Leiden University
Medical Center, Leiden, The Netherlands
S. Michel, Laboratory for Neurophysiology, Department of Molecular Cell Biology, Leiden University
Medical Center, Leiden, The Netherlands
Y. Miyazaki, Faculty of Clinical Education, Ashiya University, Hyogo, Japan
C. Moreno, University of São Paulo, School of Public Health, São Paulo, Brazil
F. Moroni, Department of Psychology, University of Rome “Sapienza,” Roma, and Department of Psy-
chology, University of Bologna, Bologna, Italy
C.J. Morris, Division of Sleep Medicine, Brigham and Women’s Hospital, and Harvard Medical School,
Boston, MA, USA
W. Nakamura, Department of Physiology, Hokkaido University Graduate School of Medicine, Sapporo, and
Department of Chronodentistry, Osaka University Graduate School of Dentistry, Suita, Osaka, Japan
T. Nisimura, College of Bioresource Sciences, Nihon University, Fujisawa, Japan
L. Nobili, Centre of Epilepsy Surgery “C. Munari,” Center of Sleep Medicine, Niguarda Hospital, Milan,
and Institute of Bioimaging and Molecular Physiology, Genoa Unit, National Research Council,
Genoa, Italy
P.M. Nolan, Neurobehavioural Genetics, MRC Harwell, Harwell Science and Innovation Campus,
Oxfordshire, United Kingdom
M. Novakova, Institute of Physiology, Academy of Sciences of the Czech Republic, Prague, Czech
Republic
H. Numata, Graduate School of Science, Kyoto University, Kyoto, Japan
D. Ono, Department of Physiology, and Advanced Photonic Bioimaging Section, Research Center for
Cooperative Projects, Hokkaido University Graduate School of Medicine, Sapporo, Japan
H. Oster, Circadian Rhythms Group, Max Planck Institute for Biophysical Chemistry, Göttingen, and
Medical Department I, University of Lübeck, Lübeck, Germany
M.J. Paul, Department of Psychology, University of Massachusetts Amherst, Amherst, MA, USA
viii

S.N. Peirson, Nuffield Department of Clinical Neuroscience, Nuffield Laboratory of Ophthalmology,


John Radcliffe Hospital, Headley Way, Oxford, United Kingdom
T. Piersma, Department of Marine Ecology, Royal Netherlands Institute for Sea Research (NIOZ), Den
Burg, Texel, The Netherlands
H.D. Piggins, Faculty of Life Sciences, University of Manchester, Manchester, United Kingdom
A. Pigorini, Department of Clinical Sciences L. Sacco Università degli Studi di Milano, Milan, Italy
L. Polidarova, Institute of Physiology, Academy of Sciences of the Czech Republic, Prague, Czech
Republic
P. Proserpio, Centre of Epilepsy Surgery “C. Munari,” Center of Sleep Medicine, Niguarda Hospital,
Milan, Italy
D. Rieger, Biocenter, University of Würzburg, Am Hubland, Würzburg, Germany
J. Ripperger, Department of Biology, Unit of Biochemistry, University of Fribourg, Fribourg,
Switzerland
J.H.T. Rohling, Laboratory for Neurophysiology, Department of Molecular Cell Biology, Leiden Univer-
sity Medical Center, Leiden, The Netherlands
F.J. Sánchez-Vázquez, Department of Physiology, Faculty of Biology, University of Murcia, Murcia,
Spain
S. Sarasso, Department of Clinical Sciences L. Sacco Università degli Studi di Milano, Milan, Italy
F.A.J.L. Scheer, Division of Sleep Medicine, Brigham and Women’s Hospital, and Harvard Medical
School, Boston, MA, USA
I. Schwabl, Max-Planck-Institut für Ornithologie, Andechs, Germany
W.J. Schwartz, Department of Neurology, University of Massachusetts Medical School, Worcester, MA,
USA
V. Simonneaux, Département de Neurobiologie des Rythmes, Institut des Neurosciences Cellulaires et
Integratives, Strasbourg, France
D.J. Skene, Centre for Chronobiology, Faculty of Health and Medical Sciences, University of Surrey,
Guildford, Surrey, United Kingdom
M. Sladek, Institute of Physiology, Academy of Sciences of the Czech Republic, Prague, Czech Republic
A.M. Spaeth, Department of Psychology, School of Arts and Sciences, University of Pennsylvania,
Philadelphia, PA, USA
D.J. Stenvers, Department of Endocrinology and Metabolism, Academic Medical Center (AMC),
University of Amsterdam, Amsterdam, The Netherlands
A. Sumova, Institute of Physiology, Academy of Sciences of the Czech Republic, Prague, Czech Republic
Y. Suzuki, Advanced Photonic Bioimaging Section, Research Center for Cooperative Projects, Hokkaido
University Graduate School of Medicine, Sapporo, Japan
A.H. Tsang, Circadian Rhythms Group, Max Planck Institute for Biophysical Chemistry, Göttingen,
Germany
M.A. Ulhôa, University of São Paulo, School of Public Health, São Paulo, Brazil
D. Vallone, Karlsruhe Institute of Technology, Institute of Toxicology and Genetics, Hermann-von-
Helmholtz-Platz, Eggenstein-Leopoldshafen, Germany
R. van der Spek, Department of Endocrinology and Metabolism, Academic Medical Center (AMC),
University of Amsterdam (UvA), Amsterdam, The Netherlands
V. van der Vinne, Chronobiology unit, Center for Behaviour and Neurosciences, University of
Groningen, Groningen, The Netherlands
ix

C. von Gall, Dr. Senckenbergische Anatomie II, Fachbereich Medizin, and Dr. Senckenbergisches
Chronomedizinisches Institut, Goethe-Universität Frankfurt, Frankfurt am Main, Germany
S.M.T. Wehrens, Centre for Chronobiology, Faculty of Health and Medical Sciences, University of
Surrey, Guildford, Surrey, United Kingdom
J.N. Yang, Division of Sleep Medicine, Brigham and Women’s Hospital, Boston, MA, USA
R. Yasenkov, Laboratory for Neurophysiology, Department of Molecular Cell Biology, Leiden Univer-
sity Medical Center, Leiden, The Netherlands
T. Yoshii, Graduate School of Natural Science and Technology, Okayama University, Okayama, Japan
T. Yoshikawa, Advanced Photonic Bioimaging Section, Research Center for Cooperative Projects,
Hokkaido University Graduate School of Medicine, Sapporo, Japan
Q.-Y. Zhou, Department of Pharmacology, University of California Irvine, CA, USA
Preface

The XII Congress of the European Biological Rhythms Society (EBRS), held at the University of Oxford
Examination Schools in Oxford, UK, from August 20–26, 2011, dealt with a broad range of chronobio-
logical disciplines, in an attempt to discuss and further the knowledge in this field.
The research aims of many branches of biology and biomedicine, including neuroscience, are to
deduce fundamental mechanisms and to understand how they are compromised under abnormal or path-
ological conditions. In this regard, the field of chronobiology—the study of periodic (cyclic) phenomena
in living organisms and their adaptation to environmentally generated rhythms—represents one of the
great success stories in biology. The best understood of these periodic phenomena are circadian
rhythms—the rhythms that organize 24-h daily activity and are often referred to as “biological rhythms”
or “the body clock.” Circadian researchers have generated a real understanding, in multiple groups
of organisms, of how genes and their protein products interact to generate stable 24-h cycles of cellular
physiology and behavior. We now also understand some of the mechanisms whereby the environment
interacts with these clocks so that internal time and solar time are appropriately coupled and aligned.
In addition, considerable insight has been gained regarding the central role of internal time in the
regulation of broad areas of physiology and behavior, including sleep and seasonal rhythms. In a little
over 40 years, we have moved from “black box” speculation to real mechanistic insight and clinical
application.
In mammals, including humans, the central biological clock resides in the central nervous system in
the bilateral hypothalamic suprachiasmatic nuclei (SCN). The SCN generates an autonomic rhythm
of electrical activity with a period of approximately (but not exactly) 24 h. This rhythm continues to
oscillate even when SCN cells are removed from a living organism and brought into culture. Approxi-
mately 20 years ago, it was discovered that the endogenous SCN rhythm is based on a molecular
clock mechanism. The core clock genes (Clock, BMAL1, Period, and Cryptochrome) are expressed in
a transcriptional–translational feedback loop with a duration of approximately 24 h. Soon after the dis-
covery of the clock genes, it became clear that not only SCN neurons but virtually every mammalian
cell contains such a molecular clock. These so-called peripheral clocks also have an autonomic cycle of
approximately 24 h that requires synchronization with the external light–dark cycle. Since most peripheral
clocks do not receive light information, the SCN communicates its rhythm to the peripheral clocks.
In addition to the direct hormonal and neuronal signals from the SCN, peripheral clocks are synchronized
by several indirect signals and are probably most strongly influenced by daily changes in energy metabolism.
It is becoming increasingly clear that the different sensitivities of the central and peripheral clocks may lie at
the heart of the discrepancy that often arises between our circadian and social clocks and all of the associated
pathologies.
As the many and diverse contributions to this special edition of Progress in Brain Research on the
“Neurobiology of Circadian Timing” prove, the field of chronobiology has come of age. Of our own spe-
cies, we now know that disruption of our internal timing may lead to major disorders and diseases, from

xi
xii

mental health issues to cancer. The medical world benefits greatly from this knowledge, not least in the
form of the delivery of therapeutics at the most opportune time in terms of good clinical outcome.
Unfortunately, however, many key questions still remain unanswered and many more links with the
clinical practice remain to be established.
Andries Kalsbeek
Martha Merrow
Till Roenneberg
Russell G. Foster
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 1

How rod, cone, and melanopsin photoreceptors


come together to enlighten the mammalian
circadian clock

Robert J. Lucas{,*, Gurprit S. Lall{, Annette E. Allen{ and Timothy M. Brown{

{
Faculty of Life Sciences, University of Manchester, Manchester, United Kingdom
{
Medway School of Pharmacy, University of Kent, Chatham, United Kingdom

Abstract: In mammals, a small number of retinal ganglion cells express melanopsin, an opsin
photopigment, allowing them to be directly photoreceptive. A major function of these so-called
intrinsically photosensitive retinal ganglion cells (ipRGCs) is to synchronize (entrain) endogenous
circadian clocks to the external light:dark cycle. Thanks to their intrinsic light response, ipRGCs can
support photoentrainment even when the other retinal photoreceptors (rods and cones) are absent or
inactive. However, in the intact retina the ipRGC light response is a composite of extrinsic (rod/cone)
and intrinsic (melanopsin) influences. As a result all three photoreceptor classes contribute to the retinal
pathways providing light information to the clock. Here, we consider what each photoreceptor type
contributes to the clock light response. We review electrophysiological and behavioral data pertinent to
this question, primarily from laboratory rodents, drawing them together to provide a conceptual model
in which each photoreceptor class plays a distinct role in encoding the light environment. We finally use
this model to highlight some of the important outstanding questions in this field.

Keywords: circadian; photoentrainment; melanopsin; ipRGCs; photoreceptor; rods; cones;


suprachiasmatic nuclei; retinohypothalamic tract.

Organisms throughout the living world use internal representation of time provided by these
endogenous circadian clocks to modulate their biological clocks is only useful insofar as it reflects
behavior and/or physiology according to the vary- external time. As a result, all circadian clocks
ing demands of the celestial day. However, the have sensory input pathways that report diurnal
variations in the physical environment in order
*Corresponding author. to synchronize (entrain) to local time.
Tel.: þ44 161 2755251 No environmental variable provides a more
E-mail: robert.lucas@manchester.ac.uk reliable indication of time of day than ambient

http://dx.doi.org/10.1016/B978-0-444-59427-3.00001-0 1
2

Rods and cones ganglion cells are themselves photoreceptive,


capable of responding directly to light even when
isolated from the rest of the retina (Berson et al.,
M1 class ipRGC 2002). The photoreceptive capacity of these so-
called intrinsically photosensitive retinal ganglion
Light cells (ipRGCs) relies upon an opsin-based photo-
SCN
EYE pigment called melanopsin (Hattar et al., 2002;
Lucas et al., 2003; Melyan et al., 2005; Panda
et al., 2005; Provencio et al., 2000; Qiu et al.,
Fig. 1. A schematic depiction of the mammalian 2005). Light absorption by melanopsin activates
photoentrainment pathway. The master circadian clock in the a G-protein signaling cascade, which causes cellu-
hypothalamic suprachiasmatic nuclei (SCN) is connected to lar depolarization and increased firing rate.
the retina via a specific component of the optic nerve (the Importantly for the purposes of this review, the
RHT). The RHT is dominated by fibers from the M1 class of
ipRGCs (shown in magnified version of the retina), which are
SCN is a primary central target of ipRGCs. In
directly photosensitive, but also receive signals from rod and fact, despite their relative scarcity in the retina,
cone photoreceptors via multiple types of bipolar and ipRGCs dominate the RHT, with this ganglion
amacrine cell. cell class providing the vast majority of fibers
reaching the SCN (Morin et al., 2003; Sollars
et al., 2003). The functional significance of this
illumination, with light levels varying over up to anatomical arrangement is confirmed by the
nine decades between midnight and midday. It is observation that photoentrainment is lost follow-
unsurprising therefore that light is the most ing selective lesion of ipRGCs (Goz et al., 2008;
important cue for circadian entrainment and that Guler et al., 2008; Hatori et al., 2008). It seems
photoreception is intimately associated with circa- then, that ipRGCs represent the main route via
dian clocks throughout the living world. In the which light information reaches the clock.
case of mammals, entrainment to the diurnal light As ipRGCs contain mechanisms both for
cycle (photoentrainment) originates in the retina, detecting light and transmitting that information
with a distinct projection of the optic nerve, to the brain, one could envisage a scenario in
known as the retinohypothalamic tract (RHT), which they act virtually autonomously to entrain
targeting the master circadian clock in the the clock. However, ipRGCs do not sit in the ret-
suprachiasmatic nuclei (SCN; Fig. 1). ina as islands entire of themselves, but receive
The conventional view of retinal function is synaptic input from several elements of the con-
that all visual information originates with rod ventional rod/cone circuitry within the retina
and/or cone photoreceptors. These specialized (Belenky et al., 2003; Dumitrescu et al., 2009;
cells absorb light and translate that event into Hoshi et al., 2009; Ostergaard et al., 2007; Perez-
neurochemical signals, which are passed through Leon et al., 2006; Viney et al., 2007). As a result,
the retinal circuitry (being sculpted along the their physiological light response (a strong and
way) before reaching the retinal ganglion cells. sustained increase in firing) is a product both of
According to this scheme, retinal ganglion cells their intrinsic phototransduction mechanism and
act principally as relays, receiving information extrinsic influences originating in rods and cones
from the outer retina and communicating it to (Dacey et al., 2005; Schmidt et al., 2008; Tu
the brain in the form of action potentials et al., 2005; Viney et al., 2007; Wong et al., 2007).
propagated along their axons (which form the Physiological and anatomical data therefore
optic nerve). The past decade, however, has seen confirm that the afferent limb of the photo-
the acceptance that a small proportion of retinal entrainment pathway extends to all three
3

photoreceptor types (rod, cone, and melanopsin). of light-evoked changes in neuronal firing in the
Importantly, mice carrying near complete lesions rodent SCN have been used to address this latter
of all three types of photoreception lack objective. Such experiments reveal the hypotha-
photoentrainment (Hattar et al., 2003; Panda lamic light response in real time and therefore
et al., 2003), confirming that no other, as yet describe the dynamic nature of light information
unidentified, photoreceptor can influence the reaching the SCN (e.g., Meijer et al., 1998). On the
clock. However, this still leaves the question of other hand, their relevance for entrainment is less
what role (if any) each of the known photo- obvious, as the mechanistic relationship between
receptors plays in photoentrainment. Here, we changes in firing of neurons within the SCN and
aim to summarize experimental data pertinent adjustments in phase or period of the circadian clock
to this question, concentrating mainly on remains unclear.
findings in mice and related laboratory rodent Several strategies have also been employed to
species because they provide the most compre- determine the contribution made by each photore-
hensive picture. We then draw this together ceptor class to these assays of circadian photosen-
around a simple conceptual model. sitivity. The most conceptually straightforward
has been to study the effects of genetic lesions of
candidate receptor types. The earliest experi-
Methods of study ments of this kind employed naturally occurring
mutations causing rod and cone degeneration.
A number of studies have now addressed the ques- By demonstrating that entrainment survived mas-
tion of under what circumstances, and to what sive loss of these conventional photoreceptors
extent, photoentrainment relies upon each photore- (Ebihara and Tsuji, 1980; Foster et al., 1991), those
ceptor class. They have achieved this using a variety efforts provided justification for the search for non-
of quantitative measures of the clock’s light rod, non-cone photoreceptors, and can be regarded
response. In some cases, photoentrainment of loco- as one of the foundations of this field. More
motor activity rhythms (an output of the SCN clock) recently, increasingly sophisticated methods have
has been studied directly, with the proportion of been developed to lesion each of the photorecep-
animals entrained under a particular light regime tive classes either singly or in combination, and
reported, or the nature of that entrainment without associated cell death (Altimus et al., 2010).
quantified, for example, in terms of its phasing with Retinally degenerate and knockout mice pro-
respect to the light cycle (e.g., Mrosovsky, 2003). vide a powerful method for asking whether each
Other studies have quantified the effect of light on photoreceptor class is necessary and sufficient to
this behavioral rhythm by measuring the account for aspects of the clock light response.
adjustments in phase or period that presumably However, they may provide an incomplete picture
underlie entrainment. These have assessed either of what occurs in the intact organism. Leaving
the period lengthening effects of constant light (LL aside the possibility of compensatory/disruptive
paradigm), or light induced adjustments in the reorganization, such reduced preparations do not
phase of rhythms free-running in constant darkness allow for emergent properties of the intact system
(phase shifting paradigm) (e.g., Mrosovsky and as, for example, the behavior of one photoreceptor
Hattar, 2005). Such behavioral analyses have the system is influenced by activity of another. Such
advantage of reporting phenomena with a fairly interactions are in fact well documented in the
obvious relevance for entrainment. However, retina (Barnard et al., 2006; Cameron and Lucas,
because these assays rely upon end point measures 2009; Cao et al., 2006; Coletta and Adams, 1984;
they do not allow a real-time readout of the clock Frumkes et al., 1992; Hankins and Lucas, 2002;
response to light. Electrophysiological recordings Peachey et al., 1992).
4

To define the function of each photoreceptor in Aspects of the clock response that are impaired at
the context of an intact visual system then, it has longer, but not shorter, wavelengths in TRß2/
been necessary to define additional strategies. Sev- compared to wild types have thus been attributed
eral researchers have achieved this by assessing to cones (see below). Our own approach has been
the spectral sensitivity of response components to use mice (Opn1mwR) in which the coding
(Fig. 2). The efficiency with which photopigments sequence for human red-cone opsin (whose spec-
absorb photons is a function of wavelength. This tral sensitivity function is shifted to much longer
relationship can be described with high precision wavelengths than rod opsin or melanopsin) has
and expressed as a spectral sensitivity profile been “knocked-in” to the mouse M-cone opsin
(Fig. 2a). Comparing such photoreceptor spectral locus (Smallwood et al., 2003). The effect is to pro-
sensitivity functions with assessments of the relative duce an intact and fully functional retina in which
sensitivity of evoked responses to different cones that would ordinarily express M-cone opsin
wavelengths can therefore be a powerful method instead express the longer wavelength human pig-
for linking candidate photoreceptors to specific ment (Fig. 3a). By carefully describing the sensitiv-
clock responses (Fig. 2b). This strategy has been ity of various response components to medium and
employed to identify rod and cone influences on long wavelength stimuli, it has been possible to
light induced firing in the rat SCN (Aggelopoulos identify aspects of the clock response attributable
and Meissl, 2000); to confirm that photoentrainment to each of the photoreceptor classes (Fig. 3b and
in rodless þ coneless mice relies upon melanopsin c; Brown et al., 2011; Lall et al., 2010).
(Hattar et al., 2003; Fig. 2b); and also to demonstrate
that cones alone cannot account for melatonin sup-
pression in humans (Brainard et al., 2001; Thapan Relying on rods
et al., 2001).
A challenge for adopting the spectral sensitivity Measured under the right circumstances, circa-
approach in rodents is that there are only slight dian clocks can be extremely sensitive to light
differences in this parameter between melanopsin, (Evans et al., 2009). As rods are the most sensi-
rod opsin and M-cone opsin (Fig. 2a). As a result, tive of all retinal photoreceptors it would make
very precise estimates of the spectral sensitivity sense if they were to provide this high sensitivity
of evoked responses are required in order to input to the clock. We have tested this predic-
distinguish between the influence of photorecep- tion by assessing the spectral sensitivity of the
tors expressing these three photopigments. Opn1mwR clock in the LL paradigm (Lall et al.,
Descriptions of circadian spectral sensitivity in 2010). We found that even very dim light (near
wild-type mice (and related rodents) thus have threshold for conventional vision) influenced
rarely been able to assign responses unambigu- the clock in this LL paradigm, and that the
ously to a single photoreceptor class (Provencio spectral sensitivity of this effect did indeed match
and Foster, 1995; Takahashi et al., 1984; that of rod, but not cone or melanopsin
Yoshimura and Ebihara, 1996). This limitation photoreceptors.
can, however, be overcome using genetic tec- The clock’s reliance on rod photoreceptors is,
hniques to change the spectral sensitivity of mouse however, not restricted to very dim light. “Rod
photoreceptors. Dkhissi-Benyahya et al. (2007) only” mice, genetically engineered to lack cone
achieved this using thyroid hormone receptor and melanopsin photoreception, entrain well
(TRß2) knockout mice. TRß2 regulates photore- under a light:dark (LD) cycle in which the light
ceptor gene expression and M-cone opsin is not portion is set at 500 photopic lux (roughly equiv-
expressed in TRß2/ mice, leaving all cones to rely alent to bright indoor lighting) (Altimus et al.,
upon S-cone opsin for their photosensitivity. 2010). This confirms that the clock can employ
5

(a) (b)

100

10–1
Relative sensitivity

10–2

10–3

S cone opsin
10–4 Melanopsin
Rod opsin
M cone opsin
10–5
400 450 500 550 600 650 700
Wavelength (nm)

(c) (d)
100

10–1
420 nm
Relative sensitivity
Response amplitude

480 nm
500 nm 10–2
550 nm
580 nm
10–3

10–4 Melanopsin
Circadian Response
10–5
400 450 500 550 600 650
Log (irradiance)
Wavelength (nm)

Fig. 2. Assigning response elements to particular photoreceptors on the basis of their spectral sensitivity. (a) The four mouse retinal
photopigments have more or less well-defined spectral sensitivity profiles, whose forms match the standard template for opsin:
vitamin A based photopigments with peak sensitivities at 360 nm (S-cone opsin), 480 nm (melanopsin), 498 nm (rod opsin), and
508 nm (M-cone opsin). Comparison of these profiles with the spectral sensitivity of specific clock responses can be used to
determine the contribution of each pigment to entrainment. Here, we illustrate this process using mock data for circadian phase
shifts in rodless þ coneless mice (Hattar et al., 2003). (b) The process relies upon measuring phase shifts induced by a discrete
15-min light pulse presented 4 h after activity onset to mice free-running in constant darkness. A typical actogram for a single
mouse exposed to such a treatment, in which bouts of wheel running activity are shown as columns (height indicating number of
wheel revolutions in each 5 min bin) and each line represents 24 h. After entraining the mouse to a light–dark cycle (phasing
shown as black:white boxes at top) for several days, mice are released into constant darkness. Seven to ten days later they
receive a single 15-min light pulse at time indicated by star. By tracing a line through activity onsets before and after the pulse, it
is possible to calculate the associated adjustment in phase. (c) Expressing phase shift magnitude as a function of stimulus
irradiance for several wavelengths of light provides a family of irradiance response curves. The relative sensitivity of the phase
shifting response at each wavelength can then be estimated by the irradiance required to drive a half saturating shift and plotted
as a so-called action spectrum to compare against the absorbance spectrum of candidate photoreceptors. (d) The example here is
for a rodless þ coneless mouse and, as a result, the spectral sensitivity of the phase shifting response closely matches that of
melanopsin, confirming that, in the absence of rods and cones it is melanopsin that supports circadian photosensitivity.
6

(a) Opn1mwR Mouse


(b)
Irradiance log (photons/cm2/s)
100
8 9 10 11 12 13 14
0
10–1

Phase shift (min)


Relative sensitivity
–25
10–2
–50
10–3
–75
500 nm
S cone opsin 650 nm
10–4 Melanopsin –100
Rod opsin
Red cone opsin
10–5
400 450 500 550 600 650 700
Wavelength (nm)

(c) Effective irradiance log (photons/cm2/s) (d)


7 8 9 10 11 12 13 14 650 nm
0

6 spikes/s
460 nm
500 nm
Phase shift (min)

–25 650 nm corrected for:


Cones
–50 Rods
Melanopsin 60 s
–75

–100

Fig. 3. Red-cone knock-in (Opn1mwR) mice. (a) Similarities in the spectral response profile of melanopsin, rod, and M-cone opsins
in wild-type mice (Fig 2a) make it difficult to use the spectral sensitivity of a particular circadian response to establish which of these
pigments it relies upon. We have addressed this problem using a red-cone knock-in mouse (Opn1mwR), in which the introduction of
the human red (L-) cone opsin coding sequence at the mouse M-cone opsin locus results in a long wavelength shift in spectral
sensitivity of the cone population. Using this mouse we have studied the photoreceptive origins of photoentrainment using LL
and phase shifting paradigms (Lall et al., 2010), as well as electrophysiological recordings from the SCN (Brown et al., 2011).
Rather than derive full action spectra for phase shifts in this genotype, spectral sensitivity can be inferred by describing
irradiance response curves at just two wavelengths (b; reproduced with permission from Lall et al., 2010). The phase shifting
response was much more sensitive to 500 than 650 nm stimuli. According to its known spectral sensitivity profile (Fig 2a), red-
cone opsin is 7.7 times more sensitive to 500 than 650 nm. If the phase shifting response were driven by red-cone opsin then we
would be expect the irradiance response curves to be superimposed when stimuli at 500 nm are 7.7 times dimmer than those at
650 nm. To test this prediction, the two curves are expressed in red-cone opsin “effective irradiance” by multiplying measured
irradiance at 650 nm by 0.13. (c) The clear distinction between the two irradiance response curves following this manipulation
indicates that phase shifts are not defined by red-cone opsin under these conditions. Similar corrections based upon the spectral
sensitivity of melanopsin and rod opsin (c), reveal rod opsin as the most likely origin for this response. A simplified version of
this strategy can be employed in which the irradiance of two stimuli of different wavelength (here 460 and 650 nm) are adjusted
so as to appear equally bright to a candidate photoreceptor (in this case red-cone opsin in Opn1mwR mice). A comparison of the
electrophysiological response in the SCN of mice exposed to these two stimuli (d; based upon data presented in Brown et al.,
2011) reveals elements (transient “on” excitation) that are equivalent at the two wavelengths and thus likely originate with red-
cone opsin. Other elements (sustained increase in firing), by contrast, are observed only at the shorter wavelengths and cannot
therefore be attributed to this photopigment. The sustained firing could in theory originate with any photopigment more
sensitive to shorter wavelengths, but for reasons outlined in Brown et al. (2011) probably relies largely on melanopsin.
7

rods to distinguish light from dark under such et al., 1999), while “cone only” mice (lacking both
conditions. Moreover, data from Opn1mwR mice rods and melanopsin) fail to entrain reliably to
indicate that the clock also relies upon rods for standard (at least in the laboratory) LD cycles
more quantitative assessments of irradiance (Lall et al., 2010; Mrosovsky and Hattar, 2005).
under relatively bright conditions, and even when Those data then indicate that cones are neither
cones are fully functional. Thus, the spectral sen- necessary nor sufficient for photoentrainment.
sitivity of phase shifting responses to a 15-min Indeed, our observation that it is rods that deter-
light pulse in this genotype matches that of rods mine the clock response to a 15-min light pulse even
(Fig. 3c) rather than cones (Lall et al., 2010), even at irradiances well within the cone sensitivity range
though the dynamic range for this response lies (Lall et al., 2010) represents further evidence that
 4–5 decimal orders above the threshold for cones have little influence on mechanisms of
rod-based vision and well within the sensitivity entrainment. Nonetheless, anatomical and electro-
range of cones (Nathan et al., 2006). physiological studies indicate that cone signals
The behavioral data thus reveal that rods are a reach the M1 class of ipRGCs responsible for
potent influence on the clock not only under very entrainment (Belenky et al., 2003; Dumitrescu
dim conditions, but also at light levels experienced et al., 2009; Hoshi et al., 2009; Ostergaard et al.,
well into the dawn/dusk transition. Interestingly, 2007; Perez-Leon et al., 2006; Schmidt et al., 2008;
however, this influence is much less apparent in Viney et al., 2007). Moreover, aspects of the
electrophysiological recordings of light responses electrophysiological light response of the SCN itself
in the rodent SCN. Light pulses can evoke large can be attributed to cones on the basis of their spec-
changes in firing (mostly excitatory) in SCN neu- tral sensitivity (Aggelopoulos and Meissl, 2000;
rones, however this requires relatively high irradi- Brown et al., 2011), indicating that cones are indeed
ance. Responses are much less pronounced at the functional components of the light input pathway.
dimmer light intensities that produce large behav- What could explain the apparent contradiction
ioral responses (Brown et al., 2011; Groos and between the ability of cones to influence behavioral
Mason, 1978; Meijer et al., 1992, 1998). This versus electrophysiological endpoints? A closer
indicates that changes in RHT activity driven by look at the electrophysiological light responses
rods may be rather slight but nonetheless are suffi- suggests a possible answer. When presented with a
cient to recruit mechanisms of entrainment. simple light pulse, both ipRGCs (Berson et al.,
2002; Dacey et al., 2005; Schmidt et al., 2008;
Inconstant cones Tu et al., 2005; Wong et al., 2007) and most light
responsive SCN neurons (Brown et al., 2011;
Although rods appear to define circadian responses Meijer et al., 1998) show a transient large increase
even at moderate light intensities, at some point in firing that relaxes over a few seconds to a lower
they are expected to saturate (Wyszecki and Stiles, level of excitation that is sustained over at least tens
1982). Under such conditions, the SCN would no of seconds of light exposure (Fig. 3d). Pharmaco-
longer be able to rely upon them to measure further logical and genetic manipulations suggest that the
increases in light intensity. Cones, by contrast, can acute phase of the ipRGC response originates in
support conventional vision even under the the outer retina implicating rods and/or cones
brightest daylight. Could they then compensate (Berson et al., 2002; Dacey et al., 2005; Schmidt
for rod saturation in the RHT? In fact, the role of et al., 2008; Tu et al., 2005; Wong et al., 2007). We
cones has been (and indeed remains) the hardest studied this aspect of the response in the SCN of
to define. Op1mwR mice (Brown et al., 2011) and observed
Photoentrainment is superficially intact follow- that wavelengths capable of activating cones, but
ing substantial cone degeneration (Freedman not melanopsin, elicited strong increases in firing
8

at lights on, but drove very little sustained activity In fact, there is little direct evidence that any
under continuous light exposure (Fig. 3d). These aspect of mouse photoentrainment relies upon
data indicate that cone-dependent excitation is melanopsin. Melanopsin knockout mice entrain
largely restricted to the first few seconds of light well to full light cycles over a range of irradiances
exposure. (Altimus et al., 2010; Morin and Studholme, 2011;
If cones activate the RHT only briefly at light Panda et al., 2005; Ruby et al., 2002). These
onset, that could explain why this photoreceptor animals do show reduced responsiveness in both
class is ill-suited to supporting entrainment to con- phase shifting and LL paradigms (Altimus et al.,
tinuous long duration ( 15 min) light pulses (Lall 2010; Morin and Studholme, 2011; Panda et al.,
et al., 2010; Mrosovsky and Hattar, 2005). In this 2005; Ruby et al., 2002). However, the simple
case, one might expect cones to make a larger con- explanation for these latter data—that they indi-
tribution to responses driven by shorter duration cate a strong melanopsin contribution—is refuted
pulses (in which the transient cone signal accounts by evidence that in animals with an intact retina
for a greater proportion of the total light-evoked both such assays of circadian light response can
activity of the RHT) and/or a series of light pulses be explained in terms of rod and/or cone activity
(in which the cones signal multiple light onsets). (Lall et al., 2010).
There is evidence to support both those pre- The likely explanation for failures to unam-
dictions. Dkhissi-Benyahya et al. (2007) reported biguously attribute aspects of behavioral pho-
that phase shifts of cone-deficient TRß2/ mice toentrainment to melanopsin is that all such
are equivalent to those of wild types when elicited response parameters saturate at relatively low light
by a 15-min light pulse, but deficient for 1 min levels. In view of the high threshold for melanopsin
stimuli. Similarly, working with Opn1mwR mice phototransduction (Do et al., 2009) one might
we found that cone-activating long wavelength expect that melanopsin’s contribution to entrain-
stimuli that drove no phase shift when given as a ment would be more apparent in paradigms better
single continuous 15 min pulse elicited large able to measure low-sensitivity aspects of the clock
responses when presented as 15 separate 1 min response. This prediction has so far not been tested
pulses presented over 43 min (Lall et al., 2010). in mouse behavioral studies (although rod/cone
The available evidence from laboratory rodents degeneration has been shown to impair entrain-
therefore indicates that while cones are a less reli- ment at low, but not high, irradiances; Morin and
able mediator of entrainment than rods, they can Studholme, 2011; Mrosovsky, 2003). The clock’s
be an important influence on the clock. They fur- reliance upon melanopsin at higher irradiances is,
ther show that cone influence is especially appar- however, very apparent in electrophysiological data
ent in clock responses elicited by one or more from the mouse SCN. SCN light responses build up
sudden increases in irradiance rather than steady, at irradiances within the predicted melanopsin sen-
extended light exposure. sitivity range (and significantly above those
saturating clock responses in most entrainment
and phase shifting studies). Under such conditions,
Melanopsin stimuli targeting cones drive only transient light
responses in the Opn1mwR SCN, while much more
If cones provide such an unreliable influence on sustained excitation is evoked by melanopsin-
the clock, which photoreceptor tells the SCN how activating shorter wavelengths (Fig. 3d; Brown
bright the world is under the daylight conditions et al., 2011). The implication that melanopsin
at which rods are saturated? Melanopsin is the makes a strong contribution to such sustained activ-
least sensitive of the three photoreceptors and ity is supported by data from melanopsin knockout
would appear well set to perform this function. mice (Mure et al., 2007) and from physiological
9

recordings of ipRGCs themselves (Berson et al., in irradiance maximize cone-dependent responses


2002; Dacey et al., 2005; Schmidt et al., 2008; Tu (Dkhissi-Benyahya et al., 2007; Lall et al., 2010),
et al., 2005; Wong et al., 2007). while (at least in rodless þ coneless preparations)
melanopsin can support irradiance dependent
responses to extended light exposure (Freedman
A conceptual model et al., 1999; Mrosovsky, 2003).
Sadly, it has proven difficult to record detect-
The behavioral and physiological data reviewed able SCN electrophysiological responses to dim
above suggest a relatively simple separation in stimuli (below estimates of melanopsin threshold)
the contributions of rods, cones, and melanopsin and thus to determine whether rods and/or cones
to photoentrainment. We describe it here in terms can support transient/sustained responses under
of light information reaching the clock under a such conditions. Aggelopoulos and Meissl (2000)
putative “naturalistic” pattern of light exposure. did record rod-derived responses to dim light in
For our purposes, we assume that superimposed the rat SCN, but used only 1 s pulses, leaving
upon the gradual, diurnal, change in light inten- the question of their temporal profile under lon-
sity are higher frequency modulations in the ger term illumination unanswered. Nonetheless,
amount of light reaching an animal’s eye as, for the evidence that rods can support irradiance
example, clouds obscure the sun or the animal dependent adjustments in circadian phase under
moves in and out of shade (including its burrow). extended light exposure (Altimus et al., 2010; Lall
A highly stylized depiction of such a light expo- et al., 2010), argues that these receptors do allow
sure pattern is provided in Fig. 4a. We can use the SCN to track low frequency changes in irradi-
the data summarized above to consider how the ance. Given the known ability of rods to also
RHT encodes such a profile, and the contribution track higher frequency events, it seems reason-
of each photoreceptor class to that process. able to assume that the SCN could rely on a com-
Our most direct available measure of retinal bination of rods and cones to encode all salient
influences on the circadian clock comes from features of the dynamic light environment at
electrophysiological recordings of the SCN. These dim-moderate irradiances (Fig. 4b).
provide a clear prediction for the separate con- In summary then, according to the conceptual
tributions made by cones and melanopsin to model each photoreceptor class contributes a dif-
encoding the dynamic light exposure profile ferent quality of information about the dynamic
presented in cartoon form in Fig. 4a. Thus, the light environment. Cones encode rapid changes
combination of a cone-derived transient and a in light intensity, leaving rods and melanopsin
melanopsin-dependent sustained SCN response to track more gradual modulations (including
implies that cones would provide information the diurnal variation) at dim and brighter irra-
about higher, and melanopsin lower, frequency diances respectively. Combining these informa-
changes in irradiance. According to this view, the tion sources allows the SCN to monitor light
signal reaching the SCN in an animal exposed to falling on the eye over a wide range of intensities
the irradiance profile in Fig. 4a would comprise a and temporal frequencies.
series of transient cone-derived excitations super-
imposed upon a gradual, melanopsin-dependent, Assumptions, implications, and uncertainties
build up in activity. This prediction, based upon
electrophysiological data (Brown et al., 2011; The cartoon depiction in Fig. 4b represents a best
Mure et al., 2007), is also qualitatively consistent guess of how the light input pathway would encode
with the behavioral data showing that light expo- “naturalistic” patterns of light exposure. It has the
sure paradigms skewed towards rapid increases advantage of relative simplicity and of accounting
10

(a) (b) ~

Input to clock (A.U.)


log (irradiance)
§

Cones

* Rods

Melanopsin

Time
Time
* Around 106 photons/cm2/s
§ Around 1011 photons/cm2/s
~ Around 1017 photons/cm2/s
(c) (d)
High pass filter Low pass filter
Activity (A.U.)

Activity (A.U.)

Time Time

Fig. 4. A conceptual model of rod, cone, and melanopsin contributions to encoding dynamic patterns of light exposure for the
circadian clock. (a) A stylized depiction of the change in corneal irradiance (arbitrary units) over the dawn transition, that
include a gradual (diurnal) increase, as well as more abrupt changes reflecting alterations in the animal’s exposure to solar
radiation as, for example, it moves into/out of shade. (b) The predicted activity (arbitrary units) of the afferent limb of the
photoentrainment pathway to the light exposure profile in (a). The diurnal increase in irradiance elicits a gradual enhancement
of activity dependent on rod (black portion of the curve) and melanopsin (blue) photoreception. More sudden increases are
encoded by rods and/or cones (in red) and appear as spikes in activity of the photoentrainment pathway. Estimates for threshold
and saturation irradiance of rod and melanopsin photoreception are provided (see text for justification). Although information
about the light environment reaching the clock thus has contributions from all three photoreceptor classes, downstream
processing employing either high (c) or low (d) pass temporal filters could exclude melanopsin or cone elements of this signal.

for the available experimental data. However, it is changes in irradiance to occur at only two rates
at best semiquantitative. A number of areas of (“fast” and “slow”), and by assuming that all fast
uncertainty are worth especial mention. changes have sufficient magnitude to evoke a
cone-dependent response. In fact, under dynamic
The role of cones conditions irradiance is likely to vary over a wide
variety of timescales and amplitudes. If we are
Figure 4b masks great uncertainty regarding the to predict the importance of cones for photo-
cone contribution to entrainment, by allowing entrainment under such conditions we therefore
11

need to know much more about the sensory cha- photoreceptor physiology. More excitingly, slight
racteristics of this cone-dependent component of changes in the characteristics of this neural path-
the SCN response. What is its temporal frequency way could fundamentally alter the range of infor-
tuning, that is, what is the relationship between its mation cones provide the clock. In particular, less
activity and the speed/size of changes in irradi- active adaptation in this pathway could allow the
ance? To what extent is the sensitivity of this sig- steady state (irradiance coding) phase of the cone
nal adjusted according to the gradual, diurnal, response to reach the clock. In this situation,
change in irradiance? Does a sudden decrease cones could contribute to the SCN’s ability to
in illumination induce a concomitant transient track lower frequency (diurnal) changes in irradi-
cone-dependent inhibition of SCN firing? ance. This could explain evidence that the clock is
A useful question is whether these properties more reliant on cones in humans than in mice
can simply be inferred from the fundamental sen- (Gooley et al., 2010). Moreover, recent reports
sory characteristics of cone photoreceptors. It is suggest that this could be the case even for differ-
our view that in fact they are likely defined down- ent cone classes in mice. The data reviewed
stream of the photoreceptors. The transience of above (and upon which the model is based)
the cone-derived SCN response to an extended describe responses elicited by cones expressing
light pulse mirrors to some extent that of the cone M-cone opsin (> 95% of the total in mice). Mice
itself. Thus, cones respond to such stimuli with an also have a small population of uniquely short
acute high amplitude hyperpolarization that wavelength sensitive cones (Haverkamp et al.,
relaxes over time. However, under extended light 2005). We have recently reported that these
exposure, cone polarization reaches a “steady S-cones encode more gradual changes in irradi-
state” level that is itself dependent upon irradi- ance for a different ipRGC related response, the
ance (at least at nonbleaching levels; Burkhardt, pupil light reflex (Allen et al., 2011). If that were
1994; Normann and Perlman, 1979; Schnapf also the case for the SCN, it could explain
et al., 1990; Valeton and van Norren, 1983). In older data that the rodent clock is especially sen-
this way, cones can encode both the contrast sitive to UV light (Provencio and Foster, 1995).
and irradiance of a light step in transient and Although a caveat for such a simple explanation
steady state phases of their response. The behav- for that latter data is that Syrian hamsters share
ioral and electrophysiological evidence reviewed the enhanced circadian sensitivity to UV
above showing that the clock relies on cones (Calderone and Jacobs, 1999; von Schantz et al.,
solely to track high frequency changes in irradi- 1997), despite lacking S-cone opsin (Calderone
ance argues that the latter, steady state compo- and Jacobs, 1999; von Schantz et al., 1997).
nent of the cone response is actively excluded Of course it is equally possible that there are
from the photoentrainment pathway. Inclusion circumstances in which the schematic in Fig. 4b
of further adaptation in the neural network bring- greatly overestimates the ability of cones to influ-
ing cone signals to the clock, as described for ence the clock. Indeed, electrophysiological data
other retinal pathways (Dunn et al., 2007), could indicate that cone-derived SCN responses are
achieve this goal. much reduced during the day and under light
We propose therefore that the inability of adapted conditions (Brown et al., 2011).
cones to support sustained aspects of the SCN
response is defined by the neural pathway bring-
ing cone signals to the clock. It follows that the Temporal frequency tuning of melanopsin
temporal frequency tuning characteristics of cone
input to the clock must be determined em- Melanopsin’s ability to track dynamic patterns of
pirically, rather than inferred from known light intensity remains largely unexplored. In the
12

absence of rod/cone input, changes in ipRGC fir- that, under most circumstances, melanopsin light
ing lag the onset and offset of light pulses by up to adaptation does not dramatically skew the clock’s
tens of seconds (Berson et al., 2002). It will be response towards higher frequency changes in illu-
important to determine whether melanopsin minance. According to this view, the most impor-
influences in the intact retina (in which ipRGC tant consequence of adaptation is extension of
firing is also defined by rod/cone signals), and the melanopsin sensitivity range (Wong et al.,
under more naturalistic modulations in irradi- 2005). This feature is discussed below.
ance, show similarly poor temporal fidelity. None-
theless, the long integration time for melanopsin
phototransduction (Do et al., 2009) is consistent Photoreceptor sensitivity ranges
with other evidence (Wong et al., 2007) that it
integrates across most high frequency events. The depiction of photoentrainment in Fig. 4 is
A more pertinent issue for the clock is whether semiquantitative by design. However, the relative
melanopsin responds preferentially to modulations sensitivities of rod, cone, and melanopsin inputs
in irradiance occurring at particular frequencies. remain an important detail. We have based these
According to the conceptual model in Fig. 4, on a combination of known photoreceptor physi-
melanopsin acts as a simple “photon counter” ology and more direct studies of circadian pho-
tracking the slow, diurnal, variation in irradiance, tosensitivity. As outlined below, a particular
irrespective of higher frequency events. In challenge in reviewing this data is that descriptions
fact, sudden increases in irradiance have been of photoreceptor sensitivity in radiometric units
reported to induce anomalously large melanopsin (energy or quanta) will vary according to the spec-
responses, presumably reflecting light adaptation tral quality of the light used in their estimation.
under continuous exposure (Wong et al., 2005). This makes it difficult to extract quantitative infor-
According to the conceptual model in Fig. 4, the mation from studies using “white” light. We have
inclusion of cones in the entrainment pathway based our estimates on data from studies using
would make the clock especially responsive to near monochromatic light at around 480–500 nm,
such an abrupt change in light exposure. Do and all irradiances in the following discussion refer
melanopsin’s light adaptation characteristics aug- to such conditions.
ment this effect? An answer to that question awaits We have estimated the threshold for melanopsin
further study of melanopsin light adaptation and its input to the clock by the minimum corneal irradi-
significance under “naturalistic” light exposure ance required for phase shifts in mice lacking rods
profiles. However, the importance of any such and cones (a little below 1011 photons/cm2/s;
effect is questioned by evidence that the photo- Hattar et al., 2003; Yoshimura and Ebihara,
entrainment pathway has an impressive ability to 1996). In comparing this figure with data from
count photons irrespective of the temporal pattern in vitro recordings from ipRGCs it is important
with which they are presented. Thus, Nelson and to account for the distinction between retinal
Takahashi (1999) showed that the magnitude of and corneal irradiance. Light measures from
hamster phase shifts were defined by the number in vitro studies reflect the former, while it is the
of photons encountered, irrespective of whether latter that is most relevant for behavioral and
they appeared in a single long, or up to 100 other in vivo studies. The relationship between
shorter light pulses. Such temporal integration is retinal and corneal irradiance is defined by the
lost for very short, very bright, pulses (Vidal and ratio of pupil to retinal area. If we assume that
Morin, 2007), and melanopsin light adaptation the mouse retina has an area of  18 mm2, and
could contribute to that process. On balance, how- the pupil area varies from 9 to 0.1 mm2
ever, this behavioral data favors the conclusion (Lyubarsky et al., 2004), this implies a correction
13

factor of between 0.5 (fully dilated pupil) and conservative estimate is that the rod phase of
0.005 (fully constricted pupil) in transforming cor- the clock light response lies between 107 and
neal to retinal irradiance. On this basis, the ability 1011 photons/cm2/s. Given that the equivalent esti-
of Schmidt et al. (2008) to record melanopsin mate for melanopsin is 1011–1017 photons/cm2/s,
responses from ipRGCs at irradiances  1010 pho- this implies that the clock can track irradiance over
tons/cm2/s is broadly consistent with the estimate a most impressive range (up to ten decades) using
from behavioral data (Hattar et al., 2003; just these two photoreceptor types.
Yoshimura and Ebihara, 1996). Although, note Cones adjust their sensitivity to any background
that for most in vitro studies brighter lights have illuminance, and so could encode high frequency
been required to elicit measurable responses modulations in irradiance at any light level above
(Berson et al., 2002; Dacey et al., 2005; Sekaran threshold (>1010 photons/cm2/s; Nathan et al.,
et al., 2003; Tu et al., 2005). 2006). Although the schematic in Fig. 4b does not
There is more uncertainty regarding allow the steady state, irradiance coding, compo-
melanopsin’s saturation point. Phase shifting nent of the cone response to reach the clock, it is
responses in rodless þ coneless mice saturate at worth considering also its range in view of the pos-
corneal irradiance  1013.5 photons/cm2/s (Hattar sibility that it could influence entrainment under
et al., 2003). However, this is unlikely to reflect different circumstances (see above). Data from
saturation in melanopsin phototransduction, in vitro recordings of cone photoreceptors
because other responses in animals with this phe- indicates that this aspect of their response
notype saturate at > 1014 photons/cm2/s even with saturates at irradiances around  1000 above
fully dilated pupils (Brown et al., 2011; Lucas threshold (Burkhardt, 1994; Normann and Per-
et al., 2003; Panda et al., 2003). In vitro studies lman, 1979; Valeton and van Norren, 1983). On
report saturation in aspects of the melanopsin- this basis, we estimate the dynamic range for any
driven light response at anything from 1013.5 putative cone-dependent low frequency signal to
(Schmidt et al., 2008) to > 1015 photons/cm2/s lie between  1010 and 1013 photons/cm2/s. As
(Sekaran et al., 2003). In view of the evidence melanopsin is capable of measuring significantly
for multiple classes of ipRGC with divergent sen- higher irradiances, this implies that melanopsin is
sory capacity (Ecker et al., 2010; Schmidt and unique among mammalian photoreceptors in
Kofuji, 2010), we have based our estimate of the being able to encode low frequency changes in
saturation point for melanopsin signals to the irradiance under brighter (daylight) conditions.
clock upon Berson and colleagues’ original data Before leaving the issue of photoreceptor sensi-
for the ipRGCs innervating the rat SCN (Berson tivity ranges it is worth making a couple of further
et al., 2002). This suggests saturation at a retinal points regarding the wider applicability of these
irradiance  1014–1015 photons/cm2/s. As the pupil estimates. Firstly, they reflect the situation with
will presumably be fully constricted under such respect to corneal irradiance in mice. In the
conditions (Lucas et al., 2001), this translates to absence of evidence to the contrary, it seems rea-
a corneal irradiance closer to 1017 photons/cm2/s. sonable to assume that the fundamental sensitiv-
Rod-dependent clock responses are evident at ity range of the various photoreceptor classes is
very low corneal irradiance (> 107 photons/cm2/s), retained across mammalian species. However,
while rods also define phase shifting responses differences in the ratio of pupil to retinal area
in the range 109–1011 photons/cm2/s (Lall et al., influence the corneal irradiances required to
2010). Rodent rods in vitro saturate around achieve comparable light levels at the retina.
1011 photons/cm2/s (equivalent corneal irradiance In the case of humans, this would likely push all
 1012 photons/cm2/s; Demontis et al., 1993; of the estimates described above for mouse sensi-
Matthews, 1991). Based upon these data a tivity ranges to slightly brighter light. Secondly,
14

while it has been helpful here to discuss sensitivity transient response to lights on/off but showed
ranges in terms of photon flux for monochromatic sustained excitation/inhibition by light. As a
light, outside of the laboratory almost all light is result their activity was strongly influenced by
polychromatic. It is possible to express these melanopsin but not by cones. As cone signals
ranges for each photoreceptor system in terms are a fundamental feature of the ipRGC light
of the photopic units relevant for measuring poly- response, their absence in these hypothalamic
chromatic light (for discussion see Enezi et al., cells implies the presence of a low pass temporal
2011). Thus, dynamic ranges for rods correspond filter somewhere in the network linking these
to 10 4–100 scotopic lux, the “steady state” cone cells to the RHT. Conversely, a minority of light
response to 10 2–101 photopic lux, and melanopsin responsive SCN units show only transient excita-
to 100–106 m-lux; (Enezi et al., 2011). Nonetheless, tion at lights on/off and lack any sustained
the relative sensitivity of the various photoreceptor response during extended light exposure. As a
types will, of course, be dependent upon the spec- result, their activity appears not to be influenced
tral composition of light, especially in species in by melanopsin at all. This response phenotype
which cones have very different spectral sensitivity might be expected if such units were downstream
profiles from rods and melanopsin. of melanopsin-negative non-photoreceptive reti-
nal ganglion cells. However, it would be equally
Moving from input to output possible to produce such a phenotype by includ-
ing a high pass temporal filter on the information
The contribution of each photoreceptor class to provided by the RHT.
light information reaching the clock summarized
in Fig. 4b has been inferred from the recorded Summary and conclusions
impact of light on electrophysiological activity in
the SCN and/or circadian locomotor activity Data from laboratory rodents indicate that each
rhythms. However, it does not follow that all of the three known retinal photoreceptors (rods,
aspects of circadian photosensitivity will be cones, and melanopsin) provides a distinct qual-
influenced by all three photoreceptor types. It is ity of information for the clock. Rods and
clear that, for example, an aspect of the clock melanopsin acting, respectively, at “low” and
response that has high sensitivity and a relatively “high” irradiance allow the clock to respond to
low saturation point (e.g., the phase shift induced gradual changes in irradiance over a very wide
in mice by a 15-min light pulse) would, in effect, range (up to ten decimal orders). Cones, on the
rely entirely on rods (Lall et al., 2010). By con- other hand, provide information about more sud-
trast, it would be surprising if suppression of den increases in light intensity. We have used
pineal melatonin in humans, which has a high these characteristics to provide a qualitative pre-
threshold, were especially reliant upon rods diction of the RHT’s response to dynamic
(Brainard et al., 2001; Thapan et al., 2001). patterns of light exposure.
A less obvious way in which the relative signif- There is an obvious requirement to test the con-
icance of each photoreceptor for a given response ceptual model summarized in Fig. 4 by recording
parameter could be modulated is through applica- the activity of the entrainment pathway under such
tion of a temporal filter (Fig. 4c). The potential a dynamic pattern of light exposure. Moreover, it
effect of such a process is apparent in electrophys- would be useful to replace the highly stylized ver-
iological recordings in the mouse hypothalamus sion of the dawn transition depicted in Fig. 4a with
(Brown et al., 2011). Thus, outside of the SCN more “natural” versions defined by recording
we found large numbers of cells that lacked any light exposure under “field” conditions. Leaving
15

aside these important issues, future work could Brown, T. M., Wynne, J., Piggins, H. D., & Lucas, R. J. (2011).
concentrate on addressing several fundamental Multiple hypothalamic cell populations encoding distinct
visual information. The Journal of Physiology, 589,
aspects of the conceptual model that remain 1173–1194.
poorly understood. In particular, the temporal Burkhardt, D. A. (1994). Light adaptation and photopigment
frequency tuning of cone input to the clock needs bleaching in cone photoreceptors in situ in the retina of
to be defined. We predict that, in fact, this charac- the turtle. The Journal of Neuroscience, 14, 1091–1105.
teristic could be labile; with differences across Calderone, J. B., & Jacobs, G. H. (1999). Cone receptor
variations and their functional consequences in two species
species and/or cone classes in the degree to which of hamster. Visual Neuroscience, 16, 53–63.
these photoreceptors could be used to track more Cameron, M. A., & Lucas, R. J. (2009). Influence of the rod
gradual (including perhaps even diurnal) changes photoresponse on light adaptation and circadian rhythmicity
in irradiance. Resolving these uncertainties will in the cone ERG. Molecular Vision, 15, 2209–2216.
be not merely of academic interest. Understanding Cao, D., Zele, A. J., & Pokorny, J. (2006). Dark-adapted rod
suppression of cone flicker detection: Evaluation of
the quality of information about the light environ- receptoral and postreceptoral interactions. Visual Neurosci-
ment provided by each photoreceptor class is a ence, 23, 531–537.
first step to designing light environments that Coletta, N. J., & Adams, A. J. (1984). Rod-cone interaction in
either maximize or minimize clock responses. As flicker detection. Vision Research, 24, 1333–1340.
such, they represent a most promising strategy Dacey, D. M., Liao, H. W., Peterson, B. B., Robinson, F. R.,
Smith, V. C., Pokorny, J., et al. (2005). Melanopsin-
for controlling circadian phase (and potentially expressing ganglion cells in primate retina signal colour
amplitude) for the benefit of human health. and irradiance and project to the LGN. Nature, 433,
749–754.
Demontis, G. C., Bisti, S., & Cervetto, L. (1993). Light sensi-
References tivity, adaptation and saturation in mammalian rods. Prog-
ress in Brain Research, 95, 15–24.
Aggelopoulos, N., & Meissl, H. (2000). Responses of neurones Dkhissi-Benyahya, O., Gronfier, C., De Vanssay, W.,
of the rat suprachiasmatic nucleus to retinal illumination Flamant, F., & Cooper, H. M. (2007). Modeling the role of
under photopic and scotopic conditions. The Journal of mid-wavelength cones in circadian responses to light. Neu-
Physiology, 523, 211–222. ron, 53, 677–687.
Allen, A. E., Brown, T. M., & Lucas, R. J. (2011). A distinct Do, M. T., Kang, S. H., Xue, T., Zhong, H., Liao, H. W.,
contribution of short-wavelength-sensitive cones to light- Bergles, D. E., et al. (2009). Photon capture and signalling
evoked activity in the mouse pretectal olivary nucleus. The by melanopsin retinal ganglion cells. Nature, 457, 281–287.
Journal of Neuroscience, 31, 16833–16843. Dumitrescu, O. N., Pucci, F. G., Wong, K. Y., & Berson, D. M.
Altimus, C. M., Guler, A. D., Alam, N. M., Arman, A. C., (2009). Ectopic retinal ON bipolar cell synapses in the OFF
Prusky, G. T., Sampath, A. P., et al. (2010). Rod photo- inner plexiform layer: Contacts with dopaminergic amacrine
receptors drive circadian photoentrainment across a wide cells and melanopsin ganglion cells. The Journal of Compar-
range of light intensities. Nature Neuroscience, 13, 1107–1112. ative Neurology, 517, 226–244.
Barnard, A. R., Hattar, S., Hankins, M. W., & Lucas, R. J. Dunn, F. A., Lankheet, M. J., & Rieke, F. (2007). Light adap-
(2006). Melanopsin regulates visual processing in the mouse tation in cone vision involves switching between receptor
retina. Current Biology, 16, 389–395. and post-receptor sites. Nature, 449, 603–606.
Belenky, M. A., Smeraski, C. A., Provencio, I., Sollars, P. J., & Ebihara, S., & Tsuji, K. (1980). Entrainment of the circadian
Pickard, G. E. (2003). Melanopsin retinal ganglion cells activity rhythm to the light cycle: Effective light intensity
receive bipolar and amacrine cell synapses. The Journal of for a Zeitgeber in the retinal degenerate C3H mouse and
Comparative Neurology, 460, 380–393. the normal C57BL mouse. Physiology & Behavior, 24,
Berson, D. M., Dunn, F. A., & Takao, M. (2002). Photo- 523–527.
transduction by retinal ganglion cells that set the circadian Ecker, J. L., Dumitrescu, O. N., Wong, K. Y., Alam, N. M.,
clock. Science, 295, 1070–1073. Chen, S. K., LeGates, T., et al. (2010). Melanopsin-
Brainard, G., Hanifin, J., Greeson, J., Byrne, B., Glickman, G., expressing retinal ganglion-cell photoreceptors: Cellular
Gerner, E., et al. (2001). Action spectrum for melatonin reg- diversity and role in pattern vision. Neuron, 67, 49–60.
ulation in humans: Evidence for a novel circadian photore- Enezi, J., Revell, V., Brown, T., Wynne, J., Schlangen, L., &
ceptor. The Journal of Neuroscience, 21, 6405–6412. Lucas, R. (2011). A "melanopic" spectral efficiency function
16

predicts the sensitivity of melanopsin photoreceptors to Haverkamp, S., Wassle, H., Duebel, J., Kuner, T.,
polychromatic lights. Journal of Biological Rhythms, 26, Augustine, G. J., Feng, G., et al. (2005). The primordial,
314–323. blue-cone color system of the mouse retina. The Journal of
Evans, J. A., Elliott, J. A., & Gorman, M. R. (2009). Neuroscience, 25, 5438–5445.
Dim nighttime illumination accelerates adjustment to time- Hoshi, H., Liu, W. L., Massey, S. C., & Mills, S. L. (2009). ON
zone travel in an animal model. Current Biology, 19, inputs to the OFF layer: Bipolar cells that break the stratifi-
R156–R157. cation rules of the retina. The Journal of Neuroscience, 29,
Foster, R. G., Provencio, I., Hudson, D., Fiske, S., De 8875–8883.
Grip, W., & Menaker, M. (1991). Circadian photoreception Lall, G. S., Revell, V. L., Momiji, H., Al Enezi, J.,
in the retinally degenerate mouse (rd/rd). Journal of Com- Altimus, C. M., Guler, A. D., et al. (2010). Distinct con-
parative Physiology. A, 169, 39–50. tributions of rod, cone, and melanopsin photoreceptors to
Freedman, M. S., Lucas, R. J., Soni, B., von Schantz, M., encoding irradiance. Neuron, 66, 417–428.
Munoz, M., David-Gray, Z., et al. (1999). Regulation of Lucas, R., Douglas, R., & Foster, R. (2001). Characterization
mammalian circadian behavior by non-rod, non-cone, ocular of an ocular photopigment capable of driving pupillary con-
photoreceptors. Science, 284, 502–504. striction in mice. Nature Neuroscience, 4, 621–626.
Frumkes, T. E., Lange, G., Denny, N., & Beczkowska, I. Lucas, R. J., Hattar, S., Takao, M., Berson, D. M.,
(1992). Influence of rod adaptation upon cone responses to Foster, R. G., & Yau, K. W. (2003). Diminished pupillary
light offset in humans: I. Results in normal observers. Visual light reflex at high irradiances in melanopsin-knockout mice.
Neuroscience, 8, 83–89. Science, 299, 245–247.
Gooley, J. J., Rajaratnam, S. M., Brainard, G. C., Lyubarsky, A. L., Daniele, L. L., & Pugh, E. N. Jr. (2004).
Kronauer, R. E., Czeisler, C. A., & Lockley, S. W. (2010). From candelas to photoisomerizations in the mouse eye by
Spectral responses of the human circadian system depend rhodopsin bleaching in situ and the light-rearing depen-
on the irradiance and duration of exposure to light. Science dence of the major components of the mouse ERG. Vision
Translational Medicine, 2, 31ra33. Research, 44, 3235–3251.
Goz, D., Studholme, K., Lappi, D. A., Rollag, M. D., Matthews, H. R. (1991). Incorporation of chelator into guinea-
Provencio, I., & Morin, L. P. (2008). Targeted destruction pig rods shows that calcium mediates mammalian photore-
of photosensitive retinal ganglion cells with a saporin conju- ceptor light adaptation. The Journal of Physiology, 436,
gate alters the effects of light on mouse circadian rhythms. 93–105.
PLoS One, 3, e3153. Meijer, J. H., Rusak, B., & Ganshirt, G. (1992). The relation
Groos, G., & Mason, R. (1978). Maintained discharge of rat between light-induced discharge in the suprachiasmatic
suprachiasmatic neurons at different adaptation levels. Neu- nucleus and phase shifts of hamster circadian rhythms. Brain
roscience Letters, 8, 59–64. Research, 598, 257–263.
Guler, A. D., Ecker, J. L., Lall, G. S., Haq, S., Altimus, C. M., Meijer, J., Watanabe, K., Schaap, J., Albus, H., & Detari, L.
Liao, H. W., et al. (2008). Melanopsin cells are the principal (1998). Light responsiveness of the suprachiasmatic nucleus:
conduits for rod-cone input to non-image-forming vision. Long-term multiunit and single-unit recordings in freely
Nature, 453, 102–105. moving rats. The Journal of Neuroscience, 18, 9078–9087.
Hankins, M. W., & Lucas, R. J. (2002). The primary visual Melyan, Z., Tarttelin, E. E., Bellingham, J., Lucas, R. J., &
pathway in humans is regulated according to long-term light Hankins, M. W. (2005). Addition of human melanopsin
exposure through the action of a nonclassical photopigment. renders mammalian cells photoresponsive. Nature, 433,
Current Biology, 12, 191–198. 741–745.
Hatori, M., Le, H., Vollmers, C., Keding, S. R., Tanaka, N., Morin, L., Blanchard, J., & Provencio, I. (2003). Retinal gan-
Schmedt, C., et al. (2008). Inducible ablation of glion cell projections to the hamster suprachiasmatic
melanopsin-expressing retinal ganglion cells reveals their nucleus, intergeniculate leaflet and visual midbrain: Bifurca-
central role in non-image forming visual responses. PLoS tion and melanopsin immunoreactivity. The Journal of Com-
One, 3, e2451. parative Neurology, 465, 401–416.
Hattar, S., Liao, H. W., Takao, M., Berson, D. M., & Morin, L. P., & Studholme, K. M. (2011). Separation of func-
Yau, K. W. (2002). Melanopsin-containing retinal ganglion tion for classical and ganglion cell photoreceptors with
cells: Architecture, projections, and intrinsic photosensitiv- respect to circadian rhythm entrainment and induction of
ity. Science, 295, 1065–1070. photosomnolence. Neurosci, 199, 213–224.
Hattar, S., Lucas, R. J., Mrosovsky, N., Thompson, S., Mrosovsky, N. (2003). Contribution of classic photoreceptors
Douglas, R. H., Hankins, M. W., et al. (2003). Melanopsin to entrainment. Journal of Comparative Physiology. A, Neu-
and rod-cone photoreceptive systems account for all major roethology, Sensory, Neural, and Behavioral Physiology,
accessory visual functions in mice. Nature, 424, 75–81. 189, 69–73.
17

Mrosovsky, N., & Hattar, S. (2005). Diurnal mice Schmidt, T. M., & Kofuji, P. (2010). Differential cone pathway
(Mus musculus) and other examples of temporal niche influence on intrinsically photosensitive retinal ganglion cell
switching. Journal of Comparative Physiology. A, Neu- subtypes. The Journal of Neuroscience, 30, 16262–16271.
roethology, Sensory, Neural, and Behavioral Physiology, Schmidt, T. M., Taniguchi, K., & Kofuji, P. (2008). Intrinsic
191, 1011–1024. and extrinsic light responses in melanopsin-expressing gan-
Mure, L. S., Rieux, C., Hattar, S., & Cooper, H. M. (2007). glion cells during mouse development. Journal of Neuro-
Melanopsin-dependent nonvisual responses: Evidence for physiology, 100, 371–384.
photopigment bistability in vivo. Journal of Biological Schnapf, J. L., Nunn, B. J., Meister, M., & Baylor, D. A.
Rhythms, 22, 411–424. (1990). Visual transduction in cones of the monkey Macaca
Nathan, J., Reh, R., Ankoudinova, I., Ankoudinova, G., fascicularis. The Journal of Physiology, 427, 681–713.
Chang, B., Heckenlively, J., et al. (2006). Scotopic and Sekaran, S., Foster, R. G., Lucas, R. J., & Hankins, M. W.
photopic visual thresholds and spatial and temporal discrim- (2003). Calcium imaging reveals a network of intrinsically
ination evaluated by behavior of mice in a water maze. light-sensitive inner-retinal neurons. Current Biology, 13,
Photochemistry and Photobiology, 82, 1489–1494. 1290–1298.
Nelson, D. E., & Takahashi, J. S. (1999). Integration and satu- Smallwood, P. M., Olveczky, B. P., Williams, G. L.,
ration within the circadian photic entrainment pathway of Jacobs, G. H., Reese, B. E., Meister, M., et al. (2003).
hamsters. The American Journal of Physiology, 277, Genetically engineered mice with an additional class of cone
R1351–R1361. photoreceptors: Implications for the evolution of color
Normann, R. A., & Perlman, I. (1979). The effects of back- vision. Proceedings of the National Academy of Sciences of
ground illumination on the photoresponses of red and green the United States of America, 100, 11706–11711.
cones. The Journal of Physiology, 286, 491–507. Sollars, P. J., Smeraski, C. A., Kaufman, J. D., Ogilvie, M. D.,
Ostergaard, J., Hannibal, J., & Fahrenkrug, J. (2007). Synaptic Provencio, I., & Pickard, G. E. (2003). Melanopsin and non-
contact between melanopsin-containing retinal ganglion melanopsin expressing retinal ganglion cells innervate the
cells and rod bipolar cells. Investigative Ophthalmology & hypothalamic suprachiasmatic nucleus. Visual Neuroscience,
Visual Science, 48, 3812–3820. 20, 601–610.
Panda, S., Nayak, S. K., Campo, B., Walker, J. R., Takahashi, J., DeCoursey, P., Bauman, L., & Menaker, M.
Hogenesch, J. B., & Jegla, T. (2005). Illumination of the (1984). Spectral sensitivity of a novel photoreceptive system
melanopsin signaling pathway. Science, 307, 600–604. mediating entrainment of mammalian circadian rhythms.
Panda, S., Provencio, I., Tu, D. C., Pires, S. S., Rollag, M. D., Nature, 308, 186–188.
Castrucci, A. M., et al. (2003). Melanopsin is required for Thapan, K., Arendt, J., & Skene, D. (2001). An action spec-
non-image-forming photic responses in blind mice. Science, trum for melatonin suppression: Evidence for a novel non-
301, 525–527. rod, non-cone photoreceptor system in humans. The Journal
Peachey, N. S., Alexander, K. R., Derlacki, D. J., & of Physiology, 535, 261–267.
Fishman, G. A. (1992). Light adaptation, rods, and the Tu, D. C., Zhang, D., Demas, J., Slutsky, E. B., Provencio, I.,
human cone flicker ERG. Visual Neuroscience, 8, 145–150. Holy, T. E., et al. (2005). Physiologic diversity and develop-
Perez-Leon, J. A., Warren, E. J., Allen, C. N., ment of intrinsically photosensitive retinal ganglion cells.
Robinson, D. W., & Lane Brown, R. (2006). Synaptic inputs Neuron, 48, 987–999.
to retinal ganglion cells that set the circadian clock. The Valeton, J. M., & van Norren, D. (1983). Light adaptation of
European Journal of Neuroscience, 24, 1117–1123. primate cones: An analysis based on extracellular data.
Provencio, I., & Foster, R. (1995). Circadian rhythms in mice Vision Research, 23, 1539–1547.
can be regulated by photoreceptors with cone-like cha- Vidal, L., & Morin, L. P. (2007). Absence of normal photic inte-
racteristics. Brain Research, 694, 183–190. gration in the circadian visual system: Response to millisec-
Provencio, I., Rodriguez, I. R., Jiang, G., Hayes, W. P., ond light flashes. The Journal of Neuroscience, 27, 3375–3382.
Moreira, E. F., & Rollag, M. D. (2000). A novel human Viney, T. J., Balint, K., Hillier, D., Siegert, S., Boldogkoi, Z.,
opsin in the inner retina. The Journal of Neuroscience, 20, Enquist, L. W., et al. (2007). Local retinal circuits of
600–605. melanopsin-containing ganglion cells identified by trans-
Qiu, X., Kumbalasiri, T., Carlson, S. M., Wong, K. Y., synaptic viral tracing. Current Biology, 17, 981–988.
Krishna, V., Provencio, I., et al. (2005). Induction of photo- von Schantz, M., Argamaso-Hernan, S., Szel, A., & Foster, R.
sensitivity by heterologous expression of melanopsin. (1997). Photopigments and photoentrainment in the Syrian
Nature, 433, 745–749. golden hamster. Brain Research, 770, 131–138.
Ruby, N. F., Brennan, T. J., Xie, X., Cao, V., Franken, P., Wong, K. Y., Dunn, F. A., & Berson, D. M. (2005). Photore-
Heller, H. C., et al. (2002). Role of melanopsin in circadian ceptor adaptation in intrinsically photosensitive retinal gan-
responses to light. Science, 298, 2211–2213. glion cells. Neuron, 48, 1001–1010.
18

Wong, K. Y., Dunn, F. A., Graham, D. M., & Berson, D. M. Yoshimura, T., & Ebihara, S. (1996). Spectral sensitivity of
(2007). Synaptic influences on rat ganglion-cell photo- photoreceptors mediating phase-shifts of circadian rhythms
receptors. The Journal of Physiology, 582, 279–296. in retinally degenerate CBA/J (rd/rd) and normal CBA/N
Wyszecki, G., & Stiles, W. (1982). Color science: Concepts and (þ/þ)mice. Journal of Comparative Physiology. A, 178,
methods, quantitative data and formulae. New York: Wiley- 797–802.
Interscience.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 2

Melanopsin phototransduction: Slowly


emerging from the dark

Steven Hughes, Mark W. Hankins, Russell G. Foster* and Stuart N. Peirson*

Nuffield Department of Clinical Neuroscience, Nuffield Laboratory of Ophthalmology,


John Radcliffe Hospital, Headley Way, Oxford, United Kingdom

Abstract: Melanopsin expressing retinal ganglion cells represent a third class of ocular photoreceptors
and are involved in irradiance detection and non-image-forming responses to light including pupil
constriction, circadian entrainment, and regulation of sleep. Over recent years, there has been a rapid
increase in our understanding of the anatomical variety of pRGC subtypes, the regions of the brain which
they innervate, and the behavioral responses of melanopsin-based light detection. However, by contrast,
our understanding of the intracellular signaling cascade initiated following activation of melanopsin has,
until recently, remained poorly characterized. This chapter focus on the melanopsin signaling pathway,
detailing the cellular mechanisms of phototransduction that occur within pRGCs, highlighting recent
advances, but also the gaps in our understanding of this important light detecting system.

Keywords: melanopsin; Opn4; phototransduction; pRGC; retina; TRPC.

Introduction the ventrolateral preoptic areas (VLPO) (Baver


et al., 2008; Chen et al., 2011; Gooley et al.,
In addition to the rods and cones, a third class of 2003; Hannibal and Fahrenkrug, 2004; Hattar
photoreceptor has been unequivocally identified et al., 2002, 2006) and are primarily involved in
within the mammalian retina, the melanopsin- non-image forming responses to light, including
expressing photosensitive retinal ganglion cell circadian entrainment (phase shifting), suppression
(pRGC) (for review see Do and Yau, 2010; of melatonin release, pupil constriction, masking
Hankins et al., 2008). pRGCs innervate numerous and induction of sleep, and negative phototaxis
regions of the brain, including the suprachiasmatic (Freedman et al., 1999; Johnson et al., 2010; Lucas
nuclei (SCN), olivary pretectal nucleus (OPN), and and Foster, 1999; Lucas et al., 2001, 2003; Lupi
et al., 2008; Mrosovsky et al., 2001; Panda et al.,
*Corresponding authors. 2002; Ruby et al., 2002; Semo et al., 2010). How-
Tel.: þ44 (01865) 234782; Fax: þ 44 (01865) 234795 ever, more recently, it has become clear that
E-mail: russell.foster@eye.ox.ac.uk; stuart.peirson@eye.ox.ac.uk pRGCs also innerve regions of the brain tasked

http://dx.doi.org/10.1016/B978-0-444-59427-3.00002-2 19
20

with processing visual information, and it would stratification of their dendrites within specific sub-
seem that melanopsin may perform roles in visual- laminae of the inner plexiform layer (IPL) (for
processing and image-forming vision (Brown et al., review see Do and Yau, 2010 and Schmidt et al.,
2010; Ecker et al., 2010). 2011). The first melanopsin cells to be identified
were the M1-type pRGCs (Berson et al., 2002;
pRGC subtypes Hattar et al., 2002, 2006; Provencio et al., 2002).
M1 cells express high levels of melanopsin protein
Since their original description, it has become clear and have sparsely branching dendritic fields that
that pRGCs are not a homogeneous population of stratify exclusively in the OFF layer of the IPL.
cells, but instead consist of numerous functionally The majority of M1-type cells have cell bodies
distinct cell types (Fig. 1). To date, the classification located in the GCL although a subset of M1 cells
of pRGC subtype has extended to include five clas- have their cell bodies located in the INL, termed
ses of pRGC, which are distinguished based on displaced M1 pRGCs. Soon after the characteriza-
levels of melanopsin expression and the tion of M1-type cells, a second class of melanopsin

(a) (c)

OS

ONL

OPL
INL
OFF
IPL
ON M1 25 mm M2 25 mm
GCL

M1 M2 M3
(b)
OFF
ON

M3 50 mm M4, M5 50 mm

M1 M2 M3 M4 M5

Fig. 1. Subtype of melanopsin-expressing pRGC. (a) Schematic showing the layers and cell types of the mammalian retina,
including the melanopsin-expressing photosensitive ganglion cells (pRGCs). (b) Schematic showing the different levels of
melanopsin expressed in pRGC subtypes (denoted by intensity of red coloring) and differing patterns of stratification within the
inner plexiform layer. (c) Images showing the pRGC subtypes indentified in the mouse retina to date. Panels show examples of
M1-, M2-, and M3-type pRGCs in wild-type mouse retina stained with an N-terminal melanopsin antibody (shown in red, UF006
antibody) and also M4- and M5-type cells that are only identified via detection of EYFP expression in Opn4-Cre mice (shown in
green). In the panel showing M4-and M5-type cells, a melanopsin-immunoreactive M1-type cell is included for comparison
(Opn4-Cre mice (Ecker et al., 2010) were kindly provided by Sammar Hattar, John Hopkins University). Outer segments (OS),
outer nuclear layer (ONL), outer plexiform layer (OPL), inner nuclear layer (INL), inner plexiform layer (IPL), ganglion cell
layer (GCL). OFF and ON denote the OFF and ON sublaminae of the IPL, respectively.
21

cell was identified. M2-type pRGCs express lower and at least two distinct types of responses have
levels of melanopsin compared to M1-type cells, been observed from adult pRGCs using multiple
and have processes that stratify in the ON layer electrode arrays (MEAs), based on sensitivity
of the IPL. M2 cells also have larger cell bodies and response latency (Tu et al., 2005). Although
and larger and more complex dendritic fields these studies did not correlate these responses
(Baver et al., 2008; Berson et al., 2010; Ecker with the different anatomical subtypes of pRGC,
et al., 2010; Hattar et al., 2006; Schmidt and subsequent reports have indicated that a range
Kofuji, 2009; Viney et al., 2007). M3-type of functional differences do exist between these
pRGCs are bistratified and have dendrites cells. These include differences in membrane pro-
located in both the OFF and ON layers of the perties and resting membrane potentials, as well
IPL (Berson et al., 2010; Schmidt and Kofuji, as their overall sensitivity to light and kinetics of
2011; Schmidt et al., 2008; Viney et al., 2007). photoresponses (Ecker et al., 2010; Schmidt and
These cells are anatomically diverse and rela- Kofuji, 2009, 2010, 2011).
tively rare, failing to tile the entire retina. It For M1 cells, there are two clear phases of the
has been suggested that these cells may repre- light-induced photoresponse. An initial fast-acting
sent a developmental anomaly and may not rep- transient followed by a smaller more sustained
resent a true class of pRGC (Berson et al., component. The initial large transient component
2010), although they do seem to have distinct to the light response is absent from M2-type
functional properties (Schmidt and Kofuji, pRGCs (and M3-, M4-, and M5-type pRGCs)
2011). Most recently, M4- and M5-type pRGCs that show small but sustained responses (Ecker
have been identified. M4- and M5-type pRGCs et al., 2010; Schmidt and Kofuji, 2009, 2011).
are broadly similar in morphology to M2-type M1 cells increase spike-firing rate with increasing
pRGCs with dendrites located in the ON layers intensities of light (Do et al., 2009), yet M1 cells
of the IPL, yet are distinguished based on the are prone to depolarization block and are only
size and complexity of their dendritic fields able to fire action potentials at relatively low rates
(Ecker et al., 2010). Levels of melanopsin (Schmidt and Kofuji, 2009; Wong et al., 2007).
expression in M4 and M5 cells are seemingly In contrast, M2 cells are 10-fold less sensitive to
very low, and below that which can be detected light than M1 cells, produce a 10-fold smaller
reliably with melanopsin antibodies (Berson maximum photocurrent and can fire action
et al., 2010; Ecker et al., 2010). Most recently, potentials at higher frequencies than M1-type
it has become clear that the M1 subclass of cells (Ecker et al., 2010; Schmidt and Kofuji,
pRGCs is further subdivided into cells that 2009). M3-type cells share many properties of
express the Brn3b transcription factor and cells M2-type cells, including more hyperpolarized
that are Brn3b negative (Chen et al., 2011). membrane potentials and reduced input resis-
Thus, the anatomical diversity of pRGCs tance compared to M1-type cells, yet their overall
subtypes continues to be refined. sensitivity to light and their propensity to achieve
depolarization block is increased compared to M2
Functional differences between pRGC subtypes cells, but is less than that observed for M1 cells
(Schmidt and Kofuji, 2011). In keeping with the
From early studies, it was clear that heterogeneity low levels of melanopsin expression, the intrinsic
exists in the responses of individual pRGCs. At photoresponses recorded from M4 and M5 cells
least three distinct types of responses have been are small, even smaller than those observed in
observed from pRGCs using calcium imaging of M2-type cells, with properties similar to that
retinal explants, classified as transient, sustained, observed for M2-type cells (Ecker et al., 2010).
and repetitive responses (Sekaran et al., 2005) Although it is clear that the biophysical properties
22

of the responses vary between pRGC subtypes, exert larger influence on M2-type cells than M1-
the properties of responses elicited by each sub- type cells. Photoresponses of M1 cells are driven
class have yet to be fully determined. primarily by melanopsin function with little influ-
ence from synaptic inputs (Schmidt and Kofuji,
2009, 2010). By contrast, inputs from cone ON
Retinal connections bipolar cells represent the major driving force
behind the generation of photoresponses in M2-
In addition to their innate photosensitivity, type pRGCs, with similar photoresponses observed
melanopsin-expressing pRGCs also receive syn- from these cells in the absence of melanopsin
aptic inputs from the outer retina, via connections expression (Schmidt and Kofuji, 2010). Similar
with bipolar cells and amacrine cells, and act as observations have been reported for M3-type
the principle conduit for delivering rod cone pRGCs (Schmidt et al., 2011) and are likely also
driven information to non-image forming areas true for M4- and M5-type pRGCs which have even
of the brain (Belenky et al., 2003; Guler et al., lower levels of melanopsin expression and elicit
2008; Hatori et al., 2008; Ostergaard et al., 2007; even smaller photoresponses than M2-type cells.
Vugler et al., 2007; Wong et al., 2007). As a result, Given the lower levels of innate photosensitivity
the photoresponses elicited by individual pRGCs compared to M1-type cells, it would appear that
are ultimately a fusion of signals generated from the functioning of M2-type cells, M3-type cells,
within rods, cones, and also the pRGCs them- and presumably M4- and M5-type cells are more
selves. As indicated by the differing location of closely linked to that of the classical photoreceptors
their dendrites in either the ON or OFF layers compared to M1-type cells.
of the IPL, it is now clear that the pRGC subtypes
receive different inputs from the outer retina. M2-
type pRGCs receive inputs from ON bipolar cells, pRGC subtypes mediate different physiological
as might be expected. Interestingly, M1-type responses to light
pRGCs receive not only OFF inputs but also
anomalous ON inputs from synaptic connections In addition to differences in cell morphology and
with cone ON bipolar cells as their dendrites pass the nature of retinal connections, there is also evi-
through the ON layer en route to the OFF layer dence that pRGC subtypes differentially innervate
of the IPL and also within the OFF layer of the specific retino-recipient brain areas (Baver et al.,
IPL itself (Dumitrescu et al., 2009; Hoshi et al., 2008; Chen et al., 2011; Hattar et al., 2002, 2006;
2009; Wong et al., 2007). Indeed, it would appear McNeill et al., 2011) and would therefore seem to
that despite the pattern of stratification, the ON be tasked with performing different physiological
pathway exerts larger effects on M1 cells than roles (Chen et al., 2011: for review see Do and
the OFF pathway (Schmidt and Kofuji, 2010; Yau, 2010; Schmidt et al., 2011). M1 cells predomi-
Wong et al., 2007). In addition to contacts with nately innervate the SCN and shell of OPN (Hattar
bipolar cells, pRGCs are also known to receive et al., 2006; McNeill et al., 2011), with these
inhibitory inputs from amacrine cells, yet again structures each innervated by distinct subtypes of
there are cell type-specific differences. M1 cells M1-type pRGCs (Chen et al., 2011). Brn3b nega-
synapse with dopaminergic amacrine cells in the tive M1-type pRGCs comprise the primary input
OFF layer of the IPL (Ostergaard et al., 2007; to the SCN and contribute to circadian entrainment
Viney et al., 2007; Vugler et al., 2007), whereas whereas Brn3b positive M1-type pRGCs innervate
M2-type cells appear to synapse with a mono- the core of the OPN and contribute to the pupillary
stratified amacrine cell type (Viney et al., 2007). light response (Chen et al., 2011). By comparison,
It is now clear that signals from the outer retina non-M1-type pRGCs (collectively) show little
23

innervation of the SCN, but do innervate the core activation of melanopsin has received far less
region of the OPN, the SC, and the dLGN (Baver attention, and to date the precise cellular
et al., 2008; Brown et al., 2010; Chen et al., 2011; mechanisms of phototransduction in pRGCs have
Ecker et al., 2010; McNeill et al., 2011). However, not been fully characterized. Until recently, our
by using currently available transgenic models, it understanding of the melanopsin signaling
is not possible to selectively label the different pathway had been informed largely by pharmaco-
non-M1 classes of pRGCs, and little is known logical studies performed in retinal explants,
regarding the specific regions of the brain innerva- disassociated pRGCs, and cellular expression sys-
ted by M2-, M3-, M4-, or M5-type pRGCs or the tems. Overall, the data from these studies suggest
specific functions associated with these cell types. a melanopsin signaling pathway starting with the
Data from retrograde labeling-based experiments activation of a Gq/11-type G-protein, leading
indicate that M2-type pRGCs (detected by to PLC activation and ultimately the influx of
melanopsin antibodies but not b-gal antibodies in Ca2 þ through TRP-type ion channels in the cell
the tau-lacZ-Opn4 mouse that reports only M1- membrane, leading to the generation of action
type cells) innervate the core of the OPN and the potentials (for review see Do and Yau, 2010 and
SC (Baver et al., 2008). Based on their innervation Hankins et al., 2008) (Fig. 2). Interestingly, all of
of the SC, it would seem that non-M1 cells, includ- the components necessary for generating light
ing M2-type pRGCs contribute to melanopsin- responses within pRGCs appear to be localized
dependent light avoidance and photophobia to a membrane-bound signaling complex, and
(Johnson et al., 2010; Semo et al., 2010) and recent the classical endpoint of the Gq/11 signaling path-
studies have suggested a role for non-M1-type way, namely, inositol 1,4,5-triphosphate (IP3)-
pRGCs in image-forming pathways and visual pro- dependent release of calcium from intracellular
cessing (Brown et al., 2010). Further work will stores, does not seem necessary for transduction
be needed to establish the precise pattern of of light stimuli by pRGCs (Graham et al., 2008;
innervations associated with each pRGC subtype, Hartwick et al., 2007; Sekaran et al., 2003).
and in turn, this information will undoubtedly From these studies, it is clear that there are
provide invaluable insight into the physiological striking differences in the phototransduction cas-
functions performed by each class of pRGC. cade employed by melanopsin-expressing pRGCs
and rod and cone photoreceptors. Melanopsin
employs a different G-protein signaling mec-
Melanopsin phototransduction hanism to the rods and cones and melanopsin
activation ultimately results in membrane depo-
The phototransduction signaling pathway of rods larization rather than cellular hyperpolarization.
and cones has been characterized extensively. It has been suggested that the melanopsin signaling
Absorption of light by an opsin/vitamin A-based cascade may be analogous to the invertebrate pho-
visual pigment leads to activation of the totransduction pathway present in rhabdomeric
G-protein transducin (a member of the Gi/o class photoreceptors (Berson, 2007; Contin et al., 2006;
of G-proteins), resulting in phosphodiesterase Graham et al., 2008; Hankins et al., 2008; Isoldi
activation and hydrolysis of cGMP to GMP et al., 2005).
(Lamb, 1996; Luo et al., 2008). Falling levels of Although these initial studies provided an
cGMP lead to closure of cyclic nucleotide-gated important framework to consider melanopsin
(CNG) cation channels and result in a membrane phototransduction, the specific identity of almost
hyperpolarization. all, if not all, the components involved remained
By contrast, the phototransduction signaling to be determined beyond doubt. More recently,
pathway initiated within pRGCs following transgenic approaches have been used to identify
24

TRPC6/7 VOCC
4 5
1

PIP2
Na+
Gq/11 Ca2+

2 β γ
PLCb4 X
3 IP3 + DAG
X

ER Ca2+
release

Fig. 2. The melanopsin signaling pathway. Schematic representation of the melanopsin signaling pathway based on available
evidence. Melanopsin activation (Step 1) leads to activation of a Gq/11-type G-protein (Step 2) followed by activation of PLCb4
(Step 3) and ultimately the downstream opening of TRPC6/7-type channels (Step 4) followed by VOCC activation and
action potential generation (Step 5). The effectors of PLCb4 are not clear, but seem not to involve IP3-mediated release of Ca2 þ
from internal stores or the direct activity of DAG (as indicated by dashed blue lines). PLC, phospholipase C; PI3K,
phosphatidylinositol 3-kinase; PIP2, phosphatidylinositol bisphosphate; IP3, inositol triphosphate; DAG, diacylglycerol; ER,
endoplasmic reticulem; VOCC, voltage-operated calcium channel.

a number of key components in the melanopsin invertebrate opsins than the classical vertebrate
signaling pathway, and have shown further visual opsins (Koyanagi and Terakita, 2008;
similarities between invertebrate and melanopsin Provencio et al., 1998). The deduced structure
phototransduction. In the following sections, we of melanopsin shows several features that are
will consider, in turn, the evidence for each stage reminiscent of invertebrate opsins, including the
of the signaling cascade and highlight recent presence of a tyrosine instead of a glutamate
advances and also areas where clear omissions counterion (E113), an elongation of the third
in our understanding remain. cytoplasmic loop, and an extended intracellular
C terminus (Provencio et al., 1998). In vivo
melanopsin binds 11-cis retinal as its chromo-
Step 1: Light absorption by melanopsin phore and converts this to all-trans retinal follow-
photopigment ing absorption of a photon (Walker et al., 2008).
Interestingly, there is strong evidence that
Melanopsin is a seven transmembrane G-protein melanopsin is capable of acting as a bistable pho-
coupled receptor (GPCR) with all the key struc- topigment (Melyan et al., 2005; Mure et al., 2007;
tural features of an opsin (Provencio et al., Panda et al., 2005), and is capable of regenerating
1998). Phylogenetic analysis suggests that mam- 11-cis chromophore from all-trans retinal. How-
malian melanopsin is more closely related to the ever, recent studies of the zebrafish melanopsin
25

family suggest that bistability is not a universal 2007). In mammals, only one melanopsin gene
property of vertebrate melanopsins (Davies has been identified, but again splice variants have
et al., 2011). The photobiology and visual cycle been reported. Two distinct isoforms of mouse
of melanopsin has been reviewed in detail melanopsin, Opn4L and Opn4S, are generated
elsewhere (Do and Yau, 2010) and shall not be by alternative splicing of the murine Opn4 gene
considered here. (Pires et al., 2009). These two isoforms of
The general consensus is that the peak spectral melanopsin differ only at the distal C-terminal
sensitivity of mammalian melanopsin is in the regions and are differentially expressed in M1-
region of 480 nm. There is strong evidence for and M2-type pRGCs of the adult mouse retina.
this from studies of melanopsin-driven light M1-type cells express both Opn4L and Opn4S
responses in cell line models (Panda et al., 2005; whereas only Opn4L can be detected in M2-type
Qiu et al., 2005), and also pRGCs (Berson cells. However, the functional relevance of these
et al., 2002; Dacey et al., 2005; Tu et al., 2005), distinct isoforms is currently unclear.
behavioral responses in animal models (Gamlin
et al., 2007; Hattar et al., 2003; Lucas et al., 2001)
and physiological measurements from human Step 2: Activation of a G-protein signaling
patients (Brainard et al., 2001; Hankins and Lucas, pathway
2002; Thapan et al., 2001; Zaidi et al., 2007).
However, there are some inconsistencies in the Once activated by light, melanopsin, like other
literature with reports of peak sensitivity for GPCRs, binds to specific G-proteins to activate
melanopsin ranging from 420 to 500 nm. Some of intracellular signaling cascades. To date, approx-
these results originate from in vitro cell expression imately 20 Ga subunits have been identified.
systems that may potentially be sensitive to They are grouped into four functionally distinct
artifacts (Melyan et al., 2005; Newman et al., subclasses: Gs, Gi/o, Gq/11, and G12/13, based on
2003). The first successful absorption spectra of the nature of the signaling pathway initiated
mouse melanopsin suggested a peak absorbance (Hubbard and Hepler, 2006; Mizuno and Itoh,
of 424 nm (Newman et al., 2003), while other 2009). By analogy to rhabdomeric transduction,
reports have indicated values of around 500 nm it was proposed that melanopsin might be specif-
(Walker et al., 2008). Spectrophotometry analysis ically coupled to the Gq/11 class of G-proteins and
of melanopsin photopigments from other species downstream activation of PLC (Provencio et al.,
have indicated a peak sensitivity in the region of 1998; Warren et al., 2003). Since then, several
480 nm, including cephalochordate (Koyanagi studies have utilized retinal explants or disso-
et al., 2005) and zebrafish (Davies et al., 2011). ciated ganglion cell cultures to study the nature
Collectively, it seems likely that the peak sensitiv- of melanopsin G-protein interactions within
ity of melanopsin is close to 480 nm, but direct pRGCs (Graham et al., 2008; Hartwick et al.,
confirmation of this fact is hindered by the techni- 2007; Sekaran et al., 2003, 2005, 2007; Tu et al.,
cal difficulties in obtaining high-quality spectro- 2005; Warren et al., 2006). Current pharmacolog-
photometry data for this class of opsin, most ical evidence does indeed suggest a melanopsin
likely related to issues of protein misfolding in signaling pathway involving the activation of a
overexpression systems. Gq/11-type G-protein (Table 1). Most striking is
In nonmammalian species, numerous distinct the finding that application of GPant-2a, a spe-
forms of melanopsin have been reported, cific inhibitor of the Gq/11 class of G-proteins,
encoded not only by separate genes, but also completely abolishes melanopsin light responses
alternate splicing (Bellingham et al., 2006; Davies in pRGCs whereas the related peptide GPant-2,
et al., 2011; Drivenes et al., 2003; Torii et al., a specific inhibitor of the Gi/o class, has no effect
26

Table 1. Summary of evidence that melanopsin phototransduction involves Gq/11-type G-proteins

Reference Cell type Assay Agent/targets Effect

Warren et al. pRGCs/retina Patch clamping GTPgS (stimulatory G-protein) Blocked response
(2006) explant GDPbS (inhibitor of GTP Blocked response
binding)
IBMX and cAMP or cGMP No effect
Hartwick et al. Isolated pRGCs Patch clamping Ca2 þ Removal of external Ca2 þ Blocked response
(2007) imaging Thapsigargin No effect
Graham et al. Isolated pRGCs Patch clamping GDPbS (inhibitor of GTP Blocked response
(2008) binding)
GPAnt-2a (inhibits Gq/11 class) Blocked response
U73122 (PLC antagonist) Blocked response
Thapsigargin No effect
GPAnt-2 (inhibits Gi/o class) No effect
PTX (inhibits Gi/o class) No effect
Cholera toxin (inhibits Gs class) No effect
OAG (active DAG analogue) No effect
Panda et al. hOpn4 in Xenopus Patch clamping Gnaq/11 antibodies Blocked response
(2005) oocytes U73122 (PLC antagonist) Blocked response
PTX (inhibits Gi/o class) No effect
Gnas, Gnai antibodies No effect
Qiu et al. (2005) mOpn4 in HEK293 Patch clamping Ca2 þ GDPbS (inhibits GTP binding) Blocked response
imaging GPant-2a (inhibits Gnaq/11 class) Blocked response
U73122 (PLC antagonist) Blocked response
PTX (inhibits Gi/o class) No effect
Melyan et al. hOpn4 in Neuro-2A Patch clamping Suramin (G-protein inhibitor) Blocked response
(2005) Thapsigargine Blocked response
8-Br-cGMP Blocked response
GTPgS (stimulatory G-protein) Induced response
U73122 (PLC antagonist) No effect
NF023 (inhibits Gi/o class) No effect

(Graham et al., 2008). In addition to studies of cells (Qiu et al., 2005), whereas blockers of Gi/o
native pRGCs, melanopsin G-protein signaling fail to inhibit melanopsin-dependent light
has also been investigated using heterologous responses in Neuro-2A cells (Melyan et al.,
expression in a variety of cell lines (Giesbers 2005). By contrast, one study has reported that
et al., 2008; Kumbalasiri et al., 2007; Melyan melanopsin can couple to transducin in a bio-
et al., 2005; Panda et al., 2005; Qiu et al., 2005). chemical assay (Newman et al., 2003), yet there
In general, the pharmacological data obtained is no evidence of transducin expression in
from these in vitro models again indicate that pRGCs and melanopsin-based signaling is nor-
melanopsin couples to Gq/11-type G-proteins mal in Gnat/ mice (Hattar et al., 2003). Collec-
(Table 1). For example, melanopsin responses tively, there is relatively strong evidence that
are greatly attenuated by antibodies against melanopsin couples to Gq/11-type G-proteins
Gq/11-type G-proteins but not by antibodies in vitro and in vivo, and a general consensus has
against Gi/o G-proteins (Panda et al., 2003). formed that this is indeed the case. Yet despite
Gq/11 agonists fully blocked the melanopsin- the growing evidence most, if not all, previous
dependent light responses in HEK293–TRPC3 studies have utilized pharmacological tools that
27

fail to distinguish between specific members of Step 3: Phospholipase C activation


the Gq/11 family (Table 1). As a result, the precise
identity of the G-protein subunit(s) involved in The Gq/11 signaling pathway classically leads to
melanopsin phototransduction remains to be the activation of PLCb isoforms and the genera-
determined. The Gq/11 family itself contains four tion of 1,2-diacylglycerol (DAG) and inositol
members: Gaq, Ga11, Ga14, and Ga15 (Ga16 in 1,4,5-trisphosphate (IP3) following the hydrolysis
humans) (Davignon et al., 1996; Wilkie et al., of phosphatidylinositol 4,5-bisphosphate (PIP2)
1992). All members of the Gq/11 subclass share a (Hubbard and Hepler, 2006; Mizuno and Itoh,
high degree of sequence homology and all exert 2009). Pharmacological evidence supports a role
their effect via the activation of the effector for PLC activity in melanopsin-driven light
enzyme PLCb (Hubbard and Hepler, 2006; Rhee, responses, both in pRGCs and cell line models
2001; Weschler et al., 2006). Gaq and Ga11 are (Table 1). Most notable is the observation that
expressed in almost all cell types (Wilkie et al., pRGC responses are blocked by the PLC inhibi-
1991) and share a near complete functional over- tor U73122, but not the inactive homolog
lap with few, if any, receptors discriminating U73343 (Graham et al., 2008). In addition, PLC
between them (Hubbard and Hepler, 2006; activity has been shown to be an important
Wettschureck and Offermanns, 2005). In com- step in melanopsin signaling in a range of non-
parison, the expression profile of Ga14 is far mammalian species, including chicken (Contin
more restricted (Wilkie et al., 1991) and typically et al., 2006), Xenopus (Isoldi et al., 2005), and
observed only in specialized cell types (Shindo Amphioxus (Nasi and del Pilar Gomez, 2009).
et al., 2008; Tizzano et al., 2008; Yule et al., However, the pharmacological evidence supports
1999). The pattern of expression observed for only the involvement of a PLC-type enzyme,
Ga14 seems to suggest specific roles in cell signal- and does not provide information regarding the
ing, however, to date there is limited information subclass of PLC involved, but by inference and
regarding the functional differences between logical assumption it is predicted to be a member
Ga14 and Gaq/11 (Nakamura et al., 1996). Ga15 of the PLCb subclass.
is a highly promiscuous G-protein subunit whose To date, four isoforms of PLCb have been
expression is almost exclusively restricted to reported. PLCb1 and PLCb3 are widely
hematopoietic cell types (Amatruda et al., 1991; expressed, whereas the expression of PLCb2 is
Davignon et al., 1996; Kostenis et al., 2005; largely restricted to hematopoietic cells. Interest-
Wilkie et al., 1991). Our own qPCR and micro- ingly, expression of PLCb4 is highest in the
array data (Steven Hughes, unpublished data), as retina, with expression highest in cones and gan-
well as several previous studies have shown that glion cells (Adamski et al., 1999; Ferreira and
Gaq, Ga11, and Ga14 but not Ga15 are expressed Pak, 1994; Ferreira et al., 1993). The profile
in the mouse retina (Graham et al., 2008; Peirson and efficiencies with which Gaq, Ga11, and Ga14
et al., 2007) and single-cell PCR studies have activate PLCb isoforms are very similar. All are
confirmed the expression of mRNA for these potent activators of PLCb1, PLCb3, and PLCb4,
G-proteins in pRGCs (Graham et al., 2008), with and all show relatively weak activation of PLCb2
Ga14 the most commonly detected. As a result, it (Hubbard and Hepler, 2006). Graham et al.
is theoretically possible that melanopsin may (2008) confirmed the expression of specific PLCb
couple to Gaq, Ga11, or Ga14 within pRGCs, or a isoforms in pRGCs using single-cell RT-PCR, and
combination of these G-protein subunits. Further indicated that levels of expression are highest for
work is required to determine the precise identity PLCb4 compared to other isoforms (Graham
of the G-protein subunit(s) involved in melanopsin et al., 2008). It is interesting to note that PLCb4
phototransduction. is 50% identical to the Drosophila PLCb norpA
28

protein, a value higher than that shared with the Drosophila can be influenced directly by polyun-
other mammalian PLCb isoforms (35–40%) saturated fatty acids (PUFAs), produced by the
(Hubbard and Hepler, 2006; Rhee, 2001). Thus, breakdown of DAG (Chyb et al., 1999b). By anal-
based on comparison with the invertebrate signal- ogy to the invertebrate cascade, it is possible that
ing cascade, PLCb4 is predicted to be the media- the activity of the light-gated channels in pRGCs
tor of melanopsin signaling. These assumptions may be influenced by PLCb4 activity via numerous
have been confirmed recently. Endogenous mechanisms, including the depletion of PIP2
photoresponses are almost entirely abolished in and the generation of PUFAs. Further work is
M1-type pRGCs from PLCb4/ mice, with only required to determine the mechanisms by which
a small residual response remaining representing PLCb4 influences the activity of ion channels posi-
< 1% of the wild-type response (Xue et al., 2011). tioned downstream in the melanopsin signaling
The downstream effectors of PLCb activity in cascade.
pRGCs are unclear, and little to nothing is known
concerning the mechanisms by which second
messengers influence the activity of the light- Step 4: TRP channel activation
sensitive ion channels. Typically, PLCb activation
leads to the hydrolysis of PIP2 to form IP3 and Electrophysiological recordings from pRGCs
DAG. However, application of DAG analogues (both whole cell and excised membrane patches)
to membrane patches excised from pRGCs fails have provided good evidence for the involvement
to induce or modify responses from pRGCs of a TRP-like channel mediating the initial cellular
suggesting that DAG itself is not required for depolarization of pRGCs following light stimula-
melanopsin phototransduction (Graham et al., tion. For example, the light-induced photocurrent
2008; Warren et al., 2006). In addition, intracellu- reverses close to 0 mV, shows inward and outward
lar application of IP3 analogues neither induces rectification, and has a small single-channel con-
nor blocks photoresponses in pRGCs (Graham ductance ( 1 pA) (Do et al., 2009; Schmidt and
et al., 2008; Hartwick et al., 2007), suggesting that Kofuji, 2009), and calcium seems to be the signifi-
IP3 is also not required. However, the generation cant charge carrier through the light-sensitive
of IP3 and DAG results in a depletion of PIP2 channels (Warren et al., 2006). These features are
from the cell membrane. It is possible that the all consistent with involvement of a TRP-like ion
light-sensitive ion channels in pRGCs are gated channel.
directly by interactions with PIP2 itself. PIP2 is Again, by analogy to the invertebrate cascade,
known to modulate the gating of several classes and the biophysical characteristics of the light
of ion channels (Hardie, 2007; Suh and Hille, currents recorded from pRGCs, initial
2008), including the light-gated channels in Dro- investigations focused on members of the TRPC
sophila (Hardie, 2003; Raghu, 2006; Suh et al., subfamily of TRP channels. Data from pharma-
2006). There is some evidence that pRGCs light cology-based studies suggested that the most
responses may indeed be directly gated by PIP2. likely candidates for this TRP-like channel are
Pharmacological inhibition of PIP2 synthesis (using indeed members of the TRPC channel subfamily,
wortmannin) leads to prolonged light responses in potentially TRPC3, TRPC6, or TRPC7 (Hartwick
pRGCs (Graham et al., 2008). It is possible that et al., 2007; Sekaran et al., 2007; Warren et al.,
the prolonged responses are caused by a delay in 2006). Light-evoked responses in pRGCs are
response termination due to a delay in restoring suppressed by lanthanides, ruthenium red, and
levels of PIP2 at the cell membrane and a failure SKF-96365, and are enhanced by flufenamic
to close the activated ion channels. There is acid and 1-oleoyl-2-acetyl-sn-glycerol (Hartwick
also evidence that the light-gated channels in et al., 2007; Warren et al., 2006). Further,
29

2-aminoethoxydiphenyl borate (2-APB), an mechanisms by which PLCb4 activation drives


inhibitor of the TRPC subclass of TRP channels, the opening of these channels and transduction of
is an extremely potent inhibitor of mouse pRGC the melanopsin signal.
light responses in vitro (Hartwick et al., 2007;
Sekaran et al., 2007), and also in vivo where
2-APB induces an acute reduction of the pupil- Step 5: Activation of voltage-gated ion channels
lary light reflex (Sekaran et al., 2007). and action potential firing
Immunolocalization studies indicate the expres-
sion of TRPC6 and TRPC7 but not TRPC3 in The initial calcium influx through the TRPC
pRGCs (Sekaran et al., 2007; Warren et al., channels results in cellular depolarization and
2006), and so these channels represent the prime ultimately leads to the activation of voltage-gated
candidates for the depolarizing current observed ion channels and the generation of action
in pRGCs. Despite these early conclusions, potentials (Graham et al., 2008; Hartwick et al.,
subsequent work suggested that the light-induced 2007; Warren et al., 2006). Initial pharmacology
channels may not be members of the TRPC chan- suggests the involvement of L-type VOCCs acting
nel subfamily. pRGC photoresponses were found downstream of TRP channel activation, as verap-
to be largely unaltered in mice lacking TRPC3, amil and cis-diltiazem significantly reduced light
TRPC6, and TRPC7, suggesting that these responses recorded from pRGCs (Hartwick
channels are not required for the generation of et al., 2007). However, it is clear that the genera-
light responses in pRGCs (Perez-Leighton et al., tion of action potentials requires the cooperative
2011). However, this study only reported the effect action of numerous classes of ion channel, includ-
of removing each of these channels in isolation, ing voltage-gated calcium, sodium, and potassium
and could not exclude the possibility that multiple channels, as well as calcium-sensitive channels,
TRPC-like channels are involved in melanopsin inward rectifiers, and background leak-type
phototransduction, potentially as functional het- channels. The precise identity of the ion channels
eromeric channels (Schaefer, 2005). Recently, that contribute to the generation and regulation
the identity of the ion channels responsible for of action potentials in pRGCs has not been fully
passing the light-sensitive current in pRGCs has determined.
been determined, at least for M1-type pRGCs. Membrane potential is, by definition, a key fac-
Photoresponses in these cells are abolished in tor in determining the activation of voltage-gated
double knockout mice lacking both TRPC6 and channels and the progression to action potential
TRPC7, but were again unaltered in single knock- firing, and as such plays a key role in regulating
out mice lacking TRPC3, TRPC6, or TRPC7 or the excitability and sensitivity of neuronal cells.
various other members of the TRPC subfamily in Single photon responses, that generate small
isolation or combination. Overall, these results currents of only few pA, are sufficient to induce
indicate that elimination of both TRPC6 and spike activity in pRGCs (Do et al., 2009) and
TRPC7 is required to eliminate photoresponses spontaneous firing of pRGCs is observed in
in M1-type pRGCs (Xue et al., 2011), although it the dark (Berson et al., 2002; Do et al., 2009;
remains to be shown conclusively whether these Schmidt et al., 2008; Tu et al., 2005). From these
channels are performing redundant roles or acting observations, it appears that the resting mem-
as functional heteromers. The definitive identifica- brane potential of pRGCs is tightly regulated
tion of the light-sensitive channels in pRGCs will and held very close to the threshold for action
no doubt prove to be an important step in our potential firing. However, the mechanisms that
understanding of melanopsin phototransduction, control resting membrane potential and cellular
and will allow more focused studies of the excitability in pRGCs remain poorly understood.
30

pRGCs are known to receive excitatory synap- and the PLC isoform and downstream TRP
tic inputs from bipolar cells and inhibitory inputs channels involved in melanopsin signaling are
from amacrine cells (Schmidt and Kofuji, 2009, the closest homologues of the respective com-
2010, 2011; Wong et al., 2007). These inputs influ- ponents in the Drosophila cascade. However,
ence the membrane potential of pRGCs by the we still only have a very basic model of the
generation of cation- and chloride-based con- melanopsin signaling pathway naive to the
ductances, respectively (Wong et al., 2007). pRGCs. The precise identity of the Ga subunit
In darkness, inhibitory inputs dominate over (s) involved still remains unknown, as does
excitatory inputs and contribute to setting the the identity and function of the Gb subunits.
resting membrane potential of pRGCs below the The intermediary step(s) in the signaling path-
threshold for action potential firing (Wong et al., way, particularly those linking PLCb4 activation
2007). However, the nature and influence of these and ion-channel opening, remain to be deter-
inputs varies between pRGC subtypes (Schmidt mined. In addition, based on our knowledge of
and Kofuji, 2009, 2010, 2011) and would seem to G-protein-signaling events, and also by more
explain in part the differences in resting mem- specific comparison to the invertebrate signaling
brane potential and excitability observed between cascade, it is clear that there are a number of
these cells. However, pRGC subtypes continue to important gaps in the current model. Most nota-
show differences in resting membrane potential ble is the lack of data concerning the regulation
under conditions where all synaptic inputs are and termination of the melanopsin signaling path-
blocked (Schmidt and Kofuji, 2009, 2010, 2011). way (Fig. 3).
Under these conditions, M1-type cells remain sig-
nificantly more depolarized than M2-type cells
(and M3- and M4-type cells) and therefore sit Role of protein kinases: Desensitization,
closer to the threshold for action potential firing. adaptation, and termination of melanopsin
These differences are seemingly the result of signaling
intrinsic differences in levels of baseline ion-
channel activity in these cells. However, the The most striking gap in our knowledge of the
mechanisms which contribute to the intrinsic rest- melanopsin signaling pathway is the lack of any
ing membrane potential of the pRGC subtypes description of protein kinase activity. Many clas-
remain largely unexplored. ses of protein kinase exist, and these have wide-
spread functions and are key to regulating the
activity of second messengers, membrane-bound
Gaps in the current model of melanopsin proteins and ion channels and ultimately act to
phototransduction fine tune the properties of cellular responses. Pro-
tein kinases may potentially exert an influence
With recent discoveries, our understanding of over every stage of the melanopsin signaling path-
the melanopsin signaling pathway has improved. way. PKA and PKC represent two of the major
There is now good evidence that melanopsin classes of protein kinases, and form core compo-
couples to Gnaq/11-type G-proteins, leading to nents of cell signaling pathways that are typically
the activation of PLCb4 and ultimately the influx involved in the regulation of GPCR signal adap-
of calcium through TRPC6 and TRPC7 channels. tation and desensitization, however little data
These results confirm the assumption that exists concerning the role of PKA or PKC in
melanopsin uses an invertebrate-like signaling melanopsin signaling.
cascade. The G-proteins involved in both pat- PKC activation is driven by DAG in a process
hways belong to the same class of G-proteins requiring Ca2 þ, and is a key feature of Gq/11-type
31

TRPC6/7 VOCC
4 5
1

Arr PIP2 PIP2


? Na
+
β γ
Gq/11 2+
GRK Ca
β γ ?
2
PLCb4 X
Gq
? IP3 + DAG
β γ 3
β γ
?
X
PKC
G11
?
β γ PLC AC PI3K ER Ca2+
? release PUFA Confirmed components and interactions
Interactions not involved
G14 PKA cAMP ? Predicted by analogy
?
β γ Common to Gq/11 signalling cascades
? ? PKCz ? Potential interactions

Scaffold proteins

Fig. 3. A potential melanopsin signaling pathway. Representation of a potential model of melanopsin phototransduction.
Components and interactions highlighted in solid blue represent confirmed components of the melanopsin signaling pathway in
pRGCs. Blue dashed lines indicate interactions or events that are known not to be involved. Items and interactions highlighted
in red represent components of the pathway that are missing or as yet unidentified, based on existing knowledge of G-protein
signaling events (i.e., arrestin and GRK activity) and also based on a direct comparison to the invertebrate signaling cascade
(i.e., involvement of PKC- and scaffold-type proteins). Items outlined in black represent signaling interactions known to occur
in classical Gq/11-type cascades and may therefore potentially perform a role in melanopsin signaling, but for which there
is currently no direct evidence. Dashed black lines represent potential interactions. Arr, arrestin; PLC, phospholipase C; P13K,
phosphatidylinositol 3-kinase; PIP2, phosphatidylinositol bisphosphate; IP3, inositol triphosphate; DAG, diacylglycerol; ER, endoplasmic
reticulem, PKA, protein kinase A; PKC, protein kinase C; PKCz, protein kinase C zeta (atypical PKC); AC, adenylate cyclase; VOCC,
voltage-operated calcium channel.

signaling cascades. In invertebrate photoreceptors, and increase in incremental responses in the pres-
PKC activity influences response inactivation and ence of steady background illumination (Wong
adaptation but not excitation of the light response et al., 2005). However, the mechanisms, if any,
(Yau and Hardie, 2009). Interestingly, adaptation by which PKC regulate the desensitization and
and desensitization of pRGC responses have adaptation of the melanopsin signaling pathway,
also been reported. Adaptation is evident by and the stage in the signaling pathway where
shortening of time to peak response in response these regulatory mechanisms act are unknown.
to increasing intensities of light (Do et al., 2009), The identification of TRPC6 and TRPC7 as the
a reduction in amplitude of response following light-sensitive channels in pRGCs will no doubt
repeated light stimulation (Wong et al., 2005), accelerate our understanding of how PKC-based
32

signaling influences the activity of these channels irradiances, as well as impaired light-induced
and modulates the nature of light responses in gene expression in the SCN. PKCz is an atypical
pRGCs. PKC and lacks binding domains for DAG and
Activation of PKA is driven by increased Ca2 þ. The involvement of PKCz is therefore com-
cAMP, which in turn is driven by adenylate patible with these second messengers exerting
cyclase activity. cAMP is not normally associated little to no effect on melanopsin signaling. How-
with the activation of Gq/11-type signaling pat- ever, there is no direct evidence for a role of
hways, but instead is the classical mediator of PKCz in pRGC light responses and further stud-
Gs- and Gi-type G-proteins. However, it is now ies are required to confirm the role of PKCz in
clear that various isoforms of adenylate cyclase melanopsin phototransduction. By analogy to
are activated by direct interactions with free Gb the Drosophila phototransduction cascade, PKCz
subunits released from activated G-protein may influence TRP ion-channel activity via
complexes (see below), and thus a direct role participation in an INAD-like signaling complex
for cAMP and PKA in the melanopsin signaling that potentially also includes PLC, PKC, and
pathway cannot be excluded. However, it is PDZ-domain-containing scaffolding proteins
more likely that the cAMP–PKA pathway (Hardie, 2001).
may represent a mechanism of external regula- To date, research on melanopsin signaling has
tion. Dopamine-mediated activation of the focused on activation of the signaling pathway.
cAMP–PKA pathway, via the Gs-coupled Drd1a There are no data concerning the deactivation
receptor, leads to an inhibition of spike firing or termination of the melanopsin signaling cas-
in RGCs (Hayashida et al., 2009). Given the cade within pRGCs. G-protein-coupled receptor
known interactions between M1-type pRGCs kinases (GRKs) are a special class of protein
and dopaminergic amacrine cells (Vugler et al., kinases (serine/threonine kinases) involved in
2007; Zhang et al., 2008), dopamine and the modulation (desensitization) and termination of
cAMP–PKA pathway may represent an impor- GPCR signaling events. GRK phosphorylation
tant mechanism for regulating pRGC function. prepares the activated receptor for arrestin bind-
Interestingly, there is evidence that cAMP and ing, and ultimately termination of G-protein
PKA activity modulate the properties of light signaling activity. Subtypes of GRKs and arrestins
responses generated in Drosophila photo- are differentially expressed in rod and cone pho-
receptors (Chyb et al., 1999a; Hardie and Raghu, toreceptors, where they perform essential roles
2001), However, currently there is no direct evi- in visual phototransduction and influence the
dence for PKA-based modulation of melanopsin nature of light responses (Makino et al., 2003).
signaling in pRGCs. Arrestins are known to perform important roles
There is evidence for at least one class of pro- in shaping the properties of light responses
tein kinase acting in the melanopsin signaling in Drosophila (Hardie, 2001), and overexpression
pathway; PKCz. The potential role for PKCz of arrestins (mouse b1 and b2 arrestin and
was first highlighted by microarray-based exper- also Drosophila arrestin-2) has been shown to
iments that investigated changes in gene tran- enhance photocurrents recorded from Xenopus
scription within the eyes of mice lacking rod and oocytes expressing melanopsin (Panda et al.,
cone photoreceptors following acute light stimula- 2005). To date, the functional roles of GRKs
tion. Remarkably, Prkcz/ mice were found to and arrestin molecules in melanopsin photo-
precisely phenocopy Opn4/ mice. They show transduction within pRGCs remain undeter-
attenuated circadian phase-shifting responses to mined. These molecules represent obvious
light, reduced period lengthening under constant candidates as regulators of melanopsin signaling
light, and attenuated pupillary responses at high and pRGC function.
33

Scaffold proteins Current evidence and data from our own gene
array studies (Hughes, S. & Butler, R., unpub-
Patch clamp experiments from excised membrane lished data) indicate that all five Gb subunits are
patches demonstrate that all the components expressed in the mouse retina. Significantly, these
required to initiate light responses within pRGCs different Gb subunits show distinct patterns of
are tightly bound to the plasma membrane. expression within cell types of the retina. Gb1 is
However, the nature of the signaling complex pro- expressed in rods and amacrine cells but not cones,
tein present in pRGCs, and how scaffold-type and Gb3 is expressed in all classes of cones, and a
proteins aid the assembly and function of this com- subset of bipolar cells but is absent from rods (Lee
plex remains unclear. Based on analogy to the et al., 1992; Peng et al., 1992). Deletions and
invertebrate signaling cascade, we might expect mutations of Gb1 and Gb3 have been associated
the involvement of an Inad-like scaffold protein with retinal degeneration (Kitamura et al., 2006;
(Hardie and Raghu, 2001; Raghu, 2006), and Tummala et al., 2006). Expression of Gb5 is
indeed previous microarray analysis has suggested restricted to ON bipolar cells and performs an
that the mammalian homologue INADL, and also important role in bipolar light responses (Rao
RACK-1 and NHERF-1 scaffold-type proteins et al., 2007) and the long isoform splice variant of
may be candidates for this role (Peirson et al., Gb5 (Gb5L) is restricted to rods where again it
2007). Further work will be required to determine contributes to the nature of light responses
the identity and role of the scaffold protein(s) that (Krispel et al., 2003; Watson et al., 1996). Gb
participate in melanopsin phototransduction. subunits also contribute to light responses in Dro-
sophila photoreceptors, acting to regulate levels
of spontaneous firing (Elia et al., 2005). However,
Gb subunits the data concerning the localization of Gb subunits
in the retina is limited and currently there are no
Functional G-protein complexes contain Ga, Gb, data concerning the identity of the Gb subunits
and Gg subunits. In addition to the activation of expressed within pRGCs, nor is there any data
Ga subunits, it is now clear that the Gb subunits concerning the roles that these subunits perform
also perform functional roles, and represent a in melanopsin phototrans-duction.
mechanism for generating secondary signaling pat- Interestingly, recent studies have reported the
hways within G-protein-signaling cascades. Gb presence of a secondary “minor” signaling path-
subunits can influence directly the activity of ion way within M1-type pRGCs that is dependent
channels, certain isoforms of adenylate cyclase on melanopsin but independent of PLCb4 and
and multiple classes and isoforms of PLC, includ- TRPC6/7 function. Small photoresponses persist
ing PLCb1, PLCb 2, and PLCb 3 but not PLCb4 in M1-type pRGCs from both PLCb4/ and
(Hubbard and Hepler, 2006; Mizuno and Itoh, TRPC6/7/ mice. These responses represent
2009; Rhee, 2001). It is possible therefore that < 1% of the wild-type response, but interestingly
Gb subunits may themselves participate in are completely absent in OPN4/ mice (Xue
melanopsin signaling, potentially activating sec- et al., 2011). The physiological relevance of this
ondary signaling pathways that are independent minor response, if any, is unknown. It is also
of PLCb4 activity. unclear whether this minor pathway is functional
Five different Gb subunits have been identified, in other pRGC subtypes, where potentially it
Gb1–5, (and 12 Gg subunits) (Cabrera-Vera et al., may exert more influence. It is possible that free
2003) and evidence from transgenic mouse models G-protein beta gamma subunits may contribute
suggests that the different Gb subunits perform to this secondary pathway, yet further work will
specific physiological roles (Dupre et al., 2009). be required to confirm this possibility.
34

Variable responses in pRGCs account, at least in part, for the differences


in levels of photosensitivity observed between
In addition to the obvious gaps in our understand- pRGC subtypes (Schmidt and Kofuji, 2009).
ing of pRGC phototransduction, it is also clear However, different response profiles observed
that the existing model fails to account for all between pRGC subtypes are not so easily
experimental observations, most notably those explained. The large fast-acting transient compo-
describing numerous distinct types of light nent of the response observed from M1-type cells
response from individual pRGCs. As discussed is likely the result of adaptation or saturation of
earlier, multiple distinct response types have been some component(s) of the phototransduction
recorded from pRGCs. Multiple response profiles pathway in these cells (Warren et al., 2003; Wong
are observed not only in adult tissue but also in et al., 2005), but the mechanisms involved have
early postnatal development prior to the emer- not been determined.
gence of rod and cone photoreceptors (Sekaran One potential explanation for the difference
et al., 2005; Tu et al., 2005). Further, distinct in light responses recorded from pRGCs is the
response profiles are recorded from pRGCs fol- differential expression of ion channels that act
lowing blockade of all synaptic inputs. Thus, it downstream in the signaling pathway, but ulti-
would seem that the mechanism for generating mately act to shape the properties of the mem-
diversity is inherent to the pRGCs themselves, brane currents and changes in membrane
and not dependent, at least not entirely, on the voltage. Interestingly, existing data suggest the
differential inputs from the outer retina. How- involvement of L-type VOCC channels in
ever, the cellular mechanisms responsible for melanopsin phototransduction, preceding the
generating this diversity are unknown. At pres- initial TRP channel-mediated depolarization
ent, we have one (incomplete) signaling model (Hartwick et al., 2007). However, L-type VOCCs
and as many as three distinct response types. Per- (but not T-type VOCCs) are reportedly
haps, multiple models of melanopsin signal trans- expressed only in OFF-type RGCs and are absent
duction are needed to explain these observations, from ON-type RGCs (Margolis et al., 2010), and
or alternatively, we need to identify the cellular therefore presumably absent from M2, M4, and
mechanisms that allow melanopsin signaling M5 pRGCs. This observation could explain, in
events to be modified in such a manner as to pro- part, the differences in photoresponses observed
duce multiple response profiles. The obvious and between these pRGC subtypes. Given the
intriguing question is whether or not melanopsin differences in resting potential of pRGC subtypes
couples to distinct and specialized signaling pat- and their different spike-firing capabilities it
hways within the different pRGC subtypes. would seem likely that these cells express differ-
Specific components of the signaling pathway ent profiles of ion channels, or at least the func-
have been identified for M1-type pRGCs, but it tioning of ion channels present in these cells are
is yet to be determined whether the same photo- differentially regulated.
transduction pathway occurs in other pRGCs Another, more intriguing possibility, is that the
subtypes. It is clear from recent evidence that dif- variant responses observed for M1- and M2-type
ferent classes of pRGC express different profiles cells are not entirely the result of differences in
of transcription factors (Chen et al., 2011), and downstream propagation of the melanopsin signal-
so it seems likely that different signaling compo- ing cascade, but instead are driven by differences
nents may be available for participation in in the function of melanopsin itself, perhaps due to
melanopsin-dependent signaling pathways. the expression of different isoforms of melanopsin
Levels of melanopsin expression and (Pires et al., 2009) or due to functional modi-
differences in resting membrane potential may fication of melanopsin proteins by glycosylation
35

(Fahrenkrug et al., 2009) and or phosphorylation most notably those describing the numerous dis-
(Pires et al., 2009; Provencio et al., 1998, 2000). tinct light responses observed from individual
The Opn4L and Opn4S isoforms of melanopsin dif- pRGCs. Overall, it is likely that many factors
fer only in their proximal C-terminal domain with contribute to the different functional responses
the extended C-terminus of Opn4L containing four observed from pRGC subtypes, including external
additional predicted PKC phosphorylation sites synaptic influences, differences in levels of
compared to the shorter tail of Opn4S (Pires et al., melanopsin expression, and differences in resting
2009). It seems likely that any functional differences membrane potential. It remains to be seen whether
between these isoforms may reside in their differen- differences also exist in the components of the sig-
tial sensitivity to PKC-based signaling, and this naling cascades themselves, or whether differential
may provide a mechanism to differentially regulate modulation of these components contributes to
melanopsin-based signaling within the pRGC the functional differences observed. Future work
subtypes. Further work will be needed to address must focus on the role of protein kinases in regula-
these possibilities. tion of melanopsin-driven signaling events, both
in response to external signals but also as part of
feedback mechanisms that are no doubt intrinsic to
Conclusions the melanopsin signaling pathway.
The melanopsin signaling pathway is broadly
Since the discovery of melanopsin-expressing similar to that observed in invertebrate photo-
pRGCs, there has been a rapid increase in our receptors. However, it is worth noting that
understanding of the anatomical variety of pRGC although there are clear similarities, the func-
subtypes; the brain regions that they innervate; tional properties of invertebrate photoreceptors
and the behavioral responses they regulate. and pRGCs are markedly different. Invertebrate
A growing body of evidence suggests that the photoreceptors are tasked with visual image pro-
pRGC subtypes are morphologically, anatomi- cessing and show the fastest signaling event of
cally, and functionally distinct with specific physi- any known GPCR, with arrestins performing an
ological roles. Although the precise functions essential role in accelerating inactivation (Hardie,
performed by each cell type remain to be 2001). Such properties are not required of the
fully determined. Our understanding of the melanopsin signaling pathway within pRGCs
melanopsin signaling pathway has also seen prog- tasked with providing background luminance
ress, but it is only very recently that the precise information. It is possible that the signaling cas-
identity of the cellular components has begun to cade of these two classes of photoreceptors has
emerge. Although there have been important diverged significantly, as has their function.
recent advances, there remain clear gaps in our In summary, it is clear that, despite recent
understanding of melanopsin phototransduction, advances, our understanding of the melanopsin
most notably the identity of the Ga subunit(s) signaling cascade remains a work in progress.
involved remains to be confirmed, as does the Microarray-based techniques have generated a
mechanisms by which PLCb activity leads to gat- number of potentially important candidates
ing of the downstream ion channels, seemingly (Peirson et al., 2007), and the confirmation that
TRPC6 and TRPC7. It is also clear that the cur- the melanopsin signaling cascade shows many
rent model fails to describe any aspects of the similarities to invertebrate signaling cascades will
regulatory pathways that undoubtedly exist, or surely accelerate the identification of other compo-
the mechanisms that regulate inactivation of nents of this pathway based on analogy and func-
melanopsin activity. Further, the existing model tional comparisons. However, validating these
fails to account for all experimental observations, candidates in the absence of robust pharmacological
36

tools remains a serious challenge. Although these Brainard, G. C., Hanifin, J. P., Greeson, J. M., Byrne, B.,
issues can be partially approached with transgenic Glickman, G., Gerner, E., et al. (2001). Action spectrum
for melatonin regulation in humans: Evidence for a novel
models and gene ablation, the study of this complex circadian photoreceptor. The Journal of Neuroscience, 21,
system will surely benefit from the application of 6405–6412.
conditional and inducible transgenic models in addi- Brown, T. M., Gias, C., Hatori, M., Keding, S. R., Semo, M.,
tion to use of techniques such as in vivo RNA Coffey, P. J., et al. (2010). Melanopsin contributions to irra-
interference. diance coding in the thalamo-cortical visual system. PLoS
Biology, 8, e1000558.
Cabrera-Vera, T. M., Vanhauwe, J., Thomas, T. O.,
Acknowledgments Medkova, M., Preininger, A., Mazzoni, M. R., et al.
(2003). Insights into G protein structure, function, and regu-
The authors would like to thank the EBRS for lation. Endocrine Reviews, 24, 765–781.
Chen, S. K., Badea, T. C., & Hattar, S. (2011). Photo-
the opportunity to contribute to this volume. entrainment and pupillary light reflex are mediated by dis-
The author’s work is supported by the Wellcome tinct populations of ipRGCs. Nature, 476, 92–95.
Trust and BBSRC. Chyb, S., Hevers, W., Forte, M., Wolfgang, W. J., Selinger, Z.,
& Hardie, R. C. (1999a). Modulation of the light response
by cAMP in Drosophila photoreceptors. The Journal of
References Neuroscience, 19, 8799–8807.
Chyb, S., Raghu, P., & Hardie, R. C. (1999b). Polyunsaturated
Adamski, F. M., Timms, K. M., & Shieh, B. H. (1999). fatty acids activate the Drosophila light-sensitive channels
A unique isoform of phospholipase Cbeta4 highly expressed TRP and TRPL. Nature, 397, 255–259.
in the cerebellum and eye. Biochimica et Biophysica Acta, Contin, M. A., Verra, D. M., & Guido, M. E. (2006). An
1444, 55–60. invertebrate-like phototransduction cascade mediates light
Amatruda, T. T., 3rd, Steele, D. A., Slepak, V. Z., & detection in the chicken retinal ganglion cells. The FASEB
Simon, M. I. (1991). G alpha 16, a G protein alpha subunit Journal, 20, 2648–2650.
specifically expressed in hematopoietic cells. Proceedings Dacey, D. M., Liao, H. W., Peterson, B. B., Robinson, F. R.,
of the National Academy of Sciences of the United States of Smith, V. C., Pokorny, J., et al. (2005). Melanopsin-
America, 88, 5587–5591. expressing ganglion cells in primate retina signal colour
Baver, S. B., Pickard, G. E., & Sollars, P. J. (2008). Two types and irradiance and project to the LGN. Nature, 433,
of melanopsin retinal ganglion cell differentially innervate 749–754.
the hypothalamic suprachiasmatic nucleus and the olivary Davies, W. I., Zheng, L., Hughes, S., Tamai, T. K., Turton, M.,
pretectal nucleus. The European Journal of Neuroscience, Halford, S., et al. (2011). Functional diversity of
27, 1763–1770. melanopsins and their global expression in the teleost retina.
Belenky, M. A., Smeraski, C. A., Provencio, I., Sollars, P. J., & Cellular and Molecular Life Sciences, 68(24), 4115–4132.
Pickard, G. E. (2003). Melanopsin retinal ganglion cells Davignon, I., Barnard, M., Gavrilova, O., Sweet, K., &
receive bipolar and amacrine cell synapses. The Journal of Wilkie, T. M. (1996). Gene structure of murine Gna11 and
Comparative Neurology, 460, 380–393. Gna15: Tandemly duplicated Gq class G protein alpha sub-
Bellingham, J., Chaurasia, S. S., Melyan, Z., Liu, C., unit genes. Genomics, 31, 359–366.
Cameron, M. A., Tarttelin, E. E., et al. (2006). Evolution Do, M. T., Kang, S. H., Xue, T., Zhong, H., Liao, H. W.,
of melanopsin photoreceptors: Discovery and characteriza- Bergles, D. E., et al. (2009). Photon capture and signalling
tion of a new melanopsin in nonmammalian vertebrates. by melanopsin retinal ganglion cells. Nature, 457, 281–287.
PLoS Biology, 4, e254. Do, M. T., & Yau, K. W. (2010). Intrinsically photosensitive
Berson, D. M. (2007). Phototransduction in ganglion-cell pho- retinal ganglion cells. Physiological Reviews, 90, 1547–1581.
toreceptors. Pflügers Archiv, 454, 849–855. Drivenes, O., Soviknes, A. M., Ebbesson, L. O., Fjose, A.,
Berson, D. M., Castrucci, A. M., & Provencio, I. (2010). Mor- Seo, H. C., & Helvik, J. V. (2003). Isolation and characteri-
phology and mosaics of melanopsin-expressing retinal gan- zation of two teleost melanopsin genes and their differential
glion cell types in mice. The Journal of Comparative expression within the inner retina and brain. The Journal of
Neurology, 518, 2405–2422. Comparative Neurology, 456, 84–93.
Berson, D. M., Dunn, F. A., & Takao, M. (2002). Photo- Dumitrescu, O. N., Pucci, F. G., Wong, K. Y., & Berson, D. M.
transduction by retinal ganglion cells that set the circadian (2009). Ectopic retinal ON bipolar cell synapses in the OFF
clock. Science, 295, 1070–1073. inner plexiform layer: Contacts with dopaminergic amacrine
37

cells and melanopsin ganglion cells. The Journal of Compar- Hankins, M. W., & Lucas, R. J. (2002). The primary visual
ative Neurology, 517, 226–244. pathway in humans is regulated according to long-term light
Dupre, D. J., Robitaille, M., Rebois, R. V., & Hebert, T. E. exposure through the action of a nonclassical photopigment.
(2009). The role of Gbetagamma subunits in the organiza- Current Biology, 12, 191–198.
tion, assembly, and function of GPCR signaling complexes. Hankins, M. W., Peirson, S. N., & Foster, R. G. (2008).
Annual Review of Pharmacology and Toxicology, 49, Melanopsin: An exciting photopigment. Trends in
31–56. Neurosciences, 31, 27–36.
Ecker, J. L., Dumitrescu, O. N., Wong, K. Y., Alam, N. M., Hannibal, J., & Fahrenkrug, J. (2004). Target areas innervated
Chen, S. K., Legates, T., et al. (2010). Melanopsin- by PACAP-immunoreactive retinal ganglion cells. Cell and
expressing retinal ganglion-cell photoreceptors: Cellular Tissue Research, 316, 99–113.
diversity and role in pattern vision. Neuron, 67, 49–60. Hardie, R. C. (2001). Phototransduction in Drosophila
Elia, N., Frechter, S., Gedi, Y., Minke, B., & Selinger, Z. melanogaster. The Journal of Experimental Biology, 204,
(2005). Excess of Gbetae over Gqalphae in vivo prevents 3403–3409.
dark, spontaneous activity of Drosophila photoreceptors. Hardie, R. C. (2003). TRP channels in Drosophila photo-
The Journal of Cell Biology, 171, 517–526. receptors: The lipid connection. Cell Calcium, 33, 385–393.
Fahrenkrug, J., Falktoft, B., Georg, B., & Rask, L. (2009). Hardie, R. C. (2007). TRP channels and lipids: From Drosoph-
N-linked deglycosylated melanopsin retains its responsiveness ila to mammalian physiology. The Journal of Physiology,
to light. Biochemistry, 48, 5142–5148. 578, 9–24.
Ferreira, P. A., & Pak, W. L. (1994). Bovine phospholipase Hardie, R., & Raghu, P. (2001). Visual transduction in Dro-
C highly homologous to the norpA protein of Drosophila sophila. Nature, 413, 186–193.
is expressed specifically in cones. The Journal of Biological Hartwick, A. T., Bramley, J. R., Yu, J., Stevens, K. T.,
Chemistry, 269, 3129–3131. Allen, C. N., Baldridge, W. H., et al. (2007). Light-evoked
Ferreira, P. A., Shortridge, R. D., & Pak, W. L. (1993). Dis- calcium responses of isolated melanopsin-expressing retinal
tinctive subtypes of bovine phospholipase C that have pref- ganglion cells. The Journal of Neuroscience, 27,
erential expression in the retina and high homology to the 13468–13480.
norpA gene product of Drosophila. Proceedings of the Hatori, M., Le, H., Vollmers, C., Keding, S. R., Tanaka, N.,
National Academy of Sciences of the United States of Buch, T., et al. (2008). Inducible ablation of melanopsin-
America, 90, 6042–6046. expressing retinal ganglion cells reveals their central
Freedman, M. S., Lucas, R. J., Soni, B., von Schantz, M., role in non-image forming visual responses. PLoS One, 3,
Munoz, M., David-Gray, Z., et al. (1999). Regulation of e2451.
mammalian circadian behavior by non-rod, non-cone, ocular Hattar, S., Kumar, M., Park, A., Tong, P., Tung, J.,
photoreceptors. Science, 284, 502–504. Yau, K. W., et al. (2006). Central projections of
Gamlin, P. D., Mcdougal, D. H., Pokorny, J., Smith, V. C., melanopsin-expressing retinal ganglion cells in the mouse.
Yau, K. W., & Dacey, D. M. (2007). Human and macaque The Journal of Comparative Neurology, 497, 326–349.
pupil responses driven by melanopsin-containing retinal Hattar, S., Liao, H. W., Takao, M., Berson, D. M., &
ganglion cells. Vision Research, 47, 946–954. Yau, K. W. (2002). Melanopsin-containing retinal ganglion
Giesbers, M. E., Shirzad-Wasei, N., Bosman, G. J., & De cells: Architecture, projections, and intrinsic photosensitiv-
Grip, W. J. (2008). Functional expression, targeting and ity. Science, 295, 1065–1070.
Ca2 þ signaling of a mouse melanopsin-eYFP fusion protein Hattar, S., Lucas, R. J., Mrosovsky, N., Thompson, S.,
in a retinal pigment epithelium cell line. Photochemistry and Douglas, R. H., Hankins, M. W., et al. (2003). Melanopsin
Photobiology, 84, 990–995. and rod-cone photoreceptive systems account for all major
Gooley, J. J., Lu, J., Fischer, D., & Saper, C. B. (2003). accessory visual functions in mice. Nature, 424, 75–81.
A broad role for melanopsin in nonvisual photoreception. Hayashida, Y., Rodriguez, C. V., Ogata, G., Partida, G. J.,
The Journal of Neuroscience, 23, 7093–7106. Oi, H., Stradleigh, T. W., et al. (2009). Inhibition of adult
Graham, D. M., Wong, K. Y., Shapiro, P., Frederick, C., rat retinal ganglion cells by D1-type dopamine receptor acti-
Pattabiraman, K., & Berson, D. M. (2008). Melanopsin gan- vation. The Journal of Neuroscience, 29, 15001–15016.
glion cells use a membrane-associated rhabdomeric photo- Hoshi, H., Liu, W. L., Massey, S. C., & Mills, S. L. (2009). ON
transduction cascade. Journal of Neurophysiology, 99, inputs to the OFF layer: Bipolar cells that break the stratifi-
2522–2532. cation rules of the retina. The Journal of Neuroscience, 29,
Guler, A. D., Ecker, J. L., Lall, G. S., Haq, S., Altimus, C. M., 8875–8883.
Liao, H. W., et al. (2008). Melanopsin cells are the principal Hubbard, K. B., & Hepler, J. R. (2006). Cell signalling diver-
conduits for rod-cone input to non-image-forming vision. sity of the Gqalpha family of heterotrimeric G proteins. Cel-
Nature, 453, 102–105. lular Signalling, 18, 135–150.
38

Isoldi, M. C., Rollag, M. D., Castrucci, A. M., & Provencio, I. light reflex at high irradiances in melanopsin-knockout mice.
(2005). Rhabdomeric phototransduction initiated by the Science, 299, 245–247.
vertebrate photopigment melanopsin. Proceedings of the Luo, D. G., Xue, T., & Yau, K. W. (2008). How vision begins:
National Academy of Sciences of the United States of An odyssey. Proceedings of the National Academy of
America, 102, 1217–1221. Sciences of the United States of America, 105, 9855–9862.
Johnson, J., Wu, V., Donovan, M., Majumdar, S., Lupi, D., Oster, H., Thompson, S., & Foster, R. G. (2008). The
Renteria, R. C., Porco, T., et al. (2010). Melanopsin- acute light-induction of sleep is mediated by OPN4-based
dependent light avoidance in neonatal mice. Proceedings photoreception. Nature Neuroscience, 11(9), 1068–1073.
of the National Academy of Sciences of the United States of Makino, C. L., Wen, X. H., & Lem, J. (2003). Piecing together
America, 107, 17374–17378. the timetable for visual transduction with transgenic
Kitamura, E., Danciger, M., Yamashita, C., Rao, N. P., animals. Current Opinion in Neurobiology, 13, 404–412.
Nusinowitz, S., Chang, B., et al. (2006). Disruption of the Margolis, D. J., Gartland, A. J., Euler, T., & Detwiler, P. B.
gene encoding the beta1-subunit of transducin in the Rd4/þ (2010). Dendritic calcium signaling in ON and OFF mouse
mouse. Investigative Ophthalmology & Visual Science, 47, retinal ganglion cells. The Journal of Neuroscience, 30,
1293–1301. 7127–7138.
Kostenis, E., Waelbroeck, M., & Milligan, G. (2005). Tec- Mcneill, D. S., Sheely, C. J., Ecker, J. L., Badea, T. C.,
hniques: Promiscuous Galpha proteins in basic research Morhardt, D., Guido, W., et al. (2011). Development of
and drug discovery. Trends in Pharmacological Sciences, melanopsin-based irradiance detecting circuitry. Neural
26, 595–602. Development, 6, 8.
Koyanagi, M., Kubokawa, K., Tsukamoto, H., Shichida, Y., & Melyan, Z., Tarttelin, E. E., Bellingham, J., Lucas, R. J., &
Terakita, A. (2005). Cephalochordate melanopsin: Evolu- Hankins, M. W. (2005). Addition of human melanopsin
tionary linkage between invertebrate visual cells and verte- renders mammalian cells photoresponsive. Nature, 433,
brate photosensitive retinal ganglion cells. Current Biology, 741–745.
15, 1065–1069. Mizuno, N., & Itoh, H. (2009). Functions and regulatory
Koyanagi, M., & Terakita, A. (2008). Gq-coupled rhodopsin mechanisms of Gq-signaling pathways. Neuro-Signals, 17,
subfamily composed of invertebrate visual pigment and 42–54.
melanopsin. Photochemistry and Photobiology, 84, Mrosovsky, N., Lucas, R., & Foster, R. (2001). Persistence of
1024–1030. masking responses to light in mice lacking rods and cones.
Krispel, C. M., Chen, C. K., Simon, M. I., & Burns, M. E. Journal of Biological Rhythms, 16, 585–587.
(2003). Prolonged photoresponses and defective adaptation Mure, L. S., Rieux, C., Hattar, S., & Cooper, H. M. (2007).
in rods of Gbeta5/ mice. The Journal of Neuroscience, Melanopsin-dependent nonvisual responses: Evidence for
23, 6965–6971. photopigment bistability in vivo. Journal of Biological
Kumbalasiri, T., Rollag, M. D., Isoldi, M. C., Castrucci, A. M., Rhythms, 22, 411–424.
& Provencio, I. (2007). Melanopsin triggers the release of Nakamura, K., Nukada, T., Imai, K., & Sugiyama, H. (1996).
internal calcium stores in response to light. Photochemistry Importance of N-terminal regions of G protein alpha
and Photobiology, 83, 273–279. subunits for the activation of phospholipase C in Xenopus
Lamb, T. D. (1996). Ida Mann Lecture. Transduction in oocytes. Journal of Biochemistry, 120, 996–1001.
human photoreceptors. Australian and New Zealand Journal Nasi, E., & Del Pilar Gomez, M. (2009). Melanopsin-mediated
of Ophthalmology, 24, 105–110. light-sensing in amphioxus: A glimpse of the microvillar
Lee, R. H., Lieberman, B. S., Yamane, H. K., Bok, D., & photoreceptor lineage within the deuterostomia. Communi-
Fung, B. K. (1992). A third form of the G protein beta sub- cative & Integrative Biology, 2, 441–443.
unit. 1. Immunochemical identification and localization to Newman, L. A., Walker, M. T., Brown, R. L., Cronin, T. W., &
cone photoreceptors. The Journal of Biological Chemistry, Robinson, P. R. (2003). Melanopsin forms a functional short-
267, 24776–24781. wavelength photopigment. Biochemistry, 42, 12734–12738.
Lucas, R. J., Douglas, R. H., & Foster, R. G. (2001). Character- Ostergaard, J., Hannibal, J., & Fahrenkrug, J. (2007). Synaptic
ization of an ocular photopigment capable of driving pupil- contact between melanopsin-containing retinal ganglion
lary constriction in mice. Nature Neuroscience, 4, 621–626. cells and rod bipolar cells. Investigative Ophthalmology &
Lucas, R. J., & Foster, R. G. (1999). Neither functional rod Visual Science, 48, 3812–3820.
photoreceptors nor rod or cone outer segments are required Panda, S., Nayak, S. K., Campo, B., Walker, J. R.,
for the photic inhibition of pineal melatonin. Endocrinology, Hogenesch, J. B., & Jegla, T. (2005). Illumination of the
140, 1520–1524. melanopsin signaling pathway. Science, 307, 600–604.
Lucas, R. J., Hattar, S., Takao, M., Berson, D. M., Panda, S., Provencio, I., Tu, D. C., Pires, S. S., Rollag, M. D.,
Foster, R. G., & Yau, K. W. (2003). Diminished pupillary Castrucci, A. M., et al. (2003). Melanopsin is required for
39

non-image-forming photic responses in blind mice. Science, Schaefer, M. (2005). Homo- and heteromeric assembly of TRP
301, 525–527. channel subunits. Pflügers Archiv, 451, 35–42.
Panda, S., Sato, T. K., Castrucci, A. M., Rollag, M. D., Schmidt, T. M., Chen, S. K., & Hattar, S. (2011). Intrinsically
Degrip, W. J., Hogenesch, J. B., et al. (2002). Melanopsin photosensitive retinal ganglion cells: Many subtypes, diverse
(Opn4) requirement for normal light-induced circadian functions. Trends in Neurosciences, 31(11), 572–580.
phase shifting. Science, 298, 2213–2216. Schmidt, T. M., & Kofuji, P. (2009). Functional and morpho-
Peirson, S. N., Oster, H., Jones, S. L., Leitges, M., logical differences among intrinsically photosensitive retinal
Hankins, M. W., & Foster, R. G. (2007). Microarray analysis ganglion cells. The Journal of Neuroscience, 29, 476–482.
and functional genomics identify novel components of Schmidt, T. M., & Kofuji, P. (2010). Differential cone pathway
melanopsin signaling. Current Biology, 17, 1363–1372. influence on intrinsically photosensitive retinal ganglion cell
Peng, Y. W., Robishaw, J. D., Levine, M. A., & Yau, K. W. subtypes. The Journal of neuroscience: the official journal of
(1992). Retinal rods and cones have distinct G protein beta the Society for Neuroscience, 30, 16262–16271.
and gamma subunits. Proceedings of the National Academy Schmidt, T. M., & Kofuji, P. (2011). Structure and function of
of Sciences of the United States of America, 89, 10882–10886. bistratified intrinsically photosensitive retinal ganglion cells
Perez-Leighton, C. E., Schmidt, T. M., Abramowitz, J., in the mouse. The Journal of Comparative Neurology, 519,
Birnbaumer, L., & Kofuji, P. (2011). Intrinsic photo- 1492–1504.
transduction persists in melanopsin-expressing ganglion cells Schmidt, T. M., Taniguchi, K., & Kofuji, P. (2008). Intrinsic
lacking diacylglycerol-sensitive TRPC subunits. The Euro- and extrinsic light responses in melanopsin-expressing gan-
pean Journal of Neuroscience, 33, 856–867. glion cells during mouse development. Journal of Neuro-
Pires, S. S., Hughes, S., Turton, M., Melyan, Z., Peirson, S. N., physiology, 100, 371–384.
Zheng, L., et al. (2009). Differential expression of two dis- Sekaran, S., Foster, R. G., Lucas, R. J., & Hankins, M. W.
tinct functional isoforms of melanopsin (Opn4) in the mam- (2003). Calcium imaging reveals a network of intrinsically
malian retina. The Journal of Neuroscience, 29, light-sensitive inner-retinal neurons. Current Biology, 13,
12332–12342. 1290–1298.
Provencio, I., Jiang, G., De Grip, W. J., Hayes, W. P., & Sekaran, S., Lall, G. S., Ralphs, K. L., Wolstenholme, A. J.,
Rollag, M. D. (1998). Melanopsin: An opsin in melanophores, Lucas, R. J., Foster, R. G., et al. (2007). 2-
brain, and eye. Proceedings of the National Academy of Aminoethoxydiphenylborane is an acute inhibitor of
Sciences of the United States of America, 95, 340–345. directly photosensitive retinal ganglion cell activity in vitro
Provencio, I., Rodriguez, I. R., Jiang, G., Hayes, W. P., and in vivo. The Journal of Neuroscience, 27, 3981–3986.
Moreira, E. F., & Rollag, M. D. (2000). A novel human Sekaran, S., Lupi, D., Jones, S. L., Sheely, C. J., Hattar, S.,
opsin in the inner retina. The Journal of Neuroscience, 20, Yau, K. W., et al. (2005). Melanopsin-dependent photore-
600–605. ception provides earliest light detection in the mammalian
Provencio, I., Rollag, M. D., & Castrucci, A. M. (2002). Photo- retina. Current Biology, 15, 1099–1107.
receptive net in the mammalian retina. This mesh of cells Semo, M., Gias, C., Ahmado, A., Sugano, E., Allen, A. E.,
may explain how some blind mice can still tell day from Lawrence, J. M., et al. (2010). Dissecting a role for
night. Nature, 415, 493. melanopsin in behavioural light aversion reveals a response
Qiu, X., Kumbalasiri, T., Carlson, S. M., Wong, K. Y., independent of conventional photoreception. PLoS One, 5,
Krishna, V., Provencio, I., et al. (2005). Induction of photo- e15009.
sensitivity by heterologous expression of melanopsin. Shindo, Y., Miura, H., Carninci, P., Kawai, J., Hayashizaki, Y.,
Nature, 433, 745–749. Ninomiya, Y., et al. (2008). G alpha14 is a candidate media-
Raghu, P. (2006). Regulation of Drosophila TRPC channels tor of sweet/umami signal transduction in the posterior
by protein and lipid interactions. Seminars in Cell & Devel- region of the mouse tongue. Biochemical and Biophysical
opmental Biology, 17, 646–653. Research Communications, 376, 504–508.
Rao, A., Dallman, R., Henderson, S., & Chen, C. K. (2007). Suh, B. C., & Hille, B. (2008). PIP2 is a necessary cofactor for
Gbeta5 is required for normal light responses and morphol- ion channel function: How and why? Annual Review of Bio-
ogy of retinal ON-bipolar cells. The Journal of Neuroscience, physics, 37, 175–195.
27, 14199–14204. Suh, B. C., Inoue, T., Meyer, T., & Hille, B. (2006). Rapid
Rhee, S. G. (2001). Regulation of phosphoinositide-specific chemically induced changes of PtdIns(4,5)P2 gate KCNQ
phospholipase C. Annual Review of Biochemistry, 70, ion channels. Science, 314, 1454–1457.
281–312. Thapan, K., Arendt, J., & Skene, D. J. (2001). An action spec-
Ruby, N. F., Brennan, T. J., Xie, X., Cao, V., Franken, P., trum for melatonin suppression: Evidence for a novel non-
Heller, H. C., et al. (2002). Role of melanopsin in circadian rod, non-cone photoreceptor system in humans. The Journal
responses to light. Science, 298, 2211–2213. of Physiology, 535, 261–267.
40

Tizzano, M., Dvoryanchikov, G., Barrows, J. K., Kim, S., Weschler, C. J., Wells, J. R., Poppendieck, D., Hubbard, H., &
Chaudhari, N., & Finger, T. E. (2008). Expression of Gal- Pearce, T. A. (2006). Workgroup report: Indoor chemistry
pha14 in sweet-transducing taste cells of the posterior and health. Environmental Health Perspectives, 114,
tongue. BMC Neuroscience, 9, 110. 442–446.
Torii, M., Kojima, D., Okano, T., Nakamura, A., Terakita, A., Wettschureck, N., & Offermanns, S. (2005). Mammalian G
Shichida, Y., et al. (2007). Two isoforms of chicken proteins and their cell type specific functions. Physiological
melanopsins show blue light sensitivity. FEBS Letters, 581, Reviews, 85, 1159–1204.
5327–5331. Wilkie, T. M., Gilbert, D. J., Olsen, A. S., Chen, X. N.,
Tu, D. C., Zhang, D., Demas, J., Slutsky, E. B., Provencio, I., Amatruda, T. T., Korenberg, J. R., et al. (1992). Evolution
Holy, T. E., et al. (2005). Physiologic diversity and develop- of the mammalian G protein alpha subunit multigene family.
ment of intrinsically photosensitive retinal ganglion cells. Nature Genetics, 1, 85–91.
Neuron, 48, 987–999. Wilkie, T. M., Scherle, P. A., Strathmann, M. P., Slepak, V. Z.,
Tummala, H., Ali, M., Getty, P., Hocking, P. M., Burt, D. W., & Simon, M. I. (1991). Characterization of G-protein alpha
Inglehearn, C. F., et al. (2006). Mutation in the guanine subunits in the Gq class: Expression in murine tissues and
nucleotide-binding protein beta-3 causes retinal degenera- in stromal and hematopoietic cell lines. Proceedings of the
tion and embryonic mortality in chickens. Investigative National Academy of Sciences of the United States of
Ophthalmology & Visual Science, 47, 4714–4718. America, 88, 10049–10053.
Viney, T. J., Balint, K., Hillier, D., Siegert, S., Boldogkoi, Z., Wong, K. Y., Dunn, F. A., & Berson, D. M. (2005). Photore-
Enquist, L. W., et al. (2007). Local retinal circuits of ceptor adaptation in intrinsically photosensitive retinal gan-
melanopsin-containing ganglion cells identified by trans- glion cells. Neuron, 48, 1001–1010.
synaptic viral tracing. Current Biology, 17, 981–988. Wong, K. Y., Dunn, F. A., Graham, D. M., & Berson, D. M.
Vugler, A. A., Redgrave, P., Semo, M., Lawrence, J., (2007). Synaptic influences on rat ganglion-cell photo-
Greenwood, J., & Coffey, P. J. (2007). Dopamine neurones receptors. The Journal of Physiology, 582, 279–296.
form a discrete plexus with melanopsin cells in normal and Xue, T., Do, M. T., Riccio, A., Jiang, Z., Hsieh, J.,
degenerating retina. Experimental Neurology, 205, 26–35. Wang, H. C., et al. (2011). Melanopsin signalling in mamma-
Walker, M. T., Brown, R. L., Cronin, T. W., & Robinson, P. R. lian iris and retina. Nature, 479, 67–73.
(2008). Photochemistry of retinal chromophore in mouse Yau, K. W., & Hardie, R. C. (2009). Phototransduction motifs
melanopsin. Proceedings of the National Academy of and variations. Cell, 139, 246–264.
Sciences of the United States of America, 105, 8861–8865. Yule, D. I., Baker, C. W., & Williams, J. A. (1999). Calcium
Warren, E. J., Allen, C. N., Brown, R. L., & Robinson, D. W. signaling in rat pancreatic acinar cells: A role for Galphaq,
(2003). Intrinsic light responses of retinal ganglion cells Galpha11, and Galpha14. The American Journal of Physiol-
projecting to the circadian system. The European Journal ogy, 276, G271–G279.
of Neuroscience, 17, 1727–1735. Zaidi, F. H., Hull, J. T., Peirson, S. N., Wulff, K., Aeschbach, D.,
Warren, E. J., Allen, C. N., Brown, R. L., & Robinson, D. W. Gooley, J. J., et al. (2007). Short-wavelength light sensitivity
(2006). The light-activated signaling pathway in SCN- of circadian, pupillary, and visual awareness in humans
projecting rat retinal ganglion cells. The European Journal lacking an outer retina. Current Biology, 17, 2122–2128.
of Neuroscience, 23, 2477–2487. Zhang, D. Q., Wong, K. Y., Sollars, P. J., Berson, D. M.,
Watson, A. J., Aragay, A. M., Slepak, V. Z., & Simon, M. I. Pickard, G. E., & Mcmahon, D. G. (2008). Intraretinal sig-
(1996). A novel form of the G protein beta subunit Gbeta5 naling by ganglion cell photoreceptors to dopaminergic
is specifically expressed in the vertebrate retina. The Journal amacrine neurons. Proceedings of the National Academy of
of Biological Chemistry, 271, 28154–28160. Sciences of the United States of America, 105, 14181–14186.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 3

Circadian clocks: Lessons from fish

M. Laura Idda{, Cristiano Bertolucci{, Daniela Vallone{, Yoav Gothilf},


Francisco Javier Sánchez-Vázquez} and Nicholas S. Foulkes{,*

{
Karlsruhe Institute of Technology, Institute of Toxicology and Genetics, Hermann-von-Helmholtz-Platz,
Eggenstein-Leopoldshafen, Germany
{
Department of Biology and Evolution, University of Ferrara, Ferrara, Italy
}
Department of Neurobiology, George S. Wise Faculty of Life Sciences, Tel Aviv University, Tel Aviv, Israel
}
Department of Physiology, Faculty of Biology, University of Murcia, Murcia, Spain

Abstract: Our understanding of the molecular and cellular organization of the circadian timing system
in vertebrates has increased enormously over the past decade. In large part, progress has been based on
genetic studies in the mouse as well as on fundamental similarities between vertebrate and Drosophila
clocks. The zebrafish was initially considered as a potentially attractive genetic model for identifying
vertebrate clock genes. However, instead, fish have ultimately proven to be valuable complementary
models for studying various aspects of clock biology. For example, many fish can shift from diurnal to
nocturnal activity implying specific flexibility in their clock function. We have learned much about the
function of light input pathways, and the ontogeny and function of the pineal organ, the fish central
pacemaker. Finally, blind cavefish have also provided new insight into the evolution of the circadian
clock under extreme environmental conditions.

Keywords: zebrafish; cavefish; genetics; clock mutants; peripheral clocks; pineal gland; clock ontogeny;
cell lines; blind clocks.

Introduction progress has been based on detailed studies of


the clock mechanism in a range of model
Our understanding of the circadian clock, how organisms amenable to detailed genetic analysis,
it is regulated by zeitgebers, and also how it in most notably Drosophila, cyanobacteria, Neuros-
turn regulates many physiological processes has pora, Arabidopsis, and the mouse. Further, in
increased enormously over the past years. This the case of the vertebrate clock, the surprising
discovery of functional circadian clocks in com-
*Corresponding author. monly used cell lines enabled detailed biochemi-
Tel.: þ49-721-60823394; Fax: þ49-721-60823354 cal analysis that has taught us much about how
E-mail: nicholas.foulkes@kit.edu core clock mechanisms work. However, a focus

http://dx.doi.org/10.1016/B978-0-444-59427-3.00003-4 41
42

on a limited number of model species inevitably tools for generating stable and transient trans-
constrains the scope of biological problems that genic lines, has made the zebrafish a favorite
can be tackled. With the advent of powerful tools model for in vivo imaging of gene expression dur-
for molecular analysis (such as complete genome ing early embryogenesis (Nüsslein-Volhard and
or transcriptome sequencing) that can theoretically Dahm, 2002). Another major advantage of the
be applied to any species, it is now easier to zebrafish is its amenability to large-scale forward
explore aspects of clock biology, not necessarily genetic screens (Haffter et al., 1996; Mullins
addressable in the classic genetic model organisms. et al., 1994). A single female zebrafish can nor-
Teleosts represent one of the most successful mally produce hundreds of offspring each week.
groups of vertebrates, showing diverse anatomical Further, robust methods now exist for generating
and physiological adaptations to a wide range of large panels of mutagenized zebrafish. By screen-
habitats. For this reason, studying these animals ing these panels, the consequent identification
provides a fascinating opportunity to explore the of specific mutants, and then the mapping of
plasticity of circadian clocks and how they adapt mutated genetic loci, we have learned much
to various environmental conditions during evolu- about basic developmental processes (Nüsslein-
tion. For example, in contrast to mammals, the Volhard and Dahm, 2002). The past few years
circadian system of fish shows impressive flexibil- have seen the use of zebrafish expanding from
ity, as the same fish species can exhibit diurnal its traditional user base and being applied to
or nocturnal behavior and shift from one type of study diverse aspects of biology including behav-
phasing behavior to another depending on the sea- ior and various aspects of physiology (Ingham,
son or during ontogeny. 2009; Petzold et al., 2009). Notably, the impres-
Detailed information on the molecular and sive capacity of most zebrafish adult tissues to
cellular basis of the clock in fish has stemmed regenerate following injury has attracted consid-
primarily from studies of the genetic model spe- erable attention from research aimed at under-
cies, the zebrafish. In this review, we will over- standing and treating certain human diseases
view our current understanding of the regulation such as heart and neurodegenerative diseases
and mechanism of the zebrafish circadian clock and cancer (Brittijn et al., 2009; Ingham, 2009).
and then describe how more recent work compar- Further, the completion of the zebrafish genome
ing the zebrafish clock with that of blind cavefish sequence has greatly facilitated the process of
provides a powerful route to exploring clock mapping and identification of mutated loci and is
evolution. now enabling accurate analysis of transcriptomes
by next generation sequencing technology. Thus,
Zebrafish: A genetic model species today the zebrafish has become an important com-
plementary model species alongside the mouse for
A model for studying embryonic development studying the molecular genetics of many aspects of
vertebrate biology and pathology.
The zebrafish was originally developed as model
for studying the genetics of embryology and
developmental biology in vertebrates (Haffter Chronobiology and the zebrafish
et al., 1996; Mullins et al., 1994). Several features
made it particularly attractive for exploring early The attention of chronobiologists originally
developmental processes. Firstly, it has transpar- turned to zebrafish many years ago (Cahill,
ent embryos that develop rapidly in a transparent 1996, 2002) during the “dark ages” of our knowl-
egg–shell outside of the mother. This basic prop- edge of the molecular workings of the vertebrate
erty, combined with the establishment of many clock. No vertebrate circadian clock genes had
43

been cloned and so the molecular mechanism circadian clock drives rhythmic synthesis of the
of the circadian clock in vertebrates was a com- hormone melatonin. Melatonin levels are high at
plete mystery. At a stage, when large-scale night and low during the day as a result of
genetic analysis of mice for complex behavioral regulated transcription and stability of serotonin-
phenotypes was prohibitive in terms of cost and N-acetyl-transferase (AANAT), a key enzyme
infrastructure, zebrafish offered the attractive in melatonin synthesis. In common with other
possibility to perform a low cost screen for clock teleosts, zebrafish possess two aanat genes:
mutants and thereby a route to identify verte- aanat1, which is expressed predominantly in
brate clock genes. Further, zebrafish also showed the retina, and aanat2, which is expressed in the
great potential to trace the origin of the circadian pineal and at lower levels in the retina (Falcon
clock during embryogenesis. In an unexpected et al., 2003; Gothilf et al., 1999). Expression of
turn, following the identification of the first clock this enzyme is regulated by the circadian clock
gene, clock, from the mouse and then the cloning and its activity is rapidly suppressed in response
and characterization of the zebrafish homolog to illumination during the night (Appelbaum
of this gene, it soon became apparent that et al., 2006; Ziv et al., 2007). In addition to the
the organization of the zebrafish circadian clock melatonin-producing photoreceptors, the pineal
mechanism and its regulation by light was fun- contains neurons (Masai et al., 1997) that inner-
damentally different from mammals. Specifically, vate a variety of brain regions, a fact that has
zebrafish peripheral clocks are directly entrainable received little attention until recently (Yáñez
by light (Whitmore et al., 2000), actually closely et al., 2009). Thus, the pineal gland transduces
resembling the situation in Drosophila (Plautz environmental light information into a neural
et al., 1997). This direct light-sensing property of and a neuroendocrine signal. Many studies now
peripheral clocks was also encountered in cell lines focus on identifying the control mechanisms
derived from zebrafish embryos such as Pac-2 directing the first appearance of rhythmic melato-
cells (Whitmore et al., 2000). This observation nin synthesis during development, its regulation
has raised many basic questions, most notably by the clock and light as well as pineal-specific
which molecules serve as the photoreceptors in patterns of gene expression.
peripheral tissues in fish? Promoter studies using a transient transgenic
Another element of the circadian timing system reporter assay as well as transfection of the Pac-2
in zebrafish that has received considerable att- cell line have identified a cluster of enhancer
ention is the pineal organ. In nonmammalian elements lying downstream of the aanat2 gene
vertebrates, including zebrafish, the pineal con- that is termed the pineal-restrictive downstream
tains all the elements required for photic entrain- module (Appelbaum et al., 2004). This includes
ment and circadian rhythm generation: it is three photoreceptor conserved elements (PCEs)
photoreceptive and contains an intrinsic circadian and an E box. The E box confers clock regulation
oscillator (Falcon et al., 2010; Korf et al., 1998). In by interacting with Clock:Bmal heterodimers (see
this regard, the pineal resembles peripheral tissue next section), while the PCEs are a target for the
clocks in zebrafish. However, importantly, the homeobox gene, otx5, and thereby determine
pineal organ is considered to serve as the central pineal-specific expression (Appelbaum et al.,
pacemaker in fish. Further, fish pineal cells are 2006). These factors appear to interact synergisti-
specialized photoreceptor cells sharing various cally (Appelbaum et al., 2005) demonstrating how
structural and functional properties with retinal the clock mechanism may interact with
photoreceptors. Indeed, pineal and retinal photo- other developmental and cellular processes. More
receptor cells express a similar set of genes, or, recently, bioinformatic analysis of the promoters
in certain cases, paralogs. In the fish pineal, the of pineal-specific genes that were identified
44

using DNA microarrays indicates that additional Kazimi and Cahill, 1999). A flow-through culture
mechanisms and enhancer sequences regulate system was successfully used to accurately mea-
the expression of pineal-specific genes (Alon sure circadian rhythms of melatonin release from
et al., 2009). individual adult zebrafish pineal organs (Cahill,
1996, 1997). Although not suitable for large-scale
analysis, this approach was used to validate
Zebrafish and the vertebrate core clock clock mutant phenotypes identified in locomotor
mechanism activity screens.
Finally, as in the case of many other genetic
Searching for new clock genes using zebrafish models, the development of transgenic biolumi-
nescent reporter lines has provided a powerful
The very first studies of circadian biology in the approach to screen for clock mutants. Specifically,
zebrafish aimed at large-scale screens to identify transgenic zebrafish lines were established where
clock mutants. Initial attention focused on moni- expression of a luciferase (luc) reporter gene is
toring behavioral rhythms (locomotor activity) driven by the clock-regulated per3 gene promoter
that had been so successfully exploited as a circa- (Kaneko and Cahill, 2005). Subsequently, when
dian clock assay in Drosophila and rodents. How- luciferin is added to the fish water of these trans-
ever, it soon became apparent that for isolated genics, a sufficient quantity diffuses into the
adult fish, these rhythms were far from robust tissues and cells to permit the oxidation of lucif-
and there was significant variability even among erin by the luc enzyme and so the generation of
populations of wild-type fish (Cahill, 2002; bioluminescence that can be measured in a scintil-
Hurd et al., 1998). An alternative, far more reli- lation counter- or camera-based assay. Impor-
able strategy was to monitor the movement of tantly, bioluminescence can be measured from
zebrafish larvae in 24-well plates using an larvae over a period of several days in small
automated video analysis system (Cahill et al., volumes of medium without requiring the addi-
1998; Hurd et al., 1998). This assay was indeed tion of extra luciferin (Kaneko and Cahill, 2005).
successfully used in a large-scale screen for clock The only potential complication of this assay is
mutant zebrafish (DeBruyne et al., 2004). More that while locomotor activity or melatonin pro-
recently, studies of locomotor activity in zebrafish duction specifically reflects central clock activity,
have also taught us about key features of the luc expression in a transgenic line will predomi-
organization of the circadian timing system. Thus, nantly assay peripheral clock function.
independent phasing between locomotor (mostly Having identified a clock mutant phenotype, the
diurnal) and feeding (nocturnal) rhythms was next stage is to pinpoint the location of the mutated
found in zebrafish, providing evidence for the genetic locus. In zebrafish, these subsequent
flexibility of their circadian system and suggesting “mapping” steps require crossing of the mutant
it possesses an underlying multioscillatory control into another genetic background displaying signifi-
(del Pozo et al., 2011). cant polymorphism relative to the strain where the
An alternative clock output that was exten- original mutagenesis was performed. Subsequent
sively tested as a screening assay is the rhythmic segregation of the mutant phenotype with poly-
release of melatonin either from individual morphic, microsatellite markers is used as a way
embryos or from explanted adult pineal organs. to map the mutation to a particular subregion or
In developing embryos raised under a light–dark locus (Nüsslein-Volhard and Dahm, 2002). This
(LD) cycle, melatonin production is first detected approach is time consuming but relatively straight-
around 24 h postfertilization (hpf), coinciding forward in the case of easily scorable phenotypes,
with the expression of aanat (Gothilf et al., 1999; for example, anatomical abnormalities in the early
45

embryo (Mullins et al., 1994). However, for long- elements (Period (Per1, 2, and 3) and
term behavioral assays and relatively subtle Cryptochrome (Cry1 and 2)), which in turn feed-
mutant phenotypes that require many repeats to back to downregulate their own expression and
achieve statistical confidence as for the identifica- so allow the start of a new cycle of the feedback
tion of circadian clock mutants, the whole mapping loop. The bHLH PAS domain transcription
process can be extremely laborious and may take activators Clock and Bmal bind as heterodimers
several years. to E box elements, located in the promoters of
To date, one major forward genetic screen has the per and cry genes and induce their transcrip-
been performed where the equivalent of 6500 tion. After translation, dimerization, and trans-
mutagenized zebrafish genomes were screened location to the nucleus, the Per and Cry proteins
for dominant mutations affecting circadian physically interact with and thereby inhibit the
rhythms of locomotor activity (DeBruyne et al., transcriptional activation driven by the Clock:
2004). Eight homozygous viable, semidominant Bmal complex. Further, an additional feedback
mutants were subsequently identified. In one loop directs the rhythmic expression of the bmal
mutant that was identified by virtue of a transcript (Emery and Reppert, 2004). The pres-
shortening of the free-running circadian period, ence of this additional loop together with the
an isoleucine to asparagine mutation was tightly regulated stability, turn over, and subcellu-
identified in the PAS domain of the clock gene lar localization of core clock components all con-
clock1a (see next section; Tan et al., 2008). tribute to conferring accuracy and robustness on
Another mutant termed lager and lime (lag(dg2)) the clock mechanism.
was also identified by a shortening of the circadian Thus, attention on the clock system in zebrafish
free-running period and by a reduction in tem- shifted from identifying elements of the core clock
perature compensation of the clock rhythm machinery, to the characterization of zebrafish
(DeBruyne et al., 2004). Although the mutated homologs of known clock genes. A genome dupli-
gene has yet to be characterized, it does not affect cation event that occurred during the evolution of
any of the known clock genes (DeBruyne et al., the teleost lineage has lead to a situation where
2004). Thus, forward genetic analysis of the there are extra clock gene copies in zebrafish com-
zebrafish clock, although laborious, may yet prove pared with the mouse (Postlethwait et al., 1998).
valuable for the identification of accessory In some cases, duplicated gene copies have subse-
elements of the circadian clock mechanism. quently been lost during evolution; however, in
many cases, the extra copies persist. This may
reflect redundancy with functions being shared by
Multiple clock genes in fish the extra genes. Alternatively, the duplicated genes
diverge in function and then multiple roles per-
Ultimately, following the cloning of the gene clock formed by the original gene may be distributed
in the mouse by Takahashi and colleagues (King among the “more specialized” copies. Thus, in the
et al., 1997), rapid progress was made in the eluci- zebrafish clock, six cry genes have been identified:
dation of the core clock mechanism in vertebrates cry1a, 1b, 2a, 2b, 3, and 4 (Kobayashi et al., 2000).
without input from the zebrafish. It emerged that Cry1a, 1b, 2a, and 2b share most sequence homol-
at the core of the mammalian circadian clock is a ogy with mammalian mcry1 and also repress
transcription–translation feedback loop that cycles Clock:Bmal directed transcriptional activation.
with a period of approximately 24 h (Albrecht, Interestingly, Cry3 and Cry4 do not repress Clock
2004; Reppert and Weaver, 2002). This regulatory and Bmal activation. While Cry3 does share
loop consists of positive elements (Clock and sequence similarity with the mammalian Crys,
Bmal) that drive the expression of negative Cry4 more closely resembles Drosophila Cry.
46

Given the role of Cry as a blue light photoreceptor spatial expression patterns (Wang, 2008a). Further,
in Drosophila, this has led to speculation that while per2 is a light-driven gene, the remaining per
zebrafish Cry4 might also function as a photorecep- genes are predominantly clock regulated. Impor-
tor (see later section). The temporal expression tantly, Per2 has been shown to function in combina-
patterns of the various cry genes vary significantly tion with Cry1a as an element of the light input
(Kobayashi et al., 2000). Thus, cry3 peaks during pathway as well as playing an important role in
the morning, cry1a, 1b, and 4 show peaks of expres- the early development of the pineal clock during
sion during the middle of the light phase, while embryogenesis (Hirayama et al., 2003; Tamai
cry2a and 2b rhythms peak during the evening. et al., 2007; Ziv and Gothilf, 2006; Ziv et al.,
Cry1a shows a light-driven pattern of expression, 2005). Thus, as in the case of the crys, combined
while the remaining cry genes are predominantly evidence seems to point to specialization of func-
clock regulated. Indeed, cry1a appears to represent tion for the various period genes.
a key element of the mechanism underlying photic
entrainment and the maintenance of high ampli-
tude cycling (Tamai et al., 2007). These functions Starting the clock during development
result from the ability of Cry1a to interact with
key regions of the Clock and Bmal activators, thus One fundamental issue addressed using the
preventing their heterodimerization and hindering zebrafish concerns when and how the clock is
their ability to transactivate from E box enhancer established during development. Clock gene tran-
elements (Tamai et al., 2007). With regard to the scripts are among the pool of maternally inherited
positive elements of the clock mechanism, three transcripts that are typically required for the very
clock genes (clock 1a, 1b, and 2 formerly identified first stages of embryonic development (Delaunay
as clock1, 2, and 3, respectively) and three bmal et al., 2003); however, the functional significance
genes (bmal1a, 1b, and 2, formerly identified as of these clock transcripts remains unclear. The
bmal 1, 3, and 2, respectively) have been identified earliest detectable circadian rhythms in the devel-
and have been shown to interact pairwise in various oping embryo are rhythms of clock gene expres-
heterodimeric combinations (Hirayama et al., 2003; sion detectable in the entire embryo as well as
Ishikawa et al., 2002; Wang, 2008b, 2009). These rhythmic melatonin synthesis and aanat2 expres-
heterodimers show different transactivation pro- sion in the pineal gland (Dekens and Whitmore,
perties and are differentially inhibited by Crys such 2008; Gothilf et al., 1999; Kazimi and Cahill,
as Cry1a. The expression pattern of the zebrafish 1999). It is also conceivable that clock genes
clock, and bmal family members also provides may perform non-clock-related functions at these
indications of significant differences in the organi- early developmental stages.
zation of the core transcription–translation feed- The establishment of circadian rhythms during
back loop in zebrafish compared with mouse early development clearly requires exposure to
(Whitmore et al., 1998). In most tissues, each of environmental zeitgebers such as light and tem-
these genes shows a robust mRNA expression perature changes. No circadian rhythms are
rhythm with a peak just after the LD transition observed at the whole animal level in embryos
(Cermakian et al., 2000). This contrasts with the sit- raised under constant temperature and constant
uation in mammals where bmal, but not clock, darkness (DD). Theoretically, this lack of rhyth-
shows rhythmic mRNA expression (Shearman micity could be explained by the presence of nor-
et al., 1999). Finally, extra per genes have also been mal clocks in each cell, which are asynchronous
encountered in fish. Two per1 homologs (per1a and with respect to each other. Alternatively, this
1b) together with single per2 and per3 genes have could result from a lack of clock function even
been cloned and characterized that exhibit different at the single-cell level. One study has addressed
47

this issue directly by investigating the regulation of RNA extracts first emerges around day 4/5 coinci-
per1b expression during early embryogenesis dent with the appearance of other key clock out-
(Dekens and Whitmore, 2008). In DD, oscillations puts such as locomotor activity rhythms (Cahill
in per1b expression in single cells were documented et al., 1998; Dekens and Whitmore, 2008; Dekens
that were asynchronous in the embryo, indicating et al., 2003). Thus, there would appear to be funda-
that even in the absence of LD cycles, circadian mental changes in the regulatory mechanisms
clocks do start ticking. underlying the core clock mechanism during
Light pulses delivered as early as the blastula development. Recent research in fish species with
stage up to early segmentation stages (4–16 hpf), a longer incubation period further supports the
before any classical photoreceptor organ appears, concept that clock systems in fish are present, light
are sufficient to set the phase of the central oscilla- responsive and fully functional during embryonic
tor in the pineal gland 2–3 days later (Ziv and development; thus continuous illumination
Gothilf, 2006). Thus, clearly light can be detected abolishes rhythmic expression of clock genes in
by nonspecialized photoreceptor cells and then these species (Davie et al., 2011).
this photic information must be preserved through
the subsequent rapid cell proliferation and differ-
entiation. Interestingly, exposure to light only Light-entrainable peripheral clocks
following the so-called mid-blastula transition is
effective at setting the phase of these pineal organ The cloning of the first mouse clock gene clock by
rhythms (Ziv and Gothilf, 2006). Given that this Joe Takahashi’s group was a milestone for the cir-
represents the step where zygotic transcription is cadian clock field (King et al., 1997). For the
initiated, this suggests that regulation of transcrip- zebrafish, this was also an important step since
tion may well be a critical step in the light the zebrafish homolog of clock became the first
response. Consistent with this prediction, light- tool for studying the general molecular organiza-
induced increases in the mRNA expression of tion of the circadian timing system (Whitmore
per2 and 6-4 DNA photolyase have been detected et al., 1998). These studies lead to the conclusion
as early as at the blastula and gastrula stages that as in the mouse, cycling clock gene expression
(Tamai et al., 2004). Further, knockdown of per2 was a property of most organs and tissues. Further,
expression (by injection of one cell stage em- these rhythms persisted in organs that were
bryos with morpholino-modified antisense oli- explanted into primary cultures showing that they
gonucleotides that block either translation were not “driven” by systemic cues—but were
initiation or correct splicing of the per2 transcript) actually generated by self-sustaining, independent
abolished the effects of early light exposure on the “peripheral” circadian oscillators. This conclusion
pineal (Ziv and Gothilf, 2006). Given the key role was entirely consistent with the first results from
for the mouse per2 homolog in the entrainment of the mouse and also Drosophila pointing to the
the clock by light (Albrecht et al., 2001; Zheng existence of peripheral clocks (Plautz et al., 1997;
et al., 2001), this points to the light-induced tran- Schibler and Sassone-Corsi, 2002). The big sur-
scription of the zebrafish per2 gene being a critical prise came when in a subsequent study, these
step in the maturation of the clock during embryo- organ cultures were exposed to various LD cycles
genesis (Ziv and Gothilf, 2006; Ziv et al., 2005). (Whitmore et al., 2000). Remarkably, the periph-
Interestingly, while both clock and bmal genes eral clocks were sustained and entrained by
show rhythmic expression at the mRNA level in the LD cycle, leading to the conclusion that in this
adult tissues, this expression is absent in the early vertebrate, as in the case of the fruit fly, tissues
embryonic stages. Rhythmic clock and bmal such as the heart and kidney were directly photo-
mRNA expression as measured in entire embryo sensitive. Later, directly light-entrainable clocks
48

were even observed in zebrafish cell lines zebrafish cell lines (Moutsaki et al., 2003).
(Whitmore et al., 2000). Thus, instead of acute Recently, functional analysis has shown that
changes in culture medium conditions such as TMT opsin serves as a photoreceptor (Cavallari
serum shocks (Balsalobre et al., 1998), clocks et al., 2011). Also exorhodopsin appears to act as
could be entrained in the zebrafish cell cultures a photoreceptor-regulating gene expression in
simply and noninvasively by exposure to light the pineal organ and in turn affects the production
(Vallone et al., 2004). Rhythms that have been of melatonin (Pierce et al., 2008). (2) Crys. The
established in zebrafish cell cultures subsequently argument for Crys being involved is based on sev-
slowly dampen when the cells are transferred to eral lines of evidence: the role played by Cry in
DD. Using single-cell imaging of bioluminescent photoreception of lateral neurons in Drosophila
reporter cell lines, this dampening has been shown (Ceriani et al., 1999), and the fact that one of the
to result from a progressive desynchronization of zebrafish cry gene homologs (cry4) shares closer
inaccurate single-cell clocks (Carr and Whitmore, sequence similarity with drosophila cry than with
2005). After extended periods in DD, the cells dis- the mammalian homologs (Kobayashi et al.,
play widely distributed phases and also marked 2000). Moreover, blue light wavelengths via
stochastic fluctuations in free-running period. retinol-independent photopigments were shown
Light exposure serves not only to reset the phase to upregulate per2 mRNA expression in cell
and thereby synchronize the single-cell oscillators culture (Cermakian et al., 2002). (3) A third model
but also to stabilize the free-running period (Carr implicates the production of hydrogen peroxide by
and Whitmore, 2005). The mechanism whereby a phototransducing flavin-containing oxidase. This
deprivation of light results in increasingly stochas- hypothesis stems from the observation that light
tic clock properties remains unclear. In this regard, triggers increases in intracellular levels of hydro-
evidence obtained from a study of per3::luc trans- gen peroxide in the zebrafish Z3 cell line
genic fish has revealed significant tissue-specific (Hirayama et al., 2007). It has been proposed that
differences in peripheral clock properties (Kaneko light-activated enzymes such as flavin-containing
et al., 2006). Clock parameters such as free- oxidases may trigger accumulation of H2O2 in
running period length, phase, light response, and the Z3 cells upon exposure to near violet–blue
temperature compensation all vary from tissue to wavelengths of light. This then results in activation
tissue, possibly reflecting differential expression of the MAPK signaling pathway that in turn drives
patterns of members of the various clock gene light-dependent activation of genes such as cry1a
families. and per2. Light was also shown to induce catalase
All these findings raise fundamental questions: expression that is then predicted to serve to
What are the widely expressed photoreceptors, downregulate the light response (Hirayama et al.,
and how do they signal to the clock? Do central 2007). Of course, a fourth possibility is the involve-
pacemakers in the fish coordinate the peripheral ment of not one but a combination of separate
clocks as they do in mammals, and if so, what is photoreceptor systems.
the underlying mechanism?
The identity of the peripheral photoreceptors
remains unclear. Three possible candidates have Light-inducible clock gene expression
been proposed: (1) Extra retinal opsins. In teleosts
and other nonmammalian vertebrates, a group How does the light-activated photoreceptor(s)
of opsin-like genes are expressed widely outside signal in the cell? Several reports have
the retina. For example, teleost multiple tissue documented light-dependent activation of the
(TMT) opsin mRNA is detected in the central ner- MAPK signaling pathway. Transient increases in
vous system, most peripheral tissues and even in pERK and pMEK levels have been documented
49

within 30 min of exposing zebrafish cell lines to promoters from other vertebrates, where interest-
light (Cermakian et al., 2002; Hirayama et al., 2007, ingly light exposure does not directly affect clock
2009). This result resembles the situation in the function. This suggests a key, ancestral role for
mammalian suprachiasmatic nucleus (SCN) where D box enhancers in clock gene regulation. In
photic signals indirectly relayed by the retino-hypo- mammals, four transcription factors have shown
thalamic tract trigger NMDA receptor activation to directly interact with and regulate gene expres-
and thereby activate the MAPK pathway in SCN sion from the D box, with three of these function-
neurons. In turn, phosphorylation of the trans- ing as activators (TEF, thyrotroph embryonic
criptional activator, CREB (cAMP-responsive factor; DBP, albumin D box-binding protein;
element binding protein), as well as chromatin and HLF, hepatic leukemia factor) and a fourth,
remodeling events have been proposed to repre- E4BP4, that serves as a repressor (see Fig. 1).
sent key steps in the light-induced activation of These factors have been studied extensively due
the per1 and 2 genes and then the subsequent to their role in clock output pathways in the
resetting of the phase of the mouse clock mouse (Gachon, 2007; Gachon et al., 2006;
(Travnickova-Bendova et al., 2002). A key obser- Lavery et al., 1999). In contrast, partly as the
vation that served as a starting point for the eluci- result of the teleost genome duplication events,
dation of the light-sensing mechanisms in in zebrafish 12 different factors with close homol-
zebrafish tissues has been that light exposure ogy to the mammalian D box regulatory factors
directly induces expression of certain clock genes have been cloned and characterized (Ben-Moshe
(per2 and cry1a) even in fish cell lines. et al., 2010) (Fig. 1). Interestingly, these factors
Interestingly, one of the sites showing the frequently show clock and light regulation. Fur-
strongest induction of per2 is the pineal organ. ther, in situ hybridization analysis has revealed
Further, it is now clear that light exposure not very different tissue specific and temporal expres-
only affects the expression of clock genes. A sys- sion patterns, frequently with highest expression
tematic microarray analysis has been performed levels in the pineal organ (Ben-Moshe et al.,
to assess the effects of light exposure on the 2010). Morpholino knockdown of one of the acti-
transcriptome of zebrafish cells and tissues; spe- vator proteins, TEF1, has been shown to severely
cifically cell lines, cultured adult zebrafish hearts, attenuate light-induced per2 expression in the
and 5-days-old larvae. The results point to a pineal, pointing to a key role for this factor in
group of around 100 genes whose expression is the pineal organ light transduction pathway
significantly upregulated upon light exposure. (Vatine et al., 2009). Clearly, an important future
These genes belong to diverse functional groups goal will be to assess the relative contribution of
including notably genes involved in repairing the various D box-binding factors to the light
DNA damage induced by exposure to the ultravi- response (Fig. 2).
olet (UV) wavelengths of sunlight. Comparison of
the predicted promoter regions of these light-
induced genes revealed an enrichment of the Blind cavefish reveal circadian clock
so-called D box enhancer elements. The impor- photoreceptors
tance of these enhancer elements in light-induced
gene expression has been confirmed by promoter Unexpected genetic evidence for the identity of
analysis of the per2 gene. A critical light-responsive the peripheral photoreceptors has come from
module within this promoter contains a D box as tackling a fundamental question concerning circa-
well as an E box whose integrity is critical dian clock evolution. What happens when animals
for light-induced gene expression. This region evolve completely isolated from the day–night
is also highly conserved in the per2 gene cycle? Does the clock persist? If so, is it still
50

tef, hlf, dbp


E4BP4

ztef ztef2 zhlf1 zhlf2 zdbp1 zdbp2


P F P P
P P H P
E H
E R
E

ze4bp4-1 ze4bp4-2 ze4bp4-5 ze4bp4-6


F P P P
R
R R

Fig. 1. Zebrafish D box-binding factors. Upper panel: schematic representation of the general structure of the activator (tef, hlf, and
dbp) PAR bZip transcription factors and the repressor E4BP4 factors. Red box represents the PAR domain; a purple box, the basic
region; and a blue box denotes the leucine zipper domain. Lower panel: in situ hybridization of zebrafish embryos with the various
D box-binding factor cDNAs. A clock or light symbol above each panel indicates whether the gene’s expression is predominantly
clock regulated or light driven, respectively. The identity of key anatomical structures within the embryo is indicated by letters.
P ¼ pineal gland, E ¼ eye, R ¼ retina, F ¼ forebrain, H ¼ hypothalamus.

entrained by light? Fish represent fascinating In the case of the zebrafish under LD cycles, a
models to address these questions. Around 100 spe- robust diurnal pattern of locomotor activity as well
cies of fish worldwide inhabit subterranean caves as rhythmic clock gene expression is observed. In
where they are isolated from daylight and have contrast, exposure of the blind cavefish to LD
experienced constant conditions in some cases cycles resulted in arrhythmicity in terms of both
for millions of years. These cavefish species locomotor activity and clock gene expression
exhibit a set of striking convergent “troglomorphic” (Fig. 3). Likewise, a cell line derived from adult
phenotypes including notably eye loss as well as loss cavefish fins and exposed to LD cycles failed to
of pigment and lowering of metabolic rate. Do exhibit the cycling clock gene expression exten-
cavefish still retain clocks? This question has been sively documented in zebrafish cell lines under
recently explored by comparing zebrafish with a identical conditions. These findings could have
species of blind cavefish at the behavioral, cellular, two alternative explanations. Either the cavefish
and molecular levels (Cavallari et al., 2011). The possess a blind circadian clock, unable to respond
Somalian cavefish, Phreatichthys andruzzii, was to light, or alternatively, they may have completely
chosen due to the severity of its troglomorphic lost circadian clock function. To distinguish
phenotype, the consequence of complete isolation between these two possibilities, the effects of
from sunlight beneath the Somalian desert for a exposing the cavefish to an alternative zeitgeber
predicted 2 million years. were tested. A regular feeding time has been well
51

Fig. 2. Schematic representation of the light input pathway in zebrafish cells. Light is perceived by opsins at the cell membrane and
relayed to the nucleus by various signal transduction pathways. Here, the light signal regulates the activity of a large family of D
box-binding factors that function as activators and repressors and can bind to the D box enhancer in various homo- and
heterodimeric combinations. D box-regulated genes include key clock genes such as per2 as well as other classes of gene
including those involved in the repair of UV-damaged DNA. The activation of a subset of the genes encoding the D box-binding
factors by light (TEF and E4BP4-6) serves as an additional regulatory loop that may contribute to defining the kinetics of light-
induced gene expression.

documented to entrain the circadian clock in many persisted even during the subsequent starvation
animals, including zebrafish, in which feeding period. In the case of the cavefish cell cultures,
cycles entrain clock gene expression in the liver, cells were transiently treated with dexamethasone,
but not in the brain (Lopez-Olmeda et al., 2010). a treatment widely used with mammalian cell lines
Thus, fish were fed regularly (once each 24 h) for to synchronize single-cell oscillators. This treat-
a period of 1–2 months followed by a period of ment led to rhythmic expression of a clock-
starvation. This led to increases in locomotor activ- regulated reporter gene, which surprisingly
ity that preceded the time of feeding, so-called exhibited a 47-h free-running period. Further,
food anticipatory activity (FAA) (see Fig. 3). Fur- changes in temperature led to alterations in the
ther, rhythmic expression of clock genes was length of the free-running period, suggesting a
established in the brain, heart, and liver that reduction in the temperature compensation of
(a) (b)

(c) (d)

Days
10

10

15

15

0 12 0 12 0 12 0 12
ZT ZT

(e) Feeding time


(f) Feeding time

5
5

10
10
Days

15
15

20
20

25
25

30
30
12 0 12 0 12 12 0 12 0 12
ZT ZT

Fig. 3. The pattern of zebrafish (Danio rerio, a) and cavefish (Phreatichthys andruzzii, b) locomotor activity during exposure to LD
cycles (c and d, respectively) and periods when food was delivered regularly at the same time each 24 h (e and f, respectively).
Zebrafish exhibit a robust diurnal pattern of activity under LD conditions and show increased activity just prior to the time of
feeding (so-called food anticipatory activity, FAA). In contrast, cavefish activity is arrhythmic under LD cycles, but like the
zebrafish, blind cavefish do show characteristic FAA. The number of days is indicated on the y-axis, while the time is plotted on
the x-axis as zeitgeber time (ZT) where ZT0 is defined either as “lights-on” or as feeding time.
53

the clock mechanism. Thus it would appear that entrainment. One unanswered question remains:
this species of cavefish does retain a clock, albeit Which mutations are responsible for the
an abnormal one that is no longer entrained by extremely long free-running period of this fish?
light but can still be entrained by feeding time So far, no mutations have been encountered in
(Cavallari et al., 2011). the core clock genes. This would tend to impli-
The molecular basis of the blindness of the cate the existence of mutations in other regu-
cavefish peripheral clocks was tracked down latory elements that may influence the stability
by the cloning and characterization of the of clock proteins or their subcellular location.
melanopsin and TMT opsin gene orthologs. Both Thus, the cloning and characterization of the
opsins are widely expressed in fish tissues and mutations responsible for this aberrant cavefish
encode seven-transmembrane domain proteins clock promise to teach us much about the
with the seventh membrane-spanning segment elements of the core circadian clock. Further,
being covalently linked with a chromophore, ret- the discovery of this aberrant clock mechanism
inaldehyde. In the cavefish, both opsin genes car- in a species of blind cavefish represents an impor-
ried mutations that lead to truncation of the tant step in studying evolution under extreme
proteins between the fifth and sixth membrane- environmental conditions. It raises many basic
spanning segments. The truncated protein would questions: Does the clock phenotype confer some
be predicted not to bind to the chromophore selective advantage for the fish in their subterra-
and therefore not to serve as a functional photo- nean cave environment? Alternatively, does this
pigment. To test whether the mutated opsins reflect the absence of positive selective pressure
could be responsible for the blind peripheral to maintain a normal clock and so are we wit-
clock phenotype, cavefish cells were transfected nessing gradual loss of a mechanism that confers
with expression vectors encoding the wild-type no selective advantage? Comparative studies of
zebrafish versions of these opsins. Light-induced other species of cavefish that have evolved inde-
expression of a per2 promoter-luc reporter gene pendently in different hypogean biotopes will
was restored in these cells, a result consistent hopefully provide some answers to these
with the opsin gene mutations contributing to questions.
the clock blindness. Interestingly, however,
the cavefish data also point to the existence of
additional mutations affecting other photopig- Concluding remarks
ments (Cavallari et al., 2011). While TMT and
melanopsin both react to blue and green light Teleosts represent the largest and one of the most
wavelengths, zebrafish cells respond to a much remarkable groups of vertebrates with unique cha-
broader range of wavelengths extending to red racteristics for chronobiological studies. With over
wavelengths. Thus, this predicts the existence of 30,000 known species, they exhibit a considerable
additional photoreceptors that have also degree of anatomical and physiological plasticity
accumulated mutations in the cavefish (see that has enabled them to occupy a wide diversity
Fig. 2). The loss of a functional light input path- of habitats and ecological niches. In the highly
way but the retention of a food-entrainable oscil- dynamic aquatic environment, one can predict that
lator make this fish an extremely attractive model during teleost evolution the circadian timing system
for studying how feeding regulates the clock. has been subjected to a range of different selective
The ability of this blind cavefish species to pressures depending on a diversity of lighting and
survive long periods of fasting adds to its suitabil- temperatures as well as the flexible patterns of
ity for studying this process of feeding behavior adopted. Thus, the initial studies in the
54

zebrafish as well as teaching us much about TMT teleost multiple tissue


key aspects of the vertebrate timing system also UV ultraviolet
promise to serve as a foundation for broader, com- ZT zeitgeber time
parative studies involving diverse species such as
the blind cavefish. These future studies should
tackle basic questions concerning the evolution of
References
the circadian timing system.
Albrecht, U. (2004). The mammalian circadian clock: A net-
work of gene expression. Frontiers in Bioscience, 9, 48–55.
Acknowledgments Albrecht, U., Zheng, B., Larkin, D., Sun, Z. S., & Lee, C. C.
(2001). MPer1 and mper2 are essential for normal resetting
M. L. I. was supported by the Italian percorso ad of the circadian clock. Journal of Biological Rhythms, 16,
100–104.
alta formazione Master and Back for the Sardin- Alon, S., Eisenberg, E., Jacob-Hirsch, J., Rechavi, G.,
ian region. D. V. and N. S. F. were supported by Vatine, G., Toyama, R., et al. (2009). A new cis-acting regu-
the CNRS, Max-Planck Institute, Tübingen, and latory element driving gene expression in the zebrafish
the Hermann-von-Helmholtz-Gemeinschaft. F. J. pineal gland. Bioinformatics, 25, 559–562.
S. V. received financial support from SENECA Appelbaum, L., Anzulovich, A., Baler, R., & Gothilf, Y. (2005).
Homeobox-clock protein interaction in zebrafish. A shared
(08743/PI/08) and MICINN (Aquagenomics and mechanism for pineal-specific and circadian gene expression.
Circasole). C. B. was supported by funding from Journal of Biological Chemistry, 280, 11544–11551.
the University of Ferrara (Italy), MIUR (Italy) Appelbaum, L., Toyama, R., Dawid, I. B., Klein, D. C.,
projects Azione Integrata Italia-Spagna, the Baler, R., & Gothilf, Y. (2004). Zebrafish serotonin-N-
acetyltransferase-2 gene regulation: Pineal-restrictive down-
VIGONI program of the DAAD, and the AIT-
stream module contains a functional E-box and three
MIUR. Y. G. was supported by Grant No. 1200/ photoreceptor conserved elements. Molecular Endocrinol-
08 from the Israel Science Foundation, Jerusalem, ogy, 18, 1210–1221.
Israel. Appelbaum, L., Vallone, D., Anzulovich, A., Ziv, L., Tom, M.,
Foulkes, N. S., et al. (2006). Zebrafish arylalkylamine-
N-acetyltransferase genes—Targets for regulation of the cir-
cadian clock. Journal of Molecular Endocrinology, 36,
Abbreviations 337–347.
Balsalobre, A., Damiola, F., & Schibler, U. (1998). A serum
AANAT serotonin-N-acetyl-transferase shock induces circadian gene expression in mammalian tis-
CREB cAMP-responsive element binding sue culture cells. Cell, 93, 929–937.
protein Ben-Moshe, Z., Vatine, G., Alon, S., Tovin, A., Mracek, P.,
Foulkes, N. S., et al. (2010). Multiple PAR and E4BP4 bZIP
Cry cryptochrome transcription factors in zebrafish: Diverse spatial and tempo-
DBP albumin D box-binding protein ral expression patterns. Chronobiology International, 27,
DD constant darkness 1509–1531.
FAA food anticipatory activity Brittijn, S. A., Duivesteijn, S. J., Belmamoune, M.,
Bertens, L. F., Bitter, W., de Bruijn, J. D., et al. (2009).
HLF hepatic leukemia factor
Zebrafish development and regeneration: New tools for
hpf hours postfertilization biomedical research. The International Journal of Develop-
LD light–dark mental Biology, 53, 835–850.
luc luciferase Cahill, G. M. (1996). Circadian regulation of melatonin pro-
PCE photoreceptor conserved duction in cultured zebrafish pineal and retina. Brain
Research, 708, 177–181.
element
Cahill, G. M. (1997). Circadian melatonin rhythms in cultured
per period zebrafish pineals are not affected by catecholamine receptor
SCN suprachiasmatic nucleus agonists. General and Comparative Endocrinology, 105,
TEF thyrotroph embryonic factor 270–275.
55

Cahill, G. M. (2002). Clock mechanisms in zebrafish. Cell and fish pineal organ and retina. Journal of Neuroendocrinology,
Tissue Research, 309, 27–34. 15, 378–382.
Cahill, G. M., Hurd, M. W., & Batchelor, M. M. (1998). Circa- Falcon, J., Migaud, H., Munoz-Cueto, J. A., & Carrillo, M.
dian rhythmicity in the locomotor activity of larval zebrafish. (2010). Current knowledge on the melatonin system in tele-
Neuroreport, 9, 3445–3449. ost fish. General and Comparative Endocrinology, 165,
Carr, A. J., & Whitmore, D. (2005). Imaging of single light- 469–482.
responsive clock cells reveals fluctuating free-running per- Gachon, F. (2007). Physiological function of PARbZip circa-
iods. Nature Cell Biology, 7, 319–321. dian clock-controlled transcription factors. Annals of Medi-
Cavallari, N., Frigato, E., Vallone, D., Frohlich, N., Lopez- cine, 39, 562–571.
Olmeda, J. F., Foa, A., et al. (2011). A blind circadian clock Gachon, F., Olela, F. F., Schaad, O., Descombes, P., &
in cavefish reveals that opsins mediate peripheral clock pho- Schibler, U. (2006). The circadian PAR-domain basic leu-
toreception. PLoS Biology, 9, e1001142. cine zipper transcription factors DBP, TEF, and HLF mod-
Ceriani, M. F., Darlington, T. K., Staknis, D., Mas, P., Petti, A. A., ulate basal and inducible xenobiotic detoxification. Cell
Weitz, C. J., et al. (1999). Light-dependent sequestration of Metabolism, 4, 25–36.
TIMELESS by CRYPTOCHROME. Science, 285, 553–556. Gothilf, Y., Coon, S. L., Toyama, R., Chitnis, A.,
Cermakian, N., Pando, M. P., Thompson, C. L., Namboodiri, M. A., & Klein, D. C. (1999). Zebrafish
Pinchak, A. B., Selby, C. P., Gutierrez, L., et al. (2002). serotonin N-acetyltransferase-2: Marker for development
Light induction of a vertebrate clock gene involves signaling of pineal photoreceptors and circadian clock function.
through blue-light receptors and MAP kinases. Current Endocrinology, 140, 4895–4903.
Biology, 12, 844–848. Haffter, P., Granato, M., Brand, M., Mullins, M. C.,
Cermakian, N., Whitmore, D., Foulkes, N. S., & Sassone- Hammerschmidt, M., Kane, D. A., et al. (1996). The identi-
Corsi, P. (2000). Asynchronous oscillations of two zebrafish fication of genes with unique and essential functions in the
CLOCK partners reveal differential clock control and func- development of the zebrafish, Danio rerio. Development,
tion. Proceedings of the National Academy of Sciences of the 123, 1–36.
United States of America, 97, 4339–4344. Hirayama, J., Cho, S., & Sassone-Corsi, P. (2007). Circadian
Davie, A., Sanchez, J. A., Vera, L. M., Sanchez-Vazquez, J., & control by the reduction/oxidation pathway: Catalase
Migaud, H. (2011). Ontogeny of the circadian system during represses light-dependent clock gene expression in the
embryogenesis in rainbow trout (Oncorhynchus mykyss) zebrafish. Proceedings of the National Academy of Sciences
and the effect of prolonged exposure to continuous illumina- of the United States of America, 104, 15747–15752.
tion on daily rhythms of per1, clock, and aanat2 expression. Hirayama, J., Fukuda, I., Ishikawa, T., Kobayashi, Y., &
Chronobiology International, 28, 177–186. Todo, T. (2003). New role of zCRY and zPER2 as regu-
DeBruyne, J., Hurd, M. W., Gutierrez, L., Kaneko, M., lators of sub-cellular distributions of zCLOCK and zBMAL
Tan, Y., Wells, D. E., et al. (2004). Isolation and phen- proteins. Nucleic Acids Research, 31, 935–943.
ogenetics of a novel circadian rhythm mutant in zebrafish. Hirayama, J., Miyamura, N., Uchida, Y., Asaoka, Y.,
Journal of Neurogenetics, 18, 403–428. Honda, R., Sawanobori, K., et al. (2009). Common light sig-
Dekens, M. P., Santoriello, C., Vallone, D., Grassi, G., naling pathways controlling DNA repair and circadian clock
Whitmore, D., & Foulkes, N. S. (2003). Light regulates the entrainment in zebrafish. Cell Cycle, 8, 2794–2801.
cell cycle in zebrafish. Current Biology, 13, 2051–2057. Hurd, M. W., Debruyne, J., Straume, M., & Cahill, G. M.
Dekens, M. P., & Whitmore, D. (2008). Autonomous onset of (1998). Circadian rhythms of locomotor activity in zebrafish.
the circadian clock in the zebrafish embryo. The EMBO Physiology and Behavior, 65, 465–472.
Journal, 27, 2757–2765. Ingham, P. W. (2009). The power of the zebrafish for disease
del Pozo, A., Sanchez-Ferez, J. A., & Sanchez-Vazquez, F. J. analysis. Human Molecular Genetics, 18, R107–112.
(2011). Circadian rhythms of self-feeding and locomotor Ishikawa, T., Hirayama, J., Kobayashi, Y., & Todo, T. (2002).
activity in zebrafish (Danio Rerio). Chronobiology Interna- Zebrafish CRY represses transcription mediated by
tional, 28, 39–47. CLOCK-BMAL heterodimer without inhibiting its binding
Delaunay, F., Thisse, C., Thisse, B., & Laudet, V. (2003). Dif- to DNA. Genes to Cells, 7, 1073–1086.
ferential regulation of Period 2 and Period 3 expression dur- Kaneko, M., & Cahill, G. M. (2005). Light-dependent develop-
ing development of the zebrafish circadian clock. Gene ment of circadian gene expression in transgenic zebrafish.
Expression Patterns, 3, 319–324. PLoS Biology, 3, e34.
Emery, P., & Reppert, S. M. (2004). A rhythmic Ror. Neuron, Kaneko, M., Hernandez-Borsetti, N., & Cahill, G. M. (2006).
43, 443–446. Diversity of zebrafish peripheral oscillators revealed by
Falcon, J., Gothilf, Y., Coon, S. L., Boeuf, G., & Klein, D. C. luciferase reporting. Proceedings of the National Academy
(2003). Genetic, temporal and developmental differences of Sciences of the United States of America, 103,
between melatonin rhythm generating systems in the teleost 14614–14619.
56

Kazimi, N., & Cahill, G. M. (1999). Development of a circa- Plautz, J. D., Kaneko, M., Hall, J. C., & Kay, S. A. (1997).
dian melatonin rhythm in embryonic zebrafish. Brain Independent photoreceptive circadian clocks throughout
Research. Developmental Brain Research, 117, 47–52. Drosophila. Science, 278, 1632–1635.
King, D. P., Zhao, Y., Sangoram, A. M., Wilsbacher, L. D., Postlethwait, J. H., Yan, Y. L., Gates, M. A., Horne, S.,
Tanaka, M., Antoch, M. P., et al. (1997). Positional cloning Amores, A., Brownlie, A., et al. (1998). Vertebrate genome
of the mouse circadian clock gene. Cell, 89, 641–653. evolution and the zebrafish gene map. Nature Genetics, 18,
Kobayashi, Y., Ishikawa, T., Hirayama, J., Daiyasu, H., 345–349.
Kanai, S., Toh, H., et al. (2000). Molecular analysis of Reppert, S. M., & Weaver, D. R. (2002). Coordination of cir-
zebrafish photolyase/cryptochrome family: Two types of cadian timing in mammals. Nature, 418, 935–941.
cryptochromes present in zebrafish. Genes to Cells, 5, Schibler, U., & Sassone-Corsi, P. (2002). A web of circadian
725–738. pacemakers. Cell, 111, 919–922.
Korf, H. W., Schomerus, C., & Stehle, J. H. (1998). The pineal Shearman, L. P., Zylka, M. J., Reppert, S. M., &
organ, its hormone melatonin, and the photoneuroendocrine Weaver, D. R. (1999). Expression of basic helix-loop-helix/
system. Advances in Anatomy, Embryology, and Cell Biol- PAS genes in the mouse suprachiasmatic nucleus. Neurosci-
ogy, 146, 1–100. ence, 89, 387–397.
Lavery, D. J., Lopez-Molina, L., Margueron, R., Fleury- Tamai, T. K., Vardhanabhuti, V., Foulkes, N. S., &
Olela, F., Conquet, F., Schibler, U., et al. (1999). Circadian Whitmore, D. (2004). Early embryonic light detection
expression of the steroid 15 alpha-hydroxylase (Cyp2a4) improves survival. Current Biology, 14, R104–105.
and coumarin 7-hydroxylase (Cyp2a5) genes in mouse Tamai, T. K., Young, L. C., & Whitmore, D. (2007). Light sig-
liver is regulated by the PAR leucine zipper transcription naling to the zebrafish circadian clock by Cryptochrome 1a.
factor DBP. Molecular and Cellular Biology, 19, Proceedings of the National Academy of Sciences of the
6488–6499. United States of America, 104, 14712–14717.
Lopez-Olmeda, J. F., Tartaglione, E. V., de la Iglesia, H. O., Tan, Y., DeBruyne, J., Cahill, G. M., & Wells, D. E. (2008).
& Sanchez-Vazquez, F. J. (2010). Feeding entrainment Identification of a mutation in the Clock1 gene affecting
of food-anticipatory activity and per1 expression in the zebrafish circadian rhythms. Journal of Neurogenetics, 22,
brain and liver of zebrafish under different lighting and 149–166.
feeding conditions. Chronobiology International, 27, Travnickova-Bendova, Z., Cermakian, N., Reppert, S. M., &
1380–1400. Sassone-Corsi, P. (2002). Bimodal regulation of
Masai, I., Heisenberg, C. P., Barth, K. A., Macdonald, R., mPeriod promoters by CREB-dependent signaling and
Adamek, S., & Wilson, S. W. (1997). Floating head and CLOCK/BMAL1 activity. Proceedings of the National Acad-
masterblind regulate neuronal patterning in the roof of the emy of Sciences of the United States of America, 99, 7728–7733.
forebrain. Neuron, 18, 43–57. Vallone, D., Gondi, S. B., Whitmore, D., & Foulkes, N. S.
Moutsaki, P., Whitmore, D., Bellingham, J., Sakamoto, K., (2004). E-box function in a period gene repressed by light.
David-Gray, Z. K., & Foster, R. G. (2003). Teleost multiple Proceedings of the National Academy of Sciences of the
tissue (tmt) opsin: A candidate photopigment regulating the United States of America, 101, 4106–4111.
peripheral clocks of zebrafish? Brain Research. Molecular Vatine, G., Vallone, D., Appelbaum, L., Mracek, P., Ben-
Brain Research, 112, 135–145. Moshe, Z., Lahiri, K., et al. (2009). Light directs zebrafish
Mullins, M. C., Hammerschmidt, M., Haffter, P., & Nüsslein- period2 expression via conserved D and E boxes. PLoS
Volhard, C. (1994). Large-scale mutagenesis in the Biology, 7, e1000223.
zebrafish: In search of genes controlling development in a Wang, H. (2008a). Comparative analysis of period genes in tel-
vertebrate. Current Biology, 4, 189–202. eost fish genomes. Journal of Molecular Evolution, 67, 29–40.
Nüsslein-Volhard, C., & Dahm, R. (2002). Zebrafish: A practi- Wang, H. (2008b). Comparative analysis of teleost fish
cal approach. Oxford: Oxford University Press. genomes reveals preservation of different ancient clock
Petzold, A. M., Balciunas, D., Sivasubbu, S., Clark, K. J., duplicates in different fishes. Marine Genomics, 1, 69–78.
Bedell, V. M., Westcot, S. E., et al. (2009). Nicotine Wang, H. (2009). Comparative genomic analysis of teleost fish
response genetics in the zebrafish. Proceedings of the bmal genes. Genetica, 136, 149–161.
National Academy of Sciences of the United States of Whitmore, D., Foulkes, N. S., & Sassone-Corsi, P. (2000).
America, 106, 18662–18667. Light acts directly on organs and cells in culture to set the
Pierce, L. X., Noche, R. R., Ponomareva, O., Chang, C., & vertebrate circadian clock. Nature, 404, 87–91.
Liang, J. O. (2008). Novel functions for Period 3 and Exo- Whitmore, D., Foulkes, N. S., Strahle, U., & Sassone-Corsi, P.
rhodopsin in rhythmic transcription and melatonin biosynthe- (1998). Zebrafish clock rhythmic expression reveals
sis within the zebrafish pineal organ. Brain Research, 1223, independent peripheral circadian oscillators. Nature
11–24. Neuroscience, 1, 701–707.
57

Yáñez, J., Busch, J., Anadón, R., & Meissl, H. (2009). Pineal Academy of Sciences of the United States of America, 103,
projections in the zebrafish (Danio rerio): Overlap with ret- 4146–4151.
inal and cerebellar projections. Neuroscience, 164, Ziv, L., Levkovitz, S., Toyama, R., Falcon, J., & Gothilf, Y.
1712–1720. (2005). Functional development of the zebrafish pineal
Zheng, B., Albrecht, U., Kaasik, K., Sage, M., Lu, W., gland: Light-induced expression of period2 is required for
Vaishnav, S., et al. (2001). Nonredundant roles of the mPer1 onset of the circadian clock. Journal of Neuroendocrinology,
and mPer2 genes in the mammalian circadian clock. Cell, 17, 314–320.
105, 683–694. Ziv, L., Tovin, A., Strasser, D., & Gothilf, Y. (2007). Spectral
Ziv, L., & Gothilf, Y. (2006). Circadian time-keeping during sensitivity of melatonin suppression in the zebrafish pineal
early stages of development. Proceedings of the National gland. Experimental Eye Research, 84, 92–99.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 4

Two clocks in the brain: An update of the morning


and evening oscillator model in Drosophila

Taishi Yoshii{, Dirk Rieger{ and Charlotte Helfrich-Förster{,*

{
Graduate School of Natural Science and Technology, Okayama University, Okayama, Japan
{
Biocenter, University of Würzburg, Am Hubland, Würzburg, Germany

Abstract: Circadian clocks play an essential role in adapting the activity rhythms of animals to the day–night
cycles on earth throughout the four seasons. In many animals, including the fruit fly Drosophila melanogaster,
two separate but mutually coupled clocks in the brain —morning (M) and evening (E) oscillators— control
the activity in the morning and evening. M and E oscillators are thought to track dawn and dusk,
respectively. This alters the phase-angle between the two oscillators under different day lengths, optimally
adapting the animal’s activity pattern to colder short and warmer long days. Using excellent genetic tools,
Drosophila researchers have addressed the neural basis of the two oscillators and could partially track
these to distinct clock cells in the brain. Nevertheless, not all data are consistent with each other and many
questions remained open. So far, most studies about M and E oscillators focused on the influence of light
(photoperiod). Here, we will review the effects of light and temperature on the two oscillators, will update
the present knowledge, discuss the limitations of the model, and raise questions that have to be addressed
in the future.

Keywords: circadian clock; seasonal adaptation; dual oscillator model; temperature; light; Drosophila
melanogaster; clock neurons.

Introduction organisms living on earth have obtained time-


measuring structures in the brain, which help them
The rotation of the earth causes cyclical alterations to predict the regularly occurring environmental
in the environment of which light and temperature cycles and thus maximize fitness. Circadian clocks
oscillations seem to be the most important. The and circannual clocks evolved to anticipate daily
and yearly environmental changes, respectively.
Both clocks may be linked, as circadian clocks
*Corresponding author.
Tel.: þ49-931-318-8823; Fax: þ49-931-318-4452
can provide the necessary reference for measuring
E-mail: charlotte.foerster@biozentrum.uni-wuerzburg.de day length (Bünning, 1936). Changes in day length

http://dx.doi.org/10.1016/B978-0-444-59427-3.00027-7 59
60

together with annual temperature cycles (T-cycles) (Aschoff, 1966): birds, fishes, and several mammals
synchronize circannual clocks to the yearly cycle, show two activity bouts, one in the morning (M)
though it is still under debate whether circadian and one in the evening (E). M and E activity bouts
clocks participate in this synchronization in all are close together under short spring days and
organisms (for a review, see Bradshaw and separated by a pronounced siesta under long sum-
Holzapfel, 2010). In order to survive in the seasonal mer days (Aschoff, 1966; see Fig. 1b). These two
environmental changes that affect organisms, activity bouts persist even under constant light
animals have to anticipate the coming winter by pre- (LL) and temperature conditions in the lab,
paring themselves in time for hibernation or dia- indicating that they are endogenous. The effects of
pause. In addition, seasonal changes have a strong temperature on the two activity bouts have not been
impact on the animal’s daily activity pattern. Espe- systematically investigated in vertebrates, but the
cially in poikilothermic insects, the entire physiol- effects of light have been studied extensively. In
ogy (including the activity) depends highly on the nature, light and temperature are closely linked:
ambient temperature. Under short cold days in more visible light (e.g., on long summer days) is
spring and autumn, it is favorable for them to be accompanied by more infrared light, resulting in
active during the warmer parts of the day, whereas higher temperatures. Thus in the lab, the animal’s
it is better to shift activity to the morning, the late circadian clock may predict a warm summer day
evening, or even the night under long, hot summer from the amount of light perceived even if no heat
days (Fig. 1). There are two mechanisms that help is produced. Indeed, under constant conditions in
insects to be active at the best time of the day: the lab, M and E activity bouts of finches shift apart
(1) They respond quickly to the acute ambient tem- with increasing light intensity (Aschoff, 1966). The
perature and (2) their circadian clocks determine same happens in Syrian hamsters and other noctur-
the daily activity pattern under spring, summer, or nal rodents (reviewed by Schwartz and de la Iglesia,
autumn conditions. The latter mechanism implies 2003). At a certain light intensity, the two compo-
that the clock entrains differently (has different nents free-run with different periods until they
waveforms) at different day lengths. Indeed, this become stably coupled 180 apart, a phenomenon
was observed in nature, even in vertebrates called “splitting.”

(a) Longer yM,E Shorter yM,E

Hot day Comfortable Cold day

(b) Longer yM,E Shorter yM,E

long day LD 12:12 Short day

Fig. 1. Temperature and photoperiod influence the phases of M and E activity peaks. (A) Under hot days (left), the two peaks shift
in opposite directions giving rise to a wide trough, allowing the flies to rest during the heat of midday. In contrast, the two peaks are
close together under cold days, allowing the fly to be active during the warmer daytime. (B) Similar effects can also be seen under
long and short photoperiods.
61

The dual oscillator model manipulate specific regions of the SCN in order
to investigate the function of distinct neuronal
Based on the hamster’s “split rhythms” in LL, populations under different photoperiods.
Pittendrigh and Daan (1976) proposed that the All this is easier to achieve in the fruit fly
clock consists of two separate oscillators with dif- D. melanogaster. While the basic neuronal func-
ferent responses to light: one oscillator accelerates tion of endogenous clocks appears well con-
and the other decelerates upon light exposure. served between flies and mammals, the entire
Under light–dark (LD) cycles, the first oscillator brain clock of Drosophila consists of only  150
tracks dawn and is therefore called morning (M) clock neurons, and there are excellent genetic
oscillator and the second tracks dusk and is called tools by which specific clock neurons can be
evening (E) oscillator. Due to the different pro- manipulated. Even at the behavioral level, the
perties of the M and E oscillators, the two activity activity rhythms of fruit flies resemble in many
bouts in the morning and evening are close aspects those of vertebrates: Fruit flies show
together under short days and far apart under long bimodal activity patterns with M and E peaks,
days. Thus, the dual oscillator model can explain and these persist to some extent under constant
that M and E activity bouts change their phase- conditions (Helfrich-Förster, 2000). Under short
angle (CM,E), leading to different waveforms of and long photoperiods, CM,E becomes smaller
the activity rhythm at different photoperiods. or larger, respectively (Rieger et al., 2003;
For many years, the dual oscillator model 2012). The effects on CM,E are enhanced when
remained gray theory, very useful to explain short or long photoperiods are combined with
behavioral phenomena but without having any low or high constant temperatures, respectively
mechanistic background. Only in the past few years (Majercak et al., 1999). Even temperature alone
has it become clear that the two oscillators have a (under constant dark (DD) conditions) affects
neuronal basis. This view was propelled by studies CM,E in the same direction: CM,E is larger under
performed in the fruit fly Drosophila melanogaster high temperatures than under low temperatures
(Grima et al., 2004; Rieger et al., 2006; Stoleru (Majercak et al., 1999). LL alone is more diffi-
et al., 2004; see below). In mammals, Jagota et al. cult to test in Drosophila, because the blue-light
(2000) demonstrated first that the mammalian photopigment cryptochrome (CRY) renders flies
central pacemaker, the suprachiasmatic nucleus arrhythmic already at very low light intensities
(SCN), can exhibit two separate peaks in electrical (Emery et al., 2000a). But, in the absence of
activity that change C in response to photoperiod. CRY, CM,E becomes larger with increasing light
Consecutive studies at the molecular and electro- intensity and finally rhythm splitting and internal
physiological level unveiled that there are two or desynchronization into two free-running compo-
more types of cell populations in the SCN that nents occurs—very similar to the hamsters’ behav-
respond differently to short and long photoperiods ior in LL (Helfrich-Förster et al., 2001; Yoshii
(e.g., Brown and Piggins, 2009; Burgoon et al., et al., 2004). Higher light intensities even increase
2004; Hazlerigg et al., 2005; Inagaki et al., 2007; CM,E under 12:12 LD-cycles (Rieger et al., 2007),
Naito et al., 2008; Sosniyenko et al., 2009; reviewed indicating that the flies interpret a day with high
in Meijer et al., 2010). Nevertheless, it remained light intensity as a hot day.
difficult to trace the different oscillators to specific After M and E oscillators were anatomically
clock neurons in the SCN and to reveal the traced to specific clock neurons in the Drosophila
mechanisms by which they follow dawn and brain by the groups of Francois Rouyer and
dusk. One reason for this difficulty is the complex- Michael Rosbash (Grima et al., 2004; Stoleru
ity of the SCN that consists of more than 10,000 et al., 2004; see below), many papers appeared
neurons. In addition, it remains hard to genetically that have modified and refined the original dual
62

oscillator model whereby the results of these stud- Other neurotransmitters followed (see Fig. 2b)
ies have been controversially reviewed (Dubruille and their role in the clock is currently investigated.
and Emery, 2008; Helfrich-Förster, 2006, 2009; Knowing the genes expressed in the clock
Helfrich-Förster et al., 2007b; Sheeba et al., neurons is the first prerequisite to genetically manip-
2008b). In the meantime, more papers appeared ulate the neurons with the gal4-uas binary expres-
that challenged the original dual oscillator model, sion system of the yeast Saccharomyces cerevisiae
and it seems high time to write an elaborate (Brand and Perrimon, 1993). GAL4 is a transcrip-
update. We will do so, discuss new aspects as well tional activator that binds to the upstream activating
as the limitations of the model and put special sequence (uas) of other genes, thus activating their
emphasis on the effects of natural-like T-cycles transcription. As flies do not have the gal4 and uas
on M and E oscillators. sequences in their genome, these can be used to
express genes of interest in specific fly cells (tissues).
All one has to do is to fuse (1) the promoter of inter-
The clock network in the Drosophila brain est (e.g., that of the Pdf gene; Park and Hall, 1998)
and the possibility to manipulate selected with the gal4 gene and (2) the gene that one wants
clock neurons to express (e.g., the apoptotic gene head involution
defect (hid); McNabb et al., 1997) with the uas
Drosophila’s 150 clock neurons are located in sequence, and (3) to insert both constructs into the
the lateral and dorsal brain and are thus called fly genome (reviewed in Duffy, 2002). Flies con-
lateral and dorsal neurons (LNs and DNs). The taining both constructs will express hid in the Pdf-
LNs and DNs are further divided into different expressing cells, and as a consequence, the PDF
subclusters that are depicted in Fig. 2. All clock neurons will die early in development. The gal4-
neurons form a neuronal network that has been uas system enables not only to kill specific neurons
described in detail and been shown in several but also to manipulate them in many different ways,
reviews (a still valid detailed schematic presenta- for example, they can be electrically silenced by
tion of the clock network can be found in expressing Kþ channels (Nitabach et al., 2002), they
Helfrich-Förster et al., 2007a). Here, we just can be overexcited by expressing Naþ channels
focus on the neurochemistry of the different (Nitabach et al., 2006), neurotransmitter release
clock neurons that helped to manipulate them can be blocked by expressing tetanus toxin (e.g.,
selectively. All clock neurons express the core Blanchardon et al., 2001; Kaneko et al., 2000), and
clock proteins Period (PER), Timeless (TIM), known genes (e.g., per) can be overexpressed (e.g.,
Clock (CLK), and Cycle (CYC) (Houl et al., Murad et al., 2007; Yang and Sehgal, 2001). It is
2006; Kaneko and Hall, 2000). However, the also possible to “rescue” the expression of a gene
blue-light photopigment CRY that works as “circa- (e.g., per) in specific cells of null-mutant flies (e.g.,
dian photoreceptor” is not uniformly present in all per01 mutants; Grima et al., 2004). In 1999, Lee and
clock neurons (Fig. 2a; Benito et al., 2008; Yoshii Luo introduced a third component—GAL80, a sup-
et al., 2008). A further divergence of the clock pressor of GAL4 activity—which can be utilized to
neurons becomes evident when one regards their refine the spatial expression of gal4 with a second
neurotransmitters (neuropeptides) (Fig. 2b). The promoter fused to gal80.
neuropeptide pigment-dispersing factor (PDF) In the meantime, many gal4 drivers are avail-
was the first neurotransmitter found in eight LNs able that drive expression in different populations
per brain hemisphere (Fig. 2b; Helfrich-Förster, of the clock neurons (Fig. 2c). The first ones
1995). PDF and the PDF-containing LNs later developed are period (per)-gal4 and timeless
turned out to be quite important for the function (tim)-gal4 lines that drive gene expression in all
of the circadian clock under constant darkness clock cells (Blau and Young, 1999; Kaneko and
(Helfrich-Förster et al., 2000; Renn et al., 1999). Hall, 2000), a Pdf-gal4 line that drives gene
(a) (b)
DN1p

DN3
DN1a
PDF+ ITP+
DN2 PDF- ITP-

LNd
LPN
5th
S-LNv
I-LNv

s-LNv CRY+
NPF+ sNPF+
CRY- NPF- sNPF-

Cha+ IPNamide+
Cha- IPNamide-

(c)
cry-gal4 Mai179-gal4 Mz520-gal4 c929-gal4

r6-gal4 Mai179-gal4, Pdf-gal80 tim-gal4, cry-gal80 Clk4.1M-gal4

Fig. 2. Circadian clock neurons of D. melanogaster and their neurochemistry. All pictures show the right brain hemisphere. The
neurons that are located in the dorsal brain are called dorsal neurons (DN1–3 s). Among the DN1s, a pair of cells lies more anterior
and is called DN1anterior (DN1a). The remaining 15 DN1s have a more posterior position and are called DN1posterior (DN1p). The
DN2 are located slightly ventral from the DN1p. The DN3s represent the largest cluster and are composed of 30 cells, most of
which have very small somata. The lateral neurons (LNs) are located in a lateral position of the central brain. This group is also
further subdivided. One cluster of three cells, the lateral posterior neurons (LPNs), lies in a more central and posterior position.
The other clusters lie in the anterior brain. Among these, the most dorsal group consisting of six cells are called LNdorsal (LNds).
64

expression specifically in the PDF neurons (Park the per-gal4, tim-gal4 lines. But again the expres-
et al., 2000; Renn et al., 1999), a Clk-gal4 line sion pattern depends on the length of the cry-
(Glossop et al., 2003), as well as several cry-gal4 promoter fused to the gal4 sequence as well as on
lines (Emery et al., 2000b; Klarsfeld et al., 2004; the insertion site in the genome that strongly affects
Zhao et al., 2003). Not all of these drivers express expression strength. Some cry-gal4 lines (Emery
exclusively in the clock neurons. Animals possess et al., 2000a; Hao et al., 2008; Klarsfeld et al., 2004;
clocks also in most body organs (so-called periph- Picot et al., 2007; Zhao et al., 2003) express strongly
eral clocks) in addition to the brain clocks in all clock neurons, glia cells, and photoreceptor
(Dibner et al., 2010; Plautz et al., 1997). Therefore, and ring neurons, and others mainly in the CRY-
the core clock genes (per, tim, Clk, and cyc) are positive clock neurons and ring neurons, whereby
naturally expressed in many tissues—not only there are differences in the expression strength
in the brain. Within the head, natural expression (Helfrich-Förster et al., 2007a,b).
is found in the photoreceptor cells of the eyes In addition to the promoter-fused gal4 lines, there
and the ocelli as well as in many glia cells of the are gal4 drivers gained by the enhancer trap tech-
brain. However, the available per- and tim-gal4 nique (reviewed in Duffy, 2002). These may not nec-
lines drive expression even in cells that do not con- essarily be related to known clock genes (in several
tain native PER or TIM. This is because the spatial cases, it is still unknown which gene they trapped).
information is not only coded by the per- and tim- Nevertheless, several of these lines are very useful
promoters but also post-transcriptionally (Kaneko because they drive gene expression in restricted
and Hall, 2000; Stanewsky et al., 1997). In addi- subsets of the clock neurons. These are mai179-
tion, the expression pattern of gal4 depends on gal4, Mz520-gal4, c929-gal4, and r6-gal4 (Fig. 2c),
the length of the clock gene promoter used. This as well as gal1118 and gal1501 that drive essentially
is very evident for the existing Clk-gal4 drivers. in the PDF neurons (Blanchardon et al., 2001;
Clk is naturally expressed in all clock neurons plus Lamaze et al., 2011; not shown) and two clock work
many other cells (Houl et al., 2006). Fusing differ- orange (cwo) enhancer trap lines that drive in all
ent parts of the Clk-promoter to gal4 has also lead neurons (Kadener et al., 2007; not shown).
to Clk-gal4 lines with more restricted expression Useful gal80 lines for refining expression to cer-
(Gummadova et al., 2009). In 2010, a very interest- tain clock neurons are cry-gal80 and Pdf-gal80
ing DN1p-specific Clk-gal4 line —Clk4.1 M-gal4— (Stoleru et al., 2004). By combining tim-gal4 with
came out that contains the  0.2 to 0.5 kb fragment cry-gal80, one can manipulate all cells that express
of the Clk gene and drives expression specifically tim but not cry (most of the DNs). By combining
in the CRY-positive DN1p (Fig. 2c; Zhang et al., tim-gal4 with Pdf-gal80, only the tim-expressing
2010a,b). cells that have no PDF can be manipulated (the
CRY is found in subsets of the clock neurons DNs and half of the LNs), and by combining cry-
(Fig. 2a), in the compound eyes, glia cells, and ring gal4 with Pdf-gal80, one can manipulate half of
neurons of the ellipsoid body. Nevertheless, the the LNs and few DNs. By combining mai179-gal4
existing cry-gal4 lines are all more specific than with Pdf-gal80, it is even possible to restrict gene

The remaining more ventrally located clusters are classically called LNventral (LNvs). Most LNv cells express the neuropeptide pigment-
dispersing factor (b). The large LNvs (l-LNvs) have large somata and send projections to optic lobe and the contralateral side of the
brain. The small LNvs (s-LNvs) send projections to the dorsal brain where DN cells lie. One of s-LNv cells does not express PDF, is
located among the l-LNvs, and is called 5th s-LNv. (a) Cryptochrome (CRY)-positive neurons are highlighted in white (Benito et al.,
2008; Yoshii et al., 2008). (b) Neurons containing different neuropeptides/neurotransmitters are highlighted in white and marked by
arrows. PDF, pigment-dispersing factor (Helfrich-Förster, 1995); NPF, long neuropeptide F (Hermann et al., 2012; Johard et al., 2009;
Lee et al., 2006); Cha, choline acetyltransferase (Johard et al., 2009); ITP, ion transport peptide (Johard et al., 2009); sNPF, small
neuropeptide F (Johard et al., 2009); IPNamide (Shafer et al., 2006). (c) Useful gal4 lines or gal4/gal80 combinations that express only
in selected clock neurons (highlighted in white).
65

expression to four LNs per hemisphere (the 5th They combined a pan-neural gal4 driver (elav-
s-LNv and three LNd; see Fig. 2c and below). gal4) with either Pdf-gal80 or cry-gal80 to drive
In general, it is wise to study first the exact uas-per in per0 mutants. In the first case, they res-
expression pattern of any used gal4-line and gal4/ cued per in all neurons except the PDF-positive
gal80 combination by expressing a reporter gene LNs; in the second case, they rescued per in all
as the green fluorescent protein (gfp) gene before neurons except all LNs plus few DN1. One would
manipulating the neurons. Unfortunately, this was expect that in the first case, the flies will lack the M
not well done in every study. anticipation, and in the second case, M and E antici-
pation. However, this was not the case, showing that
the LNs are not the only neurons that are responsi-
The original studies of Stoleru et al. (2004) ble for M and E activity. Obviously, the DNs do also
and Grima et al. (2004) contain M and E oscillators. This reminds of the
results gained with a classical mutant, disco, in which
As mentioned earlier, fruit flies show bimodal all LN cells are developmentally absent but all
activity patterns with M and E peaks. Under DN groups are intact. Most disco mutants loose
12 h:12 h LD-cycles (LD 12:12), M and E activity free-running rhythms in DD but exhibit rather nor-
bouts begin to increase before lights-on and mal M and E peaks in LD (Blanchardon et al., 2001;
lights-off, respectively. This shows that the clock Dushay et al., 1989; Hardin et al., 1992; Helfrich-
anticipates morning and evening and the two activ- Förster, 1998; Wheeler et al., 1993), suggesting
ity bouts are not mere responses to lights-on and that the DN groups are sufficient for both peaks.
lights-off. The Rosbash and Rouyer groups con- The study of Grima et al. (2004), which was
centrated on this anticipatory activity increase published on consecutive pages in the same
and defined the M oscillator to be present if M Nature issue, nicely complemented the per-rescue
anticipation is seen and the E oscillator to be pres- experiments of Stoleru et al. (2004). These
ent if E anticipation is evident (Grima et al., 2004; authors used different gal4-drivers to restrict per
Stoleru et al., 2004). This simple definition allowed to certain clock neurons (Fig. 1c): (1) mai179-
them to screen for flies that lost the anticipations gal4 that is expressed in the s-LNv, 5th s-LNv,
after genetic manipulation. and three to four LNd; (2) c929-gal4 that is
The main strategy of Stoleru et al. (2004) was to expressed in the l-LNv plus many neurosecretory
selectively ablate clock neurons by expressing hid. cells that are not clock neurons (Taghert et al.,
They ablated the PDF-positive LNs (Fig. 1b) with 2001); (3) Pdf-gal4 and Mz520-gal4 that both
Pdf-gal4;uas-hid or the other LNs by combining express in the PDF-positive LNs. When per was
Pdf-gal80 with cry-gal4;uas-hid (note that all LNd, rescued in most LNs (s-LNv, 5th s-LNv, and
not only the CRY-positive ones, plus the 5th LNd) by mai179-gal4, M and E anticipations were
s-LNv and few DN1 are ablated by this driver restored, suggesting that these cells contain the M
combination). Interestingly, the flies lacking the and E oscillators. When per was only rescued in
PDF-positive neurons lost M anticipatory activity, the eight PDF-positive LNs by Pdf-gal4 and
while the flies lacking the other LNs (plus few Mz520-gal4, M anticipation was restored, but
DN1) lost E anticipation. This clearly suggests that when per was only rescued in the four PDF-
the M oscillator resides in the PDF-positive LNs positive l-LNs by c929-gal4, the flies behaved as
and the E oscillator in the PDF-negative ones (plus per01 mutants and did not show any anticipatory
few DN1). Thus, the two oscillators seem to reside M activity, suggesting the M oscillator resides in
in distinct clock neurons (mainly LNs). the s-LNvs. Together, these results suggest that
To confirm this conclusion, Stoleru et al. (2004) the M oscillator resides in the four PDF-positive
also took an opposite approach, in which they s-LNvs and the E oscillator in three to four LNds
rescued per expression in per null (per0) mutants. (and the 5th s-LNv; however, the 5th s-LNv was
66

not detected in the two studies; it was recognized peaks under the long photoperiod. Stoleru et al.
as an E oscillator only later (Rieger et al., 2006)). (2007) concluded that the M cells dominate in
Coincidentally, the two independent groups darkness and on long nights and the E cells in
arrived at exactly the same conclusion that M light and on long days. This differential dom-
and E oscillators reside in distinct clock neurons, inance of M and E oscillators may also be
which made their discovery quite solid and con- reflected in the amplitudes of M and E peaks in
vincing. Those exciting reports raised the dual wild-type flies under short and long photoperiods
oscillator model by Pittendrigh and Daan to a (Rieger et al., 2003): under short days (LD 8:16),
hot topic in Drosophila chronobiology. their M peak was higher than their E peak,
whereas the opposite was true under long days
(LD 16:8).
Dominance of the M cells under short days
and of the E cells under long days
Light activates output from the E cells and
In order to reveal the hierarchy of the two inhibits output from the M cells
oscillators as well as the communication between
them, Stoleru et al. (2005) accelerated the oscilla- Stoleru et al. (2007) found that flies over-
tion speed of the M cells or the E cells by selec- expressing sgg in E cells remained rhythmic in
tively overexpressing the shaggy (sgg) gene in LL, a behavior that is quite different from wild-
either of the two cell clusters. SGG is the Dro- type flies that become arrhythmic in LL. This is
sophila ortholog of the mammalian GSK3 that because CRY that is activated by light interferes
seems to phosphorylate TIM and by this way with TIM and provokes its degradation; TIM in
increase the speed of the clock (Martinek et al., turn is needed to stabilize its partner PER (Busza
2001). They found that the speed of M cells (but et al., 2004; Ceriani et al., 1999; Rosato et al.,
not that of the E cells) is directly reflected in the 2001). Thus, in wild-type flies, permanent degra-
free-running activity rhythm in DD. Further, the dation of TIM and PER in LL finally stops the
M cells influenced the rhythm of tim transcription molecular feedback loops of the clock (Myers
in the E cells, suggesting that the M cells are the et al., 1996). Sgg overexpression seems to block
dominating oscillators that entrain the E cells in CRY signaling, which in turn stabilizes TIM and
DD. However, the situation was different under PER oscillations and keeps the clock running
LL conditions (Stoleru et al., 2007): Whereas flies under LL (Stoleru et al., 2007). A similar
overexpressing sgg in the M cells (Pdf-gal4;uas- LL rhythmicity is observed if per or tim is
sgg) became arrhythmic similarly to wild-type overexpressed (Murad et al., 2007; Stoleru et al.,
flies, flies overexpressing sgg in E cells (tim-gal4; 2007) or if cry is mutated (Emery et al., 2000a).
Pdf-gal80;uas-sgg) remained rhythmic in LL. All these manipulations prevent the degradation
Thus, Stoleru and coworkers assumed that M of TIM (and PER) in response to light and keep
cells are master clocks in the dark and E cells in the clock running under LL conditions. However,
the light. Next, the activity rhythm of flies with if the clock is kept running only in subsets of the
sgg overexpression in either the M cells or the E clock neurons, this does not always lead to rhyth-
cells was tested under short (LD 10:14) and long micity in LL. Stoleru et al. (2007) could not see
photoperiods (LD 14:10) to see which oscillator any behavioral rhythms after they overexpressed
dominates at the two conditions. Indeed, under sgg in the M cells (¼ PDF neurons). This is
the short photoperiod, the fast sgg-overexpressing very interesting because it shows that the PDF
M cells set the phases of M and E peaks, whereas neurons alone cannot drive rhythmicity under
the fast clock in the E cells set the phases of both LL conditions.
67

The Rouyer and Emery labs (Murad et al., 2007; exists (Cyran et al., 2003). Thus, there is a group
Picot et al., 2007) looked in more detail into the of DN1p neurons that can also drive rhythmicity
ability of different clock neurons to drive behavioral in LL, when all other clock neurons are not work-
rhythmicity under LL. Whereas Murad et al. (2007) ing. Most probably these DN1ps are CRY nega-
took a similar approach as Stoleru et al. (2007) tive, because overexpression of per or sgg with
(overexpression of per and another gene morgue) cry-gal4 could not provoke rhythmic behavior in
to make clock neurons rhythmic in LL, Picot LL (Murad et al., 2007; Stoleru et al., 2007); but
et al. (2007) restricted CRY to subsets of clock at that time, it was not clear whether they belong
neurons. Picot and coworkers took either cry to the M or the E cells. In retro perspective, these
mutant flies and rescued cry function either in CRY-negative DN1ps likely belong to the E cells,
the M or in the E cells (using Pdf-gal4 or mai179- because recently the Emery lab identified a group
gal4;Pdf-gal80) or they knocked cry down, either of CRY-positive DN1ps that unequivocally belong
in the M or in the E cells (using the same drivers). to the M cells (Zhang et al., 2010a). These neurons
As a consequence, only neurons without CRY were detected with the Clk 4.1 M-gal4 line
were oscillating in LL. Similar to Stoleru et al. (Fig. 2c), and they were not able to drive rhythmic-
(2007), they found that a working clock in the ity in LL (after compromising CRY signaling)
M cells (¼ PDF neurons) alone is not enough (Zhang et al., 2010a). When per was rescued only
to drive behavioral rhythms in LL but that a work- in these neurons, the flies exhibited a robust M
ing clock in the E cells alone could do so. This peak anticipating the lights-on in LD 12:12 and
fact let Picot et al. (2007) conclude that the only a tiny E peak (at a certain temperature and
output from the M cells is disturbed in LL. On the light intensity; see later). When the period of these
other hand, the output from the E cells seems to DN1ps was shortened (by overexpressing a mutant
be disturbed in DD because these cells alone form of doubletime, dbt S; Preuss et al., 2004), the
could not drive behavioral rhythms in DD (Grima M peak but not the tiny E peak phase-advanced
et al., 2004; Picot et al., 2007; Rieger et al., 2009). showing convincingly that the M peak is driven
It is important to note that the definition of the E by the CRY-positive DN1p cells.
cells is different in the studies of Stoleru et al. In summary, the Drosophila clock seems to
(2007) and Picot et al. (2007). In the first paper, consist of M cells (the PDF-positive s-LNvs and
LL rhythmicity was found when all clock neurons CRY-positive DN1ps) that cannot drive rhythmic-
except the PDF neurons were working, whereas ity in LL when working alone and E cells (the 5th
Picot et al. could restrict the cells responsible s-LNv, three to four LNds, and perhaps the CRY-
for LL rhythmicity to the 5th s-LNv and three to negative DN1ps) that can drive rhythmicity in LL,
four LNd. but not in DD, when working alone.
Murad et al. (2007) overexpressed per in all
clock cells (with tim-gal4) or in all clock cells
except the PDF neurons (with tim-gal4; Pdf- Light accelerates the M cells and decelerates
gal80) and found the flies remaining rhythmic in the E cells
LL, very similar to Stoleru et al. (2007) after sgg
overexpression with the same drivers. However, According to the original dual oscillator model of
Murad et al. (2007) could trace the neurons Pittendrigh and Daan (1976), M and E oscillators
responsible for LL rhythmicity to a subset of the should respond to light by accelerating and
DN1ps and not to the 5th s-LNv and three to four decelerating their velocity, respectively. Indeed,
LNd. These DN1ps were the only neurons that a detailed behavioral analysis of cryb mutants under
showed rhythmic expression of PDP1, another LL of different light intensities revealed two activity
core clock protein for which an excellent antibody components free-running with short and long
68

Fig. 3. Activity rhythms in constant light (LL) of per0;;cryb flies in which per was rescued only in subsets of the clock neurons.
Representative actograms (doubleplots) are shown for each genotype. All flies were first entrained to an LD 12:12 and then
transferred into LL. The bar on top of each actogram shows the light–dark (LD) schedule during the LD period. (a) Actogram
of a fly with per only in 4 M cells—the PDF-positive s-LNvs—(genotype: per0;Pdf-gal4/þ;cryb uas-per16. Note that per is also
present in the l-LNvs, but that these are neither M nor E cells); all these flies were arrhythmic in LL. (b) Actograms of three
flies with per expression in four E cells —the 5th s-LNv and three LNds— (per0;mai179-gal4/þ;Pdf-gal80;cryb uas-per). Most of
these flies show three activity components with long period (trimodal activity), some flies showed two activity components
(bimodal activity), and only few a unimodal activity pattern. (c) Actogram of a fly with per in four M and four E cells—the
PDF-positive s-LNvs, the 5th s-LNv, and three LNds (per0; mai179-gal4/þ;uas-per). Typically, the flies were bimodal and free-ran
with long period, and sometimes a weak short-period component splits off the long one (see arrow) (d) Actograms of two flies
with per in the same four M and four E cells and additionally in three to six DN1p-M cells (per0;cry-gal4/þ;cryb uas-per). These
flies show internal desynchronization in a short- and long-period component, though their activity pattern is not very clear.
(e) Actogram of a fly with per in all clock neurons (per0;tim-gal4/þ;cryb uas-per). This fly shows a clear internal desynchronization
in a short-period component that originates from the M peak and a long-period component that originates from the E peak. This
activity pattern is typical for cryb and cry0 mutants (Dolezelova et al., 2007; Rieger et al., 2006; Yoshii et al., 2004).

periods (Rieger et al., 2006; Yoshii et al., 2004; see The reason why the two free-running components
Fig. 3). These independent components appeared were not found in other studies (e.g., Emery et al.,
as soon as light intensity exceeded 10 lux, whereby 2000a; Picot et al., 2007) seems due to the recorded
the short-period component became faster and period and the genetic background. For unknown
the long-period component became slower with reasons, the short-period component is often barely
increasing light intensity (Yoshii et al., 2004). This visible for first few cycles but becomes apparent
phenomenon was present in the great majority after the two components crossed each other for
of flies, and —at a given light intensity— the regular the first time and dissociate again, an event occur-
criss-cross of short- and long-period components ring  10–20 days after transfer to LL. Most
occurred rather simultaneously in every fly enabling researchers recorded the activity of the flies maxi-
molecular and histological analyses (see below). mally for 10–14 days. Due to the weakness of the
69

short activity component that will not be revealed by possible that they give rise to a short-period compo-
periodogram analysis of time series < 10 days, it nent that splits off the long one? To solve this ques-
remained undetected. Further, Yoshii et al. (2004) tion, we have repeated our LL experiments with
and Rieger et al. (2006) used cryb mutants with per0;cryb flies in which per was rescued in different
red eyes that showed the two free-running compo- subsets of M and E cells: (1) only in the PDF-
nents more clearly than white-eyed mutants (w;cryb positive LNvs, (2) in the 5th s-LNv and three LNds,
or y w;cryb) that have been used by the other (3) in most LNs, (4) in the LNs plus some DNs, or
researchers. Nevertheless, Dolezelova et al. (2007) (5) in all clock neurons (see exact genotypes in the
reported in the meantime the same phenomenon legend of Fig. 3). We found that the PDF-positive
for complete knock-out mutants, cry01, cry02, and LNvs alone are indeed not capable to produce a
cry03, indicating that the splitting rhythms are real short-period component but all flies are arrhythmic
and caused by clocks free-running in LL in the in LL (Fig. 3a; only 1 out of 30 flies showed initially
absence of CRY. a short-period component before it became arrhyth-
As mentioned, the short component is often mic). Thus, they are very similar to the flies that have
not visible after the flies have been transferred a functional clock only in the CRY-positive DN1p-
from LD to LL so that it is difficult to judge M-oscillators (Zhang et al., 2010a). The flies with
where it originates from. But in most cases it per in the 5th s-LNv and three LNds (“E-oscillator
seems rather to originate from the E peak than flies without DN1p-E-oscillators”) exhibited as
from the M peak, suggesting that the E peak is expected a long period, but the minority did show
controlled by M and E oscillators (see discussion the usual unimodal activity bout continuing from
in Helfrich-Förster et al., 2007b). Nevertheless, the evening peak (13%). Most flies exhibited a tri-
in some flies, the short period originates from (66%) or bimodal (21%) activity profile and the dif-
the M peak (see Fig. 3e). Thus, light seems to ferent components did not necessarily originate
shorten the period of the M oscillator and to from the E peak in LD (Fig. 3b). This strange activ-
lengthen that of the E oscillator, in good agree- ity pattern will be discussed in more detail later.
ment with the model of Pittendrigh and Daan Adding functional PDF-positive LNvs to the 5th
(1976). Rieger et al. (2006) demonstrated that s-LNv and three LNds (“M and E-oscillator flies
when the two activity components were out of without DN1p-oscillators”) consolidated the E activ-
phase on day 5 of LL, PER and TIM oscillations ity (Fig. 3c). The majority of flies showed bimodal
in the clock neurons were also out of phase. The (75%) or unimodal (20%) activity profiles, and
s-LNv,  3 LNd, and some DN1s ran in parallel the trimodal profiles almost disappeared (5%).
to the short-period component, whereas the 5th Again the free-running rhythm had a long period,
s-LNv,  3 LNd, and other DN1s ran in parallel and only in few cases (5%) was a second weak
to the long-period component. These results fit short-period component evident. But as soon as
almost perfectly to the classification of the clock the CRY-positive DN1ps were added to the per-
neurons in M and E cells by Stoleru et al. expressing cells (“M and E-oscillator flies with
(2004), Grima et al. (2004), Picot et al. (2007), DN1p-M-oscillators”), 43% of flies showed long
Murad et al. (2007), and Zhang et al. (2010a,b) and short free-running components in their activity
except for the fact that the LNds seemed to be pattern (Fig. 3d), although the two components
split up in M and E cells (for further discussion, were never as clear as in flies that have PER in all
see Helfrich-Förster et al., 2007b). clock neurons (Fig. 3e). A possible reason for the
Nevertheless, there remains one caveat. We have weakness of the short-period component could be
lined out above that neither the s-LNvs alone nor the fact that the cry-gal4 driver used to rescue per
the CRY-positive DN1ps alone can drive rhythmic in the DN1p-M-oscillators in addition to the LNv-
behavior under LL conditions. How it is then M-oscillators is not very strong. In the majority of
70

flies, only two to four of the six DN1p cells were PER M and E oscillators under moonlit nights and
positive. Nevertheless, the results indicate that both constant moonlight
M cell groups together are able to drive rhythmicity
under LL conditions and that they do so with a short Bachleitner et al. (2007) subjected wild-type flies
period. That both groups of M cells do indeed com- to day–night cycles in which the night was not
municate with each other was shown recently completely dark but illuminated by dim light
(Zhang et al., 2010b). (0.03 lux) corresponding to the intensity of quar-
In summary, it seems quite convincing that the ter moonlight. The flies responded strongly to
above defined M and E cells (1) give rise to M the moonlight and shifted M and E peak clearly
and E peaks under LD 12:12 conditions, (2) have into the night. The M peak advanced and the E
different capabilities to control behavioral peak delayed so that the two peaks had a larger
rhythms in DD and LL, and (3) respond to light CM,E than under dark nights. As a consequence,
by either accelerating or decelerating their speed. the flies become mostly nocturnal. There may be
These are exciting discoveries, as they give a first two reasons for the nocturnal behavior. First, flies
clue about the organization of the circadian clock seem to prefer generally dim light and do neither
in the brain that seems to be composed of two like to stay under high illumination nor in com-
principle types of clock cells with quite different plete darkness as revealed by a video monitoring
properties. Nevertheless, one has always to be system (Rieger et al., 2007). This behavior is inde-
aware of the fact that LL rhythmicity is pendent of a functional clock (Kempinger et al.,
completely artificial in flies. It can occur only 2009). Second, the experiment was performed in
when photoreception via CRY is impaired or the lab where the temperature remained during
TIM and PER are otherwise stabilized. Without day and night at pleasant 20  C, allowing the flies
such a manipulation, M and E neurons will not to become nocturnal. In nature, nocturnal activity
respond differently to LL conditions with light of fruit flies is not observed, even not in full-moon
intensities above  5 lux. Both types of clock nights (Bhutani, 2009; Vanin et al., 2012). Simi-
neurons would just stop their molecular larly, fruit flies do not become nocturnal under
oscillations. In wild-type flies, even normal light moonlit nights in the lab, when natural-like
during the day will lead to complete degradation T-cycles are simulated (Yoshii et al., 2010; see
of CRY and TIM: already before midday, no also below). Although the nocturnal behavior of
CRY and TIM are left in any of the clock neurons the flies seems not natural and partly caused by
(Miyasako et al., 2007; Shafer et al., 2002; Yoshii “masking” effects that do not depend on the clock,
et al., 2008). Thus, it makes no difference for the moonlight experiment showed that M and E
wild-type behavior that some E and M cells can peaks responded in the way that was proposed
principally drive rhythmicity under LL conditions by Pittendrigh and Daan: They shifted into oppo-
and others cannot, or that some cells are site directions—the M oscillator advanced upon
accelerated and others decelerated by LL—at least moonlight and the E oscillators delayed. Most
not during the highly illuminated day. To under- important, this was also true for the molecular
stand the real responses of M and E oscillators to PER and TIM oscillations in the above described
natural day–night cycles and finally the adaptation M and E cells (Bachleitner et al., 2007). This result
to different photoperiods, one should better not shows that M and E cells do respond in the
interfere with photoreception but study flies with expected way to dim light that is naturally present,
normal light sensitivity. Further, it may be wise to if not at night then at least during early dawn and
look at the effects of dim light that is naturally dusk in the morning and evening (see also below).
present during dawn and dusk as well as during The next question was what would happen to
star- and moonlight during the night. CM,E under constant moonlight. According to
71

Wild-type
LD ® DD LR ® RR

LD/LR
5

DD/RR
10

15

20

Fig. 4. Typical locomotor activity rhythms of wild-type flies that are subjected to different illuminations. In the left actogram, a fly
was entrained in LD 12:12 for 5 days and then was transferred into DD. The bimodal rhythm in LD does not persist in DD but
becomes unimodal (green line). In the right actogram, a fly was entrained in white-light/red-light (LR) cycles for 5 days before
constant red light (RR). This fly had a larger CM,E under LR, and M and E peaks persist in RR free-running in parallel to each
other (green lines).

Pittendrigh and Daan (1976) and Aschoff (1966), started to free-run with short and long period,
CM,E should be larger under dim light than under respectively, before the flies became arrhythmic
darkness. In DD, CM,E usually gets very small in (Rieger et al., 2006). All this is in accordance with
fruit flies and sometimes M and E peaks are not the dual oscillator model.
anymore distinguishable as separate peaks (see A next logical step is to investigate how flies
Fig. 4a). Under constant moonlight, M and E with only functional M or E cells respond to dim
peaks were clearly distinguishable in each single light in the night or to constant dim light. This
fly, CM,E was  11 h, in contrast to  8 h in DD has been investigated by Rieger et al. (2009). Let
(Bachleitner et al., 2007). The same effect can be us start with the constant dim light experiments
seen when flies are subjected to constant colored because these are largely in line with the LL exper-
light of longer wavelength (> 560 nm, Fig. 4b). In iment conducted in flies with compromised CRY
this case, irradiance can even be increased to very signaling (see above). LL of 0.03 lux was already
high values without the risk that the flies become too intense for the “LNv-M-oscillator-only flies”
arrhythmic (Yoshii, Roidl, Wülbeck, and Char- (per only in the PDF-positive LNvs) (Rieger
lotte Helfrich-Förster, unpublished results). This et al., 2009). These became arrhythmic as did the
is because CRY is not activated by wavelengths same flies with compromised CRY signaling under
above 540 nm (Berndt et al., 2007; Helfrich- LL of higher light intensity (Picot et al., 2007). In
Förster et al., 2002; VanVickle-Chavez and Van contrast, the “E-oscillator-only flies” (per only in
Gelder, 2007). We are currently using constant the 5th s-LNv and three LNds) remained rhythmic;
colored light to test the effects of light on the and again this is in concert with the previously
two oscillators in more detail. But even low white mentioned results and fits to the model (Grima
light proved as useful: when light intensity was et al., 2004; Murad et al., 2007; Picot et al., 2007;
increased to only 0.5 lux, M and E components Stoleru et al., 2007).
72

Nevertheless, not every observation under simulated dawn and dusk would extend CM,E to a
moonlight was along the expectations. Instead of larger extent than long photoperiods with rectangu-
exhibiting only an E activity bout under light– lar LD-cycles. This was indeed the case for all three
moonlight (LM) cycles, the “E-oscillator-only flies” tested wild-type strains; they extended CM,E by
showed an additional prominent activity peak in 20–60 min, depending on the genetic background
the second half of the night and this activity peak (Rieger et al., 2012).
persisted under constant moonlight (Rieger et al.,
2009). As a consequence, the activity pattern of
the “E-oscillator-only flies” was clearly bimodal Adaptation of the clock to different photoperiods
and not unimodal as expected. Although the nature occurs via light input through the photoreceptor
of this activity peak is not completely clear —it organs and not via CRY
might be a very early M peak as discussed in
Helfrich-Förster et al. (2007b)— this result CRY was already mentioned as a circadian photo-
indicates that the fly clock is highly plastic and receptor that is expressed in several clock neurons
not the same under all conditions. At the tested and that interacts directly with TIM leading to its
LM-cycles, immunohistochemical analysis revealed degradation upon light. Although CRY action on
that two LNds behaved rather as M cells and the the clock is a direct and extremely efficient way
5th s-LNv and one LNd behaved likely as E cells to reset the clock by light, CRY is not involved in
(Rieger et al., 2009). Thus, the composition in M the adaptation of M and E peaks to different
and E cells may be variable and dependent on photoperiods: mutants without functional CRY
the environmental conditions. We will come to this can perfectly adapt their activity to short and long
important aspect later. photoperiod, but eyeless mutants cannot (Rieger
et al., 2003). Eyeless mutants are also not able to
shift their activity into moonlit nights (Bachleitner
Flies adapting to different photoperiods et al., 2007), suggesting that the dim light receptor
must reside in the eyes and probably in an extra-
Simulation of dawn and dusk retinal eyelet, called Hofbauer–Buchner (H–B)
eyelet, that derives from the larval photoreceptor,
Let us come back to the behavior of wild-type flies. Bolwig’s Organ (Helfrich-Förster et al., 2002;
So far this was only regarded under artificial LD- Veleri et al., 2007). Bolwig’s Organ and the H–B
cycles, in which the lights were switched on in the eyelet directly target the LNv cells (Helfrich-
morning and switched off in the evening. As the Förster et al., 2002; Malpel et al., 2002). The H–B
fly clock is very light sensitive and strongly responds eyelet and certain photoreceptor neurons in the
to moonlight, it is very likely that dim light during eye express Rhodopsin 6 that still shows sensitivity
dawn and dusk will affect M and E oscillators, under longer wavelengths and could be responsi-
respectively. Rieger et al. (2012) compared the syn- ble for the red light effects on M and E activity
chronization of three different wild-type strains peaks (Hanai et al., 2008; Fig. 4).
under conventional rectangular LD-cycles and
LD-cycles with simulated dawn and dusk at differ-
ent photoperiods. As mentioned in the introduc- The PDF-positive l-LNvs play a crucial role in
tion, M and E activity peaks of fruit flies largely mediating light input from the eyes
follow lights-on and lights-off, and as a conse-
quence, CM,E increases with increasing photoperiod. The l-LNvs are among the cells that may be directly
If M and E oscillators are mostly set by dim light contacted by the H–B eyelet and some photorecep-
(and not by bright light that stops the oscillations), tor cells of the compound eyes (Helfrich-Förster
one would expect that long photoperiods with et al., 2002). Several studies suggest that these
73

clock neurons are neither M nor E cells but play a of the per and tim genes seem to account for the
crucial role in transferring light signals to the clock. temperature effects on the phase of the E peak
The l-LNvs respond to light with increased sponta- (Boothroyd et al., 2007; Collins et al., 2004;
neous action potential firing (Sheeba et al., 2008a) Majercak et al., 1999, 2004), but no molecular
and they modulate the light arousal and sleep mechanism is known that shifts the M peak in
of Drosophila (Chung et al., 2009; Parisky et al., the opposite direction of the E peak.
2008; Shang et al., 2008; Sheeba et al., 2008c). It is also unknown how the overall temperature
Hyperexcitation of the l-LNvs increases the noctur- level will influence the activity of flies that possess
nal activity of the flies (Parisky et al., 2008) as does only M or E cells, but some data are available
dim light during the night (see above). The arousal for such flies under T-cycles. Artificial rectangular
effects of the l-LNvs are clearly mediated by PDF T-cycles in the lab (temperature immediately
because Pdf 01 mutants do not increase nocturnal increases in the morning and decreases in the even-
activity, neither after hyperexcitation of the l-LNvs ing in DD) can synchronize the activity rhythm of
(Sheeba et al., 2008b) nor after exposition to dim flies, whereby these exhibit the usual M and E
light during the night (Helfrich-Förster, 2009). activity peaks (Busza et al., 2007; Tomioka et al.,
PDF from the l-LNvs can also synchronize the 1998; Wheeler et al., 1993; Yoshii et al., 2002).
LNds and DNs to LD-cycles in the absence of Under the same conditions, per0 mutants do not
CRY (Cusumano et al., 2009). Most interestingly, show M and E peaks but are just more active dur-
an increase of PDF in the dorsal brain does lead ing the higher temperature. All this resembles the
to internal desynchronization of the free-running activity pattern observed under LD conditions.
activity rhythm into two components as does LL However, different neurons seem to be responsible
(Helfrich-Förster et al., 2000; Wülbeck et al., for M and E activity peaks under T-cycles: When
2008), and the same neurons seem to free-run with per is rescued in the PDF-positive M cells, an activ-
short and long period, respectively, as observed ity peak occurs in the middle of the warm phase
under LL (Yoshii et al., 2009a). This suggests that that seems to be neither an M nor an E peak (Busza
PDF is the factor that accelerates the speed of the et al., 2007). Even the per rescue in all CRY-
M cells and decelerates the speed of the E cells. positive neurons (with cry-gal4) did not lead to
In accordance with this, neither did Pdf01 mutants wild-type-like M and E peaks. This suggests that
show internal desynchronization upon light, nor the CRY positive are dispensable for the normal
are they able to adapt M and E peaks to different control of M and E peaks under T-cycles and that
photoperiods (Yoshii et al., 2009a; see also discus- the CRY negative are more important. Indeed,
sion in Helfrich-Förster, 2006, 2009). flies that lack the CRY-positive clock neurons still
synchronize to T-cycles (Busza et al., 2007; Yoshii
et al., 2010). Among the CRY-negative clock
The effects of temperature on M and E oscillators neurons, the LPNs seem to be most important
(Busza et al., 2007; Yoshii et al., 2005): The PER/
As mentioned in the introduction, the environ- TIM oscillations in the LPNs do only synchronize
mental temperature is extremely important for to T-cycles but not to LD-cycles. The DN2 cells
poikilothermic animals, as it directly influences also seem to be more sensitive to T-cycles than to
their activity, reproduction, and survival (Garrity LD-cycles (Miyasako et al., 2007; Picot et al.,
et al., 2010). Majercak et al. (1999) nicely demon- 2009; Yoshii et al., 2009b). In contrast, the PDF-
strated that the overall temperature strongly positive l-LNvs do only entrain to LD-cycles, but
affects the phases of M and E peaks in Drosophila not at all to T-cycles (Yoshii et al., 2009b). This
in the same way as the photoperiod: The M peak result strongly underlines the above-mentioned
advances and the E peak delays with increasing role of the latter cells (and PDF) in the light-input
temperature (see Fig. 1). Thermosensitive splicing pathway.
74

Flies under natural-like temperature cycles —a fact that was reported already a long time ago
(Pittendrigh et al., 1958)— and (2) it underlines
In nature, T-cycles are not rectangular, but tem- the above-mentioned importance of the CRY-
perature slowly increases in the morning until it positive neurons for entrainment to light.
reaches its maximum  2 h after midday, and then But what happens when the PDF-positive l-LNvs
it decreases again reaching the lowest point in the that seem to be most important for light input are
morning at sunset (Bhutani, 2009; Yoshii et al., ablated? Busza et al. (2007) performed this impor-
2009b). When such natural-like T-cycles are tant experiment and found that such flies entrain
simulated in the lab, the flies show their M peak faster to T-cycles than wild-type flies. Busza et al.
few hours after the temperature minimum and concluded that the PDF cells may even inhibit the
the temperature maximum, respectively (Currie temperature entrainment of CRY-negative neurons
et al., 2009; Yoshii et al., 2009b). Recently, we and that this inhibition is important for entrainment
subjected flies to simulated short and long days in nature because it guarantees that the clock does
of natural T-cycles of 26/16  C under DD not overreact to erratic temperature changes caused
conditions (Bywalez et al., 2012). The two peaks by weather changes.
nicely changed their phase-angle to adapt to dif- More complex results are found when per ex-
ferent day lengths as was observed under differ- pression is restricted to the CRY-positive DN1ps
ent photoperiods: CM,E was shorter in shorter that are designated as M cells under LD conditions
days and broader in longer days. This suggests (Zhang et al., 2010a; see above). Such flies show a
that CM,E is similarly flexible under T-cycles as prominent M peak under LD 12:12 (500 lux) and a
it is under LD-cycles. Thus, the dual oscillator constant temperature of 25  C. However, when
model seems to be valid also for T-cycles, light intensity was reduced to 50 lux or the temper-
although different clock neurons may be responsi- ature decreased to 20  C, an additional E peak
ble for M and E activity under LD- and T-cycles. appeared. The same happened when the flies were
Which CRY-negative clock neurons control M subjected to T-cycles in DD. All these suggest that
activity and which control E activity under T- the temperature interacts with light in a complex
cycles has to be determined in the future. way. The most important result of the study of
Zhang et al. (2010a) is that the CRY-positive
DN1ps are sufficient to generate M and E peaks
Interaction of light and temperature but that the expression of the two peaks depends
on the environment. A possible explanation is that
When natural-like LD- and T-cycles are applied in the E peak of the flies may be suppressed at higher
phase with each other, they work synergistically on light intensities and temperatures, but this has to
entrainment: the flies show narrow M and E peaks be proven by future studies.
and the molecular clock cycles with high amplitude A completely different approach to dissect the
in all clock neurons (Yoshii et al., 2009b). Interest- temperature-sensitive clock is to make the light-
ing effects are observed when LD- and T-cycles sensitive clock arrhythmic by applying LL of rather
are applied out of phase with each other or when high light intensity (Tomioka et al., 1998; Yoshii
the clock is restricted to certain clock neurons et al., 2002). Amazingly, T-cycles can override the
(Busza et al., 2007; Yoshii et al., 2010; Zhang LL effect on the clock, although the amplitudes of
et al., 2010a). PER and TIM oscillations under such conditions
Under conflicting LD- and T-cycles, wild-type are quite low (Glaser and Stanewsky, 2005; Yoshii
flies entrain to the LD-cycles, but CRY-less flies to et al., 2005). Nevertheless, the low-amplitude
the T-cycles (Yoshii et al., 2010). This shows that molecular rhythms restored by T-cycles in LL are
(1) light is a stronger Zeitgeber than temperature sufficient for robust behavioral rhythms (Yoshii
75

(a) (b)
DD T-cycles LL T-cycles
26 °C/16 °C 26 °C/16 °C

2
n=30 n=32
2

Activity
1
1

0 0
18 0 6 12 18 18 0 6 12 18
Zeitgeber time (hr) Zeitgeber time (hr)

(c) (d)
(n=27) (n=28)

5 5

6 hr
10 10 delay
Days

15 15

6 hr
20 20 advance

25 25

Fig. 5. Activity rhythms under natural-like temperature cycles (T-cycles) in DD (a, c) and LL (b, d). (a) The averaged activity
pattern in DD T-cycles indicates that one main activity peak occurs in the evening and another peak occurs in the morning
(arrowheads; Yoshii et al., 2009b). Those two peaks run in parallel when T-cycles were shifted by 6 h delay and advance (c). In
contrast, only one robust peak at evening is visible in LL T-cycles, and the phase of E peak is delayed compared with that in
DD T-cycles (b, d). (c, d) Average actograms from 27 flies for DD T-cycles and from 28 flies for LL T-cycles, respectively.

et al., 2002; 2007). Interestingly, the restored behav- The dual oscillator model appears too simple
ioral rhythms in LL do not display an M peak except
for a startle response at the sudden temperature Though the M and E oscillator model turned out
increase in the morning (these studies were per- to be very useful to explain the adaptation of flies
formed under rectangular T-cycles). More informa- to long and short days, it became clear that it is
tive are natural-like T-cycles that show clearly that too simple to understand all observations made
the M peak is absent in LL but is present in DD so far. Several studies show that there are more
conditions (Fig. 5). In contrast, the E peak is more than two clocks in the fly brain. These should
robust under T-cycles in LL than in DD. These be briefly summarized in the following.
findings fit to the idea mentioned in the beginning
that the M oscillator runs in DD and the E oscillator
in LL (Picot et al., 2007; Stoleru et al., 2007). Never- E cells alone can drive two or even more
theless, the relationship between light and tempera- activity components
ture is clearly more complicated. The phases of
the E peak are different in DD and LL (Fig. 5), Sheeba et al. (2010) ablated specifically the PDF-
and the above-mentioned results for flies with positive s-LNvs by expressing the human neurotoxic
restricted clock function (especially the ones with protein Q128-Huntingtin. One may expect that
the DN1p-only flies) clearly show that we do not such flies lack the M peak, but this was clearly not
yet completely understand how light and tempera- the case: M and E peaks remained present. In this
ture act on M and E oscillators. study, still the CRY-positive DN1ps may account
76

for the remaining M peak, but this cannot be the People may think that only 150 clock neurons of
explanation for the two activity peaks present in Drosophila should be easy to study. Yes, perhaps
per0 flies in which per was rescued only in the four they are easier to study than the 10,000 neurons in
LNs (the 5th s-LNv and three LNds; Rieger et al., the mammalian clock, but still more complicated
2009). Most convincing that these four putative E than we wished. Nevertheless, we should not capit-
cells can drive several activity components are per- ulate. To reach the goal, we need much more
haps the results shown in Fig. 3b. Here, the same pieces in the puzzle. People are successfully
flies were put additionally into the cryb background searching for further useful gal4 drivers and devel-
and subjected to LL. Instead of the expected single oping new efficient methods to manipulate subsets
activity band, up to three free-running components of neurons (Choi et al., 2009; Gummadova et al.,
were present and these did not even originate from 2009; Sheeba et al., 2010; Wu et al., 2008). The bac-
the E peak in LD (Fig. 3b). Rieger et al. (2009) terial LexA-LexAop-system and the Q-system of
suggested that not all of these neurons are E cells, Neurospora are new binary expression methods
but alternatively, the nature of the neurons could that can be combined with the gal4-uas/gal80 sys-
be influenced by the environment (here the light tem allowing to dissect the clock network further
intensity). (Lai and Lee, 2006; Potter and Luo, 2011; Shang
et al., 2008). Although we can only expect minor
effects on the flies’ activity rhythm of the flies after
Under certain circumstances, M cells alone can
manipulating smaller subsets of clock neurons, one
also drive two activity components
day these may represent extremely useful pieces
for the puzzle we need to solve in order to under-
As outlined above, even the M cells alone —in this
stand the whole system. We would like to give a
case, the CRY-positive DN1ps— can drive M and
recent example: We have just ablated the NPF-
E peaks. This is the case at high light intensities
positive clock cells, which are three LNd, the 5th
and higher temperatures (Zhang et al., 2010a).
s-LNv, and few l-LNv, and found subtle but signifi-
Either these six or seven cells represent not exclu-
cant opposing effects on the E activity peak and
sively M oscillators as was suggested for the four
the free-running period in DD (Hermann et al.,
E cells above, or again the environment (here tem-
2012) that have not been reported before (Lee
perature and light intensity) determines whether
et al., 2006): the knock-out advanced the E peak
they behave as M or E cells.
but prolonged the free-running period—an effect
All these results suggest that we are dealing with a
that is hard to reconcile with the present M and E
flexible multi-oscillator system that may be com-
model. Clearly, more pieces of the puzzle are
posed of interacting M and E cells (see also
needed to understand what this means.
Umezaki and Tomioka, 2008). Alternatively, the
Tiny differences after defined manipulations
activity pattern may be determined by a flexible net-
ask also for careful observations and analyzing
work of interacting clock neurons without the neces-
methods. Defining the presence of M and E
sity of the designated M and E cells. The latter view
peaks by looking at anticipatory activity before
is strongly favored by Sheeba et al. (2008b).
lights-on and lights-off is probably not enough.
An activity peak can be delayed or masked by
Welcome to complexity certain circumstances, but this does not mean
that the peak is absent (see Helfrich-Förster,
Understanding nature and biological mechanisms 2009 for further discussion).
is hard. Seven years after the Stoleru and Grima It is also important to choose the best experi-
papers, we have to realize that it is very difficult mental conditions for studying M and E
to define M and E cells in the clock network. oscillators. The dual oscillator system is important
77

to adapt the activity rhythm to different E evening


photoperiods and temperatures, meaning that LD light–dark
investigations under different photoperiods and LL constant light
temperatures are warranted (Bywalez et al., LN lateral neuron
2012; Hermann et al., 2012; Majercak et al., M morning
1999; Rieger et al., 2003, 2012). Monitoring phase PDF pigment-dispersing factor
response curves for each oscillator to light pulses PER period
is also an important issue as was suggested by T-cycles temperature cycles
Schwartz (2004) in his comments on the Stoleru TIM timeless
and Grima papers. Given the predictions by Daan CM,E the phase-angle between M and E
et al. (2001), the M clock should only show phase- peak
advances upon light pulses, whereas the E clock
should show only phase-delays. Such fundamental
investigations are still lacking. References
Stoleru and Grima and the following papers
provided a good start how to find clock neurons Aschoff, J. (1966). Circadian activity pattern with two peaks.
in the fly brain that control different aspects of Ecology, 47, 657–662.
Bachleitner, W., Kempinger, L., Wülbeck, C., Rieger, D., &
rhythmic behavior. Now we have to continue this Helfrich-Förster, C. (2007). Moonlight shifts the endoge-
research—a task that is possible due to the unique nous clock of Drosophila melanogaster. Proceedings of the
genetic methods available in the fruit fly. Under- National Academy of Sciences of the United States of
standing the fly’s dual or multiple oscillators and America, 104, 3538–3543.
Benito, J., Houl, J. H., Roman, G. W., & Hardin, P. E. (2008).
their role in seasonal adaptation will not be easy,
The blue-light photoreceptor CRYPTOCHROME is
but will certainly remain fascinating. expressed in a subset of circadian oscillator neurons in the
Drosophila CNS. Journal of Biological Rhythms, 23, 296–307.
Berndt, A., Kottke, T., Breitkreuz, H., Dvorsky, R.,
Acknowledgments Hennig, S., Alexander, M., et al. (2007). A novel photoreac-
tion mechanism for the circadian blue light photoreceptor
Drosophila cryptochrome. The Journal of Biological Chem-
We thank Pamela Menegazzi, Francois Rouyer, istry, 282, 13011–13021.
and Irina Stahl for commenting on and editing Bhutani, S. (2009). Natural entrainment of the Drosophila
the chapter and Francois Rouyer for providing melanogaster circadian clock. PhD Thesis, University of
Leicester.
the fly strains used in the LL experiment shown
Blanchardon, E., Grima, B., Klarsfeld, A., Chelot, E.,
in Fig. 3. The work from our groups was finan- Hardin, P. E., Preat, T., et al. (2001). Defining the role of
cially supported by the European Community Drosophila lateral neurons in the control of circadian
(the 6th Framework Project EUCLOCK no. rhythms in motor activity and eclosion by targeted genetic
018741) and that of the C.H.F. group additionally ablation and PERIOD protein overexpression. The Euro-
by the DFG (Fo207/10). pean Journal of Neuroscience, 13, 871–888.
Blau, J., & Young, M. W. (1999). Cycling vrille expression is
required for a functional Drosophila clock. Cell, 99, 661–671.
Boothroyd, C. E., Wijnen, H., Naef, F., Saez, L., & Young, M. W.
Abbreviations (2007). Integration of light and temperature in the regulation of
circadian gene expression in Drosophila. PLoS Genetics, 3, e54.
CLK clock Bradshaw, W. E., & Holzapfel, C. M. (2010). What season is it
CRY cryptochrome anyway? Circadian tracking vs. photoperiodic anticipation
in insects. Journal of Biological Rhythms, 25, 155–165.
CYC cycle Brand, A. H., & Perrimon, N. (1993). Targeted gene expres-
DD constant dark sion as a means of altering cell fates and generating domi-
DN dorsal neuron nant phenotypes. Development, 118, 401–415.
78

Brown, T. M., & Piggins, H. D. (2009). Spatiotemporal hetero- Dibner, C., Schibler, U., & Albrecht, U. (2010). The mamma-
geneity in the electrical activity of suprachiasmatic nuclei lian circadian timing system: Organization and coordination
neurons and their response to photoperiod. Journal of of central and peripheral clocks. Annual Review of Physiol-
Biological Rhythms, 24, 44–54. ogy, 72, 517–549.
Bünning, E. (1936). Die endonome Tagesrhythmik als Dolezelova, E., Dolezel, D., & Hall, J. C. (2007). Rhythm
Grundlage der photoperiodischen Reaktion. Ber d Deutsch defects caused by newly engineered null mutations in
Bot Ges, 54, 590–607. Drosophila’s cryptochrome gene. Genetics, 177, 329–345.
Burgoon, P. W., Lindberg, P. T., & Gillette, M. U. (2004). Dif- Dubruille, R., & Emery, P. (2008). A plastic clock: How circa-
ferent patterns of circadian oscillation in the suprachiasmatic dian rhythms respond to environmental cues in Drosophila.
nucleus of hamster, mouse, and rat. Journal of Comparative Molecular Neurobiology, 38, 129–145.
Physiology. A, Neuroethology, Sensory, Neural, and Behav- Duffy, J. B. (2002). GAL4 system in Drosophila: A fly
ioral Physiology, 190, 167–171. geneticist’s Swiss army knife. Genesis, 34, 1–15.
Busza, A., Emery-Le, M., Rosbash, M., & Emery, P. (2004). Dushay, M. S., Rosbash, M., & Hall, J. C. (1989). The discon-
Roles of the two Drosophila CRYPTOCHROME structural nected visual system mutations in Drosophila melanogaster
domains in circadian photoreception. Science, 304, 1503–1506. drastically disrupt circadian rhythms. Journal of Biological
Busza, A., Murad, A., & Emery, P. (2007). Interactions between Rhythms, 4, 1–27.
circadian neurons control temperature synchronization of Emery, P., Stanewsky, R., Hall, J. C., & Rosbash, M. (2000).
Drosophila behavior. The Journal of Neuroscience, 27, Drosophila cryptochromes: A unique circadian-rhythm pho-
10722–10733. toreceptor. Nature, 404, 456–457.
Bywalez, W., Menegazzi, P., Rieger, D., Schmid, B., Helfrich- Emery, P., Stanewsky, R., Helfrich-Förster, C., Emery-Le, M.,
Förster, C., & Yoshii, T. (2012). The dual oscillator system Hall, J. C., & Rosbash, M. (2000). Drosophila CRY is a
of Drosophila melanogaster under natural-like temperature deep brain circadian photoreceptor. Neuron, 26, 493–504.
cycles. Chronobiology International, 29, 395–407. Garrity, P. A., Goodman, M. B., Samuel, A. D., &
Ceriani, M. F., Darlington, T. K., Staknis, D., Mas, P., Sengupta, P. (2010). Running hot and cold: Behavioral
Petti, A. A., Weitz, C. J., et al. (1999). Light-dependent strategies, neural circuits, and the molecular machinery for
sequestration of TIMELESS by CRYPTOCHROME. thermotaxis in C. elegans and Drosophila. Genes & Develop-
Science, 285, 553–556. ment, 24, 2365–2382.
Choi, C., Fortin, J. P., McCarthy, E., Oksman, L., Kopin, A. S., Glaser, F. T., & Stanewsky, R. (2005). Temperature synchroni-
& Nitabach, M. N. (2009). Cellular dissection of circadian zation of the Drosophila circadian clock. Current Biology,
peptide signals with genetically encoded membrane-tethered 15, 1352–1363.
ligands. Current Biology, 19, 1167–1175. Glossop, N. R., Houl, J. H., Zheng, H., Ng, F. S., Dudek, S. M.,
Chung, B. Y., Kilman, V. L., Keath, J. R., Pitman, J. L., & & Hardin, P. E. (2003). VRILLE feeds back to control cir-
Allada, R. (2009). The GABA(A) receptor RDL acts in cadian transcription of Clock in the Drosophila circadian
peptidergic PDF neurons to promote sleep in Drosophila. oscillator. Neuron, 37, 249–261.
Current Biology, 19, 386–390. Grima, B., Chelot, E., Xia, R., & Rouyer, F. (2004). Morning
Collins, B. H., Rosato, E., & Kyriacou, C. P. (2004). Seasonal and evening peaks of activity rely on different clock neurons
behavior in Drosophila melanogaster requires the photo- of the Drosophila brain. Nature, 431, 869–873.
receptors, the circadian clock, and phospholipase C. Gummadova, J. O., Coutts, G. A., & Glossop, N. R. (2009).
Proceedings of the National Academy of Sciences of the Analysis of the Drosophila Clock promoter reveals hetero-
United States of America, 101, 1945–1950. geneity in expression between subgroups of central oscilla-
Currie, J., Goda, T., & Wijnen, H. (2009). Selective entrain- tor cells and identifies a novel enhancer region. Journal of
ment of the Drosophila circadian clock to daily gradients Biological Rhythms, 24, 353–367.
in environmental temperature. BMC Biology, 7, 49. Hanai, S., Hamasaka, Y., & Ishida, N. (2008). Circadian
Cusumano, P., Klarsfeld, A., Chelot, E., Picot, M., Richier, B., entrainment to red light in Drosophila: Requirement of
& Rouyer, F. (2009). PDF-modulated visual inputs and Rhodopsin 1 and Rhodopsin 6. Neuroreport, 19, 1441–1444.
cryptochrome define diurnal behavior in Drosophila. Nature Hao, Z., Ng, F., Yixiao, L., & Hardin, P. E. (2008). Spatial and
Neuroscience, 12, 1431–1437. circadian regulation of cry in Drosophila. Journal of
Cyran, S. A., Buchsbaum, A. M., Reddy, K. L., Lin, M. C., Biological Rhythms, 23, 283–295.
Glossop, N. R., Hardin, P. E., et al. (2003). vrille, Pdp1, Hardin, P. E., Hall, J. C., & Rosbash, M. (1992). Behavioral
and dClock form a second feedback loop in the Drosophila and molecular analyses suggest that circadian output is dis-
circadian clock. Cell, 112, 329–341. rupted by disconnected mutants in D. melanogaster. EMBO
Daan, S., Albrecht, U., van der Horst, G. T., Illnerova, H., Journal, 11, 1–6.
Roenneberg, T., Wehr, T. A., et al. (2001). Assembling a Hazlerigg, D. G., Ebling, F. J., & Johnston, J. D. (2005). Photo-
clock for all seasons: Are there M and E oscillators in the period differentially regulates gene expression rhythms in the
genes? Journal of Biological Rhythms, 16, 105–116. rostral and caudal SCN. Current Biology, 15, R449–R450.
79

Helfrich-Förster, C. (1995). The period clock gene is expressed suprachiasmatic nucleus couple photoperiodically to the onset
in central nervous system neurons which also produce a neu- and end of daily activity. Proceedings of the National Academy
ropeptide that reveals the projections of circadian pace- of Sciences of the United States of America, 104, 7664–7669.
maker cells within the brain of Drosophila melanogaster. Jagota, A., de la Iglesia, H. O., & Schwartz, W. J. (2000). Morn-
Proceedings of the National Academy of Sciences of the ing and evening circadian oscillations in the suprachiasmatic
United States of America, 92, 612–616. nucleus in vitro. Nature Neuroscience, 3, 372–376.
Helfrich-Förster, C. (1998). Robust circadian rhythmicity of Johard, H. A., Yoshii, T., Dircksen, H., Cusumano, P.,
Drosophila melanogaster requires the presence of lateral Rouyer, F., Helfrich-Förster, C., et al. (2009). Peptidergic
neurons: A brain-behavioral study of disconnected mutants. clock neurons in Drosophila: Ion transport peptide and
Journal of Comparative Physiology. A, 182, 435–453. short neuropeptide F in subsets of dorsal and ventral lat-
Helfrich-Förster, C. (2000). Differential control of morning eral neurons. The Journal of Comparative Neurology, 516,
and evening components in the activity rhythm of Dro- 59–73.
sophila melanogaster—Sex-specific differences suggest a Kadener, S., Stoleru, D., McDonald, M., Nawathean, P., &
different quality of activity. Journal of Biological Rhythms, Rosbash, M. (2007). Clockwork Orange is a transcriptional
15, 135–154. repressor and a new Drosophila circadian pacemaker com-
Helfrich-Förster, C. (2006). The neural basis of Drosophila’s ponent. Genes & Development, 21, 1675–1686.
circadian clock. Sleep and Biological Rhythms, 4, 224–234. Kaneko, M., & Hall, J. C. (2000). Neuroanatomy of cells
Helfrich-Förster, C. (2009). Does the morning and evening expressing clock genes in Drosophila: Transgenic manipu-
oscillator model fit better for flies or mice? Journal of lation of the period and timeless genes to mark the peri-
Biological Rhythms, 24, 259–270. karya of circadian pacemaker neurons and their
Helfrich-Förster, C., Edwards, T., Yasuyama, K., Wisotzki, B., projections. The Journal of Comparative Neurology, 422,
Schneuwly, S., Stanewsky, R., et al. (2002). The extraretinal 66–94.
eyelet of Drosophila: Development, ultrastructure, and Kaneko, M., Park, J. H., Cheng, Y., Hardin, P. E., & Hall, J. C.
putative circadian function. The Journal of Neuroscience, (2000). Disruption of synaptic transmission or clock-
22, 9255–9266. gene-product oscillations in circadian pacemaker cells of
Helfrich-Förster, C., Shafer, O. T., Wülbeck, C., Drosophila cause abnormal behavioral rhythms. Journal of
Grieshaber, E., Rieger, D., & Taghert, P. (2007). Develop- Neurobiology, 43, 207–233.
ment and morphology of the clock-gene-expressing lateral Kempinger, L., Dittmann, R., Rieger, D., & Helfrich-Förster, C.
neurons of Drosophila melanogaster. The Journal of Com- (2009). The nocturnal activity of fruit flies exposed to artifi-
parative Neurology, 500, 47–70. cial moonlight is partly caused by direct light effects on the
Helfrich-Förster, C., Tauber, M., Park, J. H., Muhlig- activity level that bypass the endogenous clock. Chronobiol-
Versen, M., Schneuwly, S., & Hofbauer, A. (2000). Ectopic ogy International, 26, 151–166.
expression of the neuropeptide pigment-dispersing factor Klarsfeld, A., Malpel, S., Michard-Vanhee, C., Picot, M.,
alters behavioral rhythms in Drosophila melanogaster. The Chelot, E., & Rouyer, F. (2004). Novel features of
Journal of Neuroscience, 20, 3339–3353. cryptochrome-mediated photoreception in the brain circa-
Helfrich-Förster, C., Winter, C., Hofbauer, A., Hall, J. C., & dian clock of Drosophila. The Journal of Neuroscience, 24,
Stanewsky, R. (2001). The circadian clock of fruit flies is 1468–1477.
blind after elimination of all known photoreceptors. Neuron, Lai, S. L., & Lee, T. (2006). Genetic mosaic with dual binary
30, 249–261. transcriptional systems in Drosophila. Nat Neurosci, 9,
Helfrich-Förster, C., Yoshii, T., Wülbeck, C., Grieshaber, E., 703–709.
Rieger, D., Bachleitner, W., et al. (2007). The lateral and Lamaze, A., Lamouroux, A., Vias, C., Hung, H. C., Weber, F.,
dorsal neurons of Drosophila melanogaster: New insights & Rouyer, F. (2011). The E3 ubiquitin ligase CTRIP con-
about their morphology and function. Cold Spring Harbor trols CLOCK levels and PERIOD oscillations in Drosoph-
Symposia on Quantitative Biology, 72, 517–525. ila. EMBO Reports, 12, 549–557.
Hermann, C., Yoshii, T., Dusik, V., & Helfrich-Förster, C. Lee, G., Bahn, J. H., & Park, J. H. (2006). Sex- and clock-
(2012). Neuropeptide F immunoreactive clock neurons controlled expression of the neuropeptide F gene in
modify evening locomotor activity and free-running period Drosophila. Proceedings of the National Academy of
in Drosophila melanogaster. The Journal of Comparative Sciences of the United States of America, 103, 12580–12585.
Neurology, 520, 970–987. Lee, T., & Luo, L. (1999). Mosaic analysis with a repressible
Houl, J. H., Yu, W., Dudek, S. M., & Hardin, P. E. (2006). cell marker for studies of gene function in neuronal morpho-
Drosophila CLOCK is constitutively expressed in circadian genesis. Neuron, 22, 451–461.
oscillator and non-oscillator cells. Journal of Biological Majercak, J., Chen, W. F., & Edery, I. (2004). Splicing of the
Rhythms, 21, 93–103. period gene 3’-terminal intron is regulated by light, circadian
Inagaki, N., Honma, S., Ono, D., Tanahashi, Y., & Honma, K. clock factors, and phospholipase C. Molecular and Cellular
(2007). Separate oscillating cell groups in mouse Biology, 24, 3359–3372.
80

Majercak, J., Sidote, D., Hardin, P. E., & Edery, I. (1999). Proceedings of the National Academy of Sciences of the
How a circadian clock adapts to seasonal decreases in tem- United States of America, 97, 3608–3613.
perature and day length. Neuron, 24, 219–230. Picot, M., Cusumano, P., Klarsfeld, A., Ueda, R., & Rouyer, F.
Malpel, S., Klarsfeld, A., & Rouyer, F. (2002). Larval optic (2007). Light activates output from evening neurons and
nerve and adult extra-retinal photoreceptors sequentially inhibits output from morning neurons in the Drosophila cir-
associate with clock neurons during Drosophila brain devel- cadian clock. PLoS Biology, 5, e315.
opment. Development, 129, 1443–1453. Picot, M., Klarsfeld, A., Chelot, E., Malpel, S., & Rouyer, F.
Martinek, S., Inonog, S., Manoukian, A. S., & Young, M. W. (2009). A role for blind DN2 clock neurons in temperature
(2001). A role for the segment polarity gene shaggy/GSK-3 entrainment of the Drosophila larval brain. The Journal of
in the Drosophila circadian clock. Cell, 105, 769–779. Neuroscience, 29, 8312–8320.
McNabb, S. L., Baker, J. D., Agapite, J., Steller, H., Pittendrigh, C. S., Bruce, V. G., & Kaus, P. (1958). On the sig-
Riddiford, L. M., & Truman, J. W. (1997). Disruption of a nificance of transients in daily rhythms. Proceedings of the
behavioral sequence by targeted death of peptidergic National Academy of Sciences of the United States of
neurons in Drosophila. Neuron, 19, 813–823. America, 44, 965–973.
Meijer, J. H., Michel, S., Vanderleest, H. T., & Rohling, J. H. Pittendrigh, C. S., & Daan, S. (1976). A functional analysis of
(2010). Daily and seasonal adaptation of the circadian circadian pacemakers in nocturnal rodents. V. Pacemaker
clock requires plasticity of the SCN neuronal network. The structure: A clock for all seasons. The Journal of Compara-
European Journal of Neuroscience, 32, 2143–2151. tive Physiology A, 106, 333–355.
Miyasako, Y., Umezaki, Y., & Tomioka, K. (2007). Separate Plautz, J. D., Kaneko, M., Hall, J. C., & Kay, S. A. (1997).
sets of cerebral clock neurons are responsible for light and Independent photoreceptive circadian clocks throughout
temperature entrainment of Drosophila circadian locomotor Drosophila. Science, 278, 1632–1635.
rhythms. Journal of Biological Rhythms, 22, 115–126. Potter, C. J., & Luo, L. (2011). Using the Q system in
Murad, A., Emery-Le, M., & Emery, P. (2007). A subset of Drosophila melanogaster. Nature Protocols, 6, 1105–1120.
dorsal neurons modulates circadian behavior and light Preuss, F., Fan, J. Y., Kalive, M., Bao, S., Schuenemann, E.,
responses in Drosophila. Neuron, 53, 689–701. Bjes, E. S., et al. (2004). Drosophila doubletime mutations
Myers, M. P., Wager-Smith, K., Rothenfluh-Hilfiker, A., & which either shorten or lengthen the period of circadian
Young, M. W. (1996). Light-induced degradation of TIME- rhythms decrease the protein kinase activity of casein kinase
LESS and entrainment of the Drosophila circadian clock. I. Molecular and Cellular Biology, 24, 886–898.
Science, 271, 1736–1740. Renn, S. C., Park, J. H., Rosbash, M., Hall, J. C., &
Naito, E., Watanabe, T., Tei, H., Yoshimura, T., & Ebihara, S. Taghert, P. H. (1999). A pdf neuropeptide gene mutation
(2008). Reorganization of the suprachiasmatic nucleus and ablation of PDF neurons each cause severe
coding for day length. Journal of Biological Rhythms, 23, abnormalities of behavioral circadian rhythms in Drosoph-
140–149. ila. Cell, 99, 791–802.
Nitabach, M. N., Blau, J., & Holmes, T. C. (2002). Electrical Rieger, D., Fraunholz, C., Popp, J., Bichler, D., Dittmann, R.,
silencing of Drosophila pacemaker neurons stops the free- & Helfrich-Förster, C. (2007). The fruit fly Drosophila
running circadian clock. Cell, 109, 485–495. melanogaster favors dim light and times its activity peaks
Nitabach, M. N., Wu, Y., Sheeba, V., Lemon, W. C., to early dawn and late dusk. Journal of Biological Rhythms,
Strumbos, J., Zelensky, P. K., et al. (2006). Electrical 22, 387–399.
hyperexcitation of lateral ventral pacemaker neurons Rieger, D., Peschel, N., Dusik, V., Glotz, S., & Helfrich-
desynchronizes downstream circadian oscillators in the fly Förster, C. (2012). The ability to entrain to long
circadian circuit and induces multiple behavioral periods. photoperiods differs between three D. melanogaster wild-
The Journal of Neuroscience, 26, 479–489. type strains and is modified by twilight simulation. Journal
Parisky, K. M., Agosto, J., Pulver, S. R., Shang, Y., Kuklin, E., of Biological Rhythms, 27, 37–47.
Hodge, J. J., et al. (2008). PDF cells are a GABA-responsive Rieger, D., Shafer, O. T., Tomioka, K., & Helfrich-Förster, C.
wake-promoting component of the Drosophila sleep circuit. (2006). Functional analysis of circadian pacemaker neurons
Neuron, 60, 672–682. in Drosophila melanogaster. The Journal of Neuroscience,
Park, J. H., & Hall, J. C. (1998). Isolation and chronobiological 26, 2531–2543.
analysis of a neuropeptide pigment-dispersing factor gene in Rieger, D., Stanewsky, R., & Helfrich-Förster, C. (2003).
Drosophila melanogaster. Journal of Biological Rhythms, 13, Cryptochrome, compound eyes, Hofbauer-Buchner eyelets,
219–228. and ocelli play different roles in the entrainment and mas-
Park, J. H., Helfrich-Förster, C., Lee, G., Liu, L., Rosbash, M., king pathway of the locomotor activity rhythm in the fruit
& Hall, J. C. (2000). Differential regulation of circadian fly Drosophila melanogaster. Journal of Biological Rhythms,
pacemaker output by separate clock genes in Drosophila. 18, 377–391.
81

Rieger, D., Wülbeck, C., Rouyer, F., & Helfrich-Förster, C. Stanewsky, R., Frisch, B., Brandes, C., Hamblen-Coyle, M. J.,
(2009). Period gene expression in four neurons is sufficient Rosbash, M., & Hall, J. C. (1997). Temporal and spatial
for rhythmic activity of Drosophila melanogaster under expression patterns of transgenes containing increasing
dim light conditions. Journal of Biological Rhythms, 24, amounts of the Drosophila clock gene period and a lacZ
271–282. reporter: Mapping elements of the PER protein involved
Rosato, E., Codd, V., Mazzotta, G., Piccin, A., Zordan, M., in circadian cycling. The Journal of Neuroscience, 17,
Costa, R., et al. (2001). Light-dependent interaction 676–696.
between Drosophila CRY and the clock protein PER medi- Stoleru, D., Nawathean, P., Fernandez, M. P., Menet, J. S.,
ated by the carboxy terminus of CRY. Current Biology, 11, Ceriani, M. F., & Rosbash, M. (2007). The Drosophila circa-
909–917. dian network is a seasonal timer. Cell, 129, 207–219.
Schwartz, W. J. (2004). Sunrise and sunset in fly brains. Stoleru, D., Peng, Y., Agosto, J., & Rosbash, M. (2004). Cou-
Nature, 431, 751–752. pled oscillators control morning and evening locomotor
Schwartz, W. J., & de la Iglesia, H. O. (2003). Dual circadian behaviour of Drosophila. Nature, 431, 862–868.
oscillators in the suprachiasmatic nucleus. In K. Homna & Stoleru, D., Peng, Y., Nawathean, P., & Rosbash, M. (2005).
S. Honma (Eds.), Circadian clock as multi-oscillation system. A resetting signal between Drosophila pacemakers synchro-
Sapporo, Japan: Hokkaido University Press. nizes morning and evening activity. Nature, 438, 238–242.
Shafer, O. T., Helfrich-Förster, C., Renn, S. C., & Taghert, P. H., Hewes, R. S., Park, J. H., O’Brien, M. A., Han, M.,
Taghert, P. H. (2006). Reevaluation of Drosophila & Peck, M. E. (2001). Multiple amidated neuropeptides are
melanogaster’s neuronal circadian pacemakers reveals new required for normal circadian locomotor rhythms in Drosoph-
neuronal classes. The Journal of Comparative Neurology, ila. The Journal of Neuroscience, 21, 6673–6686.
498, 180–193. Tomioka, K., Sakamoto, M., Harui, Y., Matsumoto, N., &
Shafer, O. T., Rosbash, M., & Truman, J. W. (2002). Sequen- Matsumoto, A. (1998). Light and temperature cooperate to
tial nuclear accumulation of the clock proteins period regulate the circadian locomotor rhythm of wild type and
and timeless in the pacemaker neurons of Drosophila period mutants of Drosophila melanogaster. Journal of
melanogaster. The Journal of Neuroscience, 22, 5946–5954. Insect Physiology, 44, 587–596.
Shang, Y., Griffith, L. C., & Rosbash, M. (2008). Light-arousal Umezaki, Y., & Tomioka, K. (2008). Behavioral dissection of
and circadian photoreception circuits intersect at the large the Drosophila circadian multioscillator system that regulates
PDF cells of the Drosophila brain. Proceedings of the locomotor rhythms. Zoological Science, 25, 1146–1155.
National Academy of Sciences of the United States of Vanin, S., Butani, S., Montelli, S., Menegazzi, P.,
America, 105, 19587–19594. Green, E. W., Pegoraro, M., et al. (2012). Unexpected
Sheeba, V., Fogle, K. J., & Holmes, T. C. (2010). Persistence features of Drosophila circadian behavioural rhythms under
of morning anticipation behavior and high amplitude morn- natural conditions. Nature, 484, 371–375.
ing startle response following functional loss of small ventral VanVickle-Chavez, S. J., & Van Gelder, R. N. (2007). Action
lateral neurons in Drosophila. PLoS One, 5, e11628. spectrum of Drosophila cryptochrome. The Journal of
Sheeba, V., Fogle, K. J., Kaneko, M., Rashid, S., Chou, Y. T., Biological Chemistry, 282, 10561–10566.
Sharma, V. K., et al. (2008). Large ventral lateral neurons Veleri, S., Rieger, D., Helfrich-Förster, C., & Stanewsky, R.
modulate arousal and sleep in Drosophila. Current Biology, (2007). Hofbauer-Buchner eyelet affects circadian photosen-
18, 1537–1545. sitivity and coordinates TIM and PER expression in Dro-
Sheeba, V., Gu, H., Sharma, V. K., O’Dowd, D. K., & sophila clock neurons. Journal of Biological Rhythms, 22,
Holmes, T. C. (2008). Circadian- and light-dependent regu- 29–42.
lation of resting membrane potential and spontaneous Wheeler, D. A., Hamblen-Coyle, M. J., Dushay, M. S., &
action potential firing of Drosophila circadian pacemaker Hall, J. C. (1993). Behavior in light-dark cycles of Drosoph-
neurons. Journal of Neurophysiology, 99, 976–988. ila mutants that are arrhythmic, blind, or both. Journal of
Sheeba, V., Kaneko, M., Sharma, V. K., & Holmes, T. C. Biological Rhythms, 8, 67–94.
(2008). The Drosophila circadian pacemaker circuit: Pas Wu, Y., Cao, G., Pavlicek, B., Luo, X., & Nitabach, M. N.
De Deux or Tarantella? Critical Reviews in Biochemistry (2008). Phase coupling of a circadian neuropeptide with
and Molecular Biology, 43, 37–61. rest/activity rhythms detected using a membrane-tethered
Sosniyenko, S., Hut, R. A., Daan, S., & Sumova, A. (2009). spider toxin. PLoS Biology, 6, e273.
Influence of photoperiod duration and light–dark trans- Wülbeck, C., Grieshaber, E., & Helfrich-Förster, C. (2008).
itions on entrainment of Per1 and Per2 gene and protein Pigment-dispersing factor (PDF) has different effects on
expression in subdivisions of the mouse suprachiasmatic Drosophila’s circadian clocks in the accessory medulla and
nucleus. The European Journal of Neuroscience, 30, in the dorsal brain. Journal of Biological Rhythms, 23,
1802–1814. 409–424.
82

Yang, Z., & Sehgal, A. (2001). Role of molecular oscillations Yoshii, T., Todo, T., Wülbeck, C., Stanewsky, R., & Helfrich-
in generating behavioral rhythms in Drosophila. Neuron, Förster, C. (2008). Cryptochrome is present in the compound
29, 453–467. eyes and a subset of Drosophila’s clock neurons. The Journal
Yoshii, T., Fujii, K., & Tomioka, K. (2007). Induction of of Comparative Neurology, 508, 952–966.
Drosophila behavioral and molecular circadian rhythms by Yoshii, T., Vanin, S., Costa, R., & Helfrich-Förster, C. (2009).
temperature steps in constant light. Journal of Biological Synergic entrainment of Drosophila’s circadian clock by
Rhythms, 22, 103–114. light and temperature. Journal of Biological Rhythms, 24,
Yoshii, T., Funada, Y., Ibuki-Ishibashi, T., Matsumoto, A., 452–464.
Tanimura, T., & Tomioka, K. (2004). Drosophila cryb muta- Yoshii, T., Wülbeck, C., Sehadova, H., Veleri, S., Bichler, D.,
tion reveals two circadian clocks that drive locomotor Stanewsky, R., et al. (2009). The neuropeptide pigment-
rhythm and have different responsiveness to light. Journal dispersing factor adjusts period and phase of Drosophila’s
of Insect Physiology, 50, 479–488. clock. The Journal of Neuroscience, 29, 2597–2610.
Yoshii, T., Hermann, C., & Helfrich-Förster, C. (2010). Zhang, L., Chung, B. Y., Lear, B. C., Kilman, V. L., Liu, Y.,
Cryptochrome-positive and -negative clock neurons in Mahesh, G., et al. (2010). DN1(p) circadian neurons
Drosophila entrain differentially to light and temperature. coordinate acute light and PDF inputs to produce
Journal of Biological Rhythms, 25, 387–398. robust daily behavior in Drosophila. Current Biology, 20,
Yoshii, T., Heshiki, Y., Ibuki-Ishibashi, T., Matsumoto, A., 591–599.
Tanimura, T., & Tomioka, K. (2005). Temperature cycles Zhang, Y., Liu, Y., Bilodeau-Wentworth, D., Hardin, P. E., &
drive Drosophila circadian oscillation in constant light that Emery, P. (2010). Light and temperature control the contri-
otherwise induces behavioural arrhythmicity. The European bution of specific DN1 neurons to Drosophila circadian
Journal of Neuroscience, 22, 1176–1184. behavior. Current Biology, 20, 600–605.
Yoshii, T., Sakamoto, M., & Tomioka, K. (2002). A tempera- Zhao, J., Kilman, V. L., Keegan, K. P., Peng, Y., Emery, P.,
ture-dependent timing mechanism is involved in the circa- Rosbash, M., et al. (2003). Drosophila clock can generate
dian system that drives locomotor rhythms in the fruit fly ectopic circadian clocks. Cell, 113, 755–766.
Drosophila melanogaster. Zoological Science, 19, 841–850.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 5

Circadian system from conception till adulthood

Alena Sumova*, Martin Sladek, Lenka Polidarova, Marta Novakova and


Pavel Houdek

Institute of Physiology, Academy of Sciences of the Czech Republic, Prague, Czech Republic

Abstract: In mammals, the circadian system is composed of the central clock in the hypothalamic
suprachiasmatic nuclei and of peripheral clocks that are located in other neural structures and in cells of
the peripheral tissues and organs. In adults, the system is hierarchically organized so that the central
clock provides the other clocks in the body with information about the time of day. This information is
needed for the adaptation of their functions to cyclically changing external conditions. During
ontogenesis, the system undergoes substantial development and its sensitivity to external signals
changes. Perinatally, maternal cues are responsible for setting the phase of the developing clock, while
later postnatally, the LD cycle is dominant. The central clock attains its functional properties during a
gradual and programmed process. Peripheral clocks begin to exhibit rhythmicity independent of each
other at various developmental stages. During the early developmental stages, the peripheral clocks are
set or driven by maternal feeding, but later the central clock becomes fully functional and begins to
entrain the periphery. During the perinatal period, the central and peripheral clocks seem to be
vulnerable to disturbances in external conditions. Further studies are needed to understand the
processes of how the circadian system develops and what degree of plasticity and resilience it possesses
during ontogenesis. These data may lead to an assessment of the contribution of disturbances of the
circadian system during early ontogenesis to the occurrence of circadian diseases in adulthood.

Keywords: ontogenesis; suprachiasmatic nucleus; peripheral circadian clocks; clock gene.

Introduction organism grows and matures. Increasing awareness


of the mechanisms of how the system operates at
The mammalian internal timekeeping system, molecular level facilitates studies of the develop-
which temporally controls many processes in our ment of the system from its very early stages, that
body, undergoes developmental changes as an is, far before the overt rhythms driven by the
system can be used as a marker of its functional
*Corresponding author. stage. These studies may uncover the mystery of
Tel.: þ420-24106-2528; Fax: þ420-24106-2488 when the circadian clock begins to tick indepen-
E-mail: sumova@biomed.cas.cz dent of the maternal environment. It is possible

http://dx.doi.org/10.1016/B978-0-444-59427-3.00005-8 83
84

that the circadian clock is functional early prena- the ventral hypothalamus just above the optic chi-
tally when it resides in the maternal body and the asm (Klein et al., 1991). The SCN is a complex
mother’s role is just to adjust its phase according structure, which is composed of a cluster of inde-
to the external light/dark (LD) cycle to ensure that pendent cellular oscillators (Welsh et al., 1995)
her offspring are born with fully entrained clocks. that are mutually coupled by a web of synapses.
Alternatively, it is also possible that the developing This intercellular communication is responsible
circadian clock is first directly driven by maternal for synchrony among the cells (Liu et al., 2007).
signals as a slave oscillator in utero and develops Moreover, the SCN cells are organized into
into a self-sustained clock that is sensitive to morphologically and functionally distinct sub-
the external LD cycle only gradually after birth. populations (Klein et al., 1991) which also
Both possibilities would ensure that the newborn communicate among each other. The most prom-
leaves the maternal body with its clock properly inent subpopulations are located in the ventrolat-
entrained with the external environment, but it eral (core) and dorsomedial (shell) regions of the
remains unclear which of these possibilities would SCN (for a review, see Welsh et al., 2010), but
be evolutionally more advantageous. It is also pos- there are also other subdivisions, including a func-
sible that different species utilize various scenarios tional specialization in populations of cells in the
of the development. This issue is important for rostro-caudal dimension of the SCN (Brown and
understanding the significance of mother-to-fetus Piggins, 2009; Morin, 2007; Sosniyenko et al.,
and mother-to-neonate communication for proper 2009). As a result of this intercellular communica-
development of the circadian system. Enormous tion, the SCN generates a synchronized output
amounts of data have been accumulated to dem- signal to the rest of the body that entrains the
onstrate that a malfunction of the circadian system peripheral clocks with the external LD cycles.
leads to the distortion of temporal control of phys- This chapter provides a summary of our latest
iology and results in the development of serious knowledge and offers some hypotheses on the pro-
diseases. Studies on the ontogenesis of the circa- cesses by which the circadian system develops in
dian system may reveal whether its malfunction mammals from its origin through weaning and
in adulthood may result from an insult during early adulthood. The summary includes mostly data from
developmental stages, which might arise from the rodent models. For a more extensive overview of
disruption of the endogenous timekeeping system the subject, the readers are encouraged to refer to
during pregnancy. previous reviews (e.g., Weinert, 2005; Sumova
The circadian system is composed of hierar- et al., 2006a,b, 2008; Seron-Ferre et al., 2007).
chically organized circadian clocks forming an
integral regulatory system that gives the organism
the daily temporal program and helps it adapt to Developing circadian system in utero
the external environment. Circadian clocks reside
in nearly every, if not all, mammalian cells (for a Prefetal development
review, see Schibler et al., 2003). These clock cells
are equipped with a set of so-called clock genes, The development of circadian clocks in mammals
which are genes that are indispensable for circa- is intrinsically programmed and does not depend
dian function. In mammals, Per1, Per2, Cry1, on presence of a functional maternal clock (Davis
Cry2, Bmal1, Clock, Rev-erba, and Rora have and Gorski, 1988; Jud and Albrecht, 2006;
been recognized as clock genes (for a review, Reppert and Schwartz, 1986; Shibata and Moore,
see Takahashi et al., 2008). Cellular clocks are 1988). At the cellular level, the first appearance
regularly entrained by the master pacemaker of circadian rhythmicity has been detected in
located in the suprachiasmatic nuclei (SCN) of differentiated, multipotent cells derived from
85

embryonic stem cells but not at earlier stages such Rat


as pluripotent cells (Kowalska et al., 2010; Yagita
E0 Conception
et al., 2010). Unfertilized mouse oocytes contain
transcripts of the clock genes (Johnson et al.,
2002), but their levels do not exhibit rhythms in
oocytes and preimplantation embryos (Amano
et al., 2009). The embryonic stem cells first differ-
entiate into neural precursor cells, which further
divide and may differentiate into neurons and E15 SCN formation
astrocytes (Gotz and Sommer, 2005). The differ-
entiated cells possess the circadian clock, which E22/P0 Birth
is lost if they are forced to dedifferentiate back
into embryonic stem cells (Yagita et al., 2010).
P5
Thus, the earliest detection of the circadian clock Photic entrainment
is in studies employing differentiated neural synaptogenesis completed
precursor cell lines. The presence of a clock in P10
these studies was detected by rhythmic clock
and clock-controlled gene transcription, using P15 Eye opening
Bmal1- and Dbp-promoter-driven luciferase
reporters (Kowalska et al., 2010; Yagita et al., P21
2010) and RT-PCR (Kowalska et al., 2010). How- Weaning
ever, information about the levels of protein
products of the clock genes at these stages is P28
lacking. Also, these studies must rely on employing
cell lines and it is possible that the cells and the Fig. 1. Schematic diagram of the main circadian milestones of the
process of their differentiation may not exactly rat ontogenesis. Vertical bar represents a timescale from
match the situation in vivo. Nevertheless, these embryonic day (E) 0 till postnatal day (P) 28 (for details see text).
studies provide insight into the origin of the cellu-
lar circadian clock in mammalian development.
It is intriguing that the presence of transcripts during the prenatal period depends on the differ-
of the clock gene Bmal1 was confirmed already entiation and maturation of the structure in ques-
in unfertilized eggs at 2-cell and 16-cell stages tion. The main circadian milestones of the rat
(Ko et al., 2000). This finding may suggest that this ontogenesis are depicted in Fig. 1.
gene is expressed intermittently, skipping some Morphologically, the rodent SCN develops
developmental stages. Moreover, this finding gradually (Moore, 1991). For example, in the
may imply that Bmal1 may be involved in various rat, neurogenesis of the SCN begins on embry-
processes already during ontogenesis, such as the onic day (E) 14 and continues through E17,
timing of cell division in the mouse embryo (Ko originating from a specialized zone of the ventral
et al., 2000). diencephalic germinal epithelium as a component
of periventricular cell groups. Neurons of the ven-
trolateral SCN are generated at E15–E16 and
Development of the fetal SCN clock those of the dorsomedial SCN at E16–E17. Neu-
rogenesis is completed at E18, but the morpho-
At the structural level, the development of the logical maturation of the SCN neurons gradually
circadian clocks in the SCN or peripheral organs proceeds until postnatal day (P) 10. The genetic
86

program that is responsible for the formation of of synapses. The presence of a functional network
the SCN seems to be dependent on the presence in the SCN seems to be one of the basic
of the Six3 transcriptional factor (VanDunk mechanisms of how the SCN generates circadian
et al., 2011). rhythms. Interestingly, the network has been
Well before the circadian molecular clock recently attributed to maintenance of noise-
mechanism was deciphered, studies in mammals induced stochastic circadian rhythms that may
using overt circadian rhythms as a marker of the arise as its emergent property (Ko et al., 2010;
functional state of the circadian clock strongly Mohawk and Takahashi, 2011). It was hypothe-
suggested the presence of a functional circadian sized that due to this function, circadian rhythms
clock during the fetal stage, and this statement may exist even in the absence of cell-autonomous
has been widely accepted for decades (for a circadian oscillators (Ko et al., 2010). Accord-
review, see Weinert, 2005). The question of when ingly, the population of clock cells with low
exactly during the fetal stage the molecular degree of intercellular communication, and there-
machinery begins to operate was reinitiated by a fore with a low level of the stochastic rhythmicity,
rather surprising finding that clock gene expres- should generate rhythms with lower amplitude
sion profiles in the rat SCN of 19-day-old fetuses, than the population with a high degree of commu-
that is, 3–4 days before delivery, do not exhibit nication. This is in agreement with the finding
circadian rhythms (Sladek et al., 2004). Moreover, that, during fetal stage, the rodent SCN with an
levels of protein products of the clock genes undeveloped intercellular communication system
PER1, PER2, and CRY1, which are indispens- exhibited no or only low-amplitude rhythms in
able for clock function, could not be detected by clock gene expression within the population of
conventional immunocytochemistry in the SCN the cells (Ansari et al., 2009; Li and Davis, 2005;
of these fetuses (Sladek et al., 2004). This finding Sladek et al., 2004). Moreover, the amplitudes of
suggests that the population of SCN cells might the clock gene expression rhythms gradually
not be able to generate synchronized rhythms at increased as a function of the web development
this developmental stage and facilitated further during the postnatal stage and achieved adult-like
studies to identify the developmental stage at levels only when the development was completed
which the circadian clock begins to tick. (Kovacikova et al., 2006). These data suggest that
After completing neurogenesis, the SCN fur- the stochastic rhythms likely develop later during
ther develops gradually during ontogenesis ontogenesis together with the formation of the
(Moore and Bernstein, 1989). In rodents, the intercellular web as an additional mechanism to
newly formed SCN neurons begin to spread their fortify the cell-autonomous rhythms and the clock
processes across the nuclei to make contact with rhythmicity. It remains to be deciphered whether
other neurons and to create an intercellular web the stochastic oscillations are conditional for the
during the late prenatal and mainly early postna- clock function.
tal period. In the rat SCN, very sparse synapses Studies where gene expression was recorded in
may be observed at E19–E22, but not earlier, so real time from SCN slices in vitro have suggested
indicating that fetal SCN cells are virtually devoid that the fetal SCN as a whole structure might have
of synapses (Moore and Bernstein, 1989). a potential to oscillate. The SCN slices explanted
Similarly, astroglia were only sparse in the SCN from 22-day-old fetuses of transgenic rat exhibit
during the embryonic stage (Munekawa et al., rhythm in Per1:luciferase when cultured in vitro
2000). Therefore, during the prenatal period, (Ohta et al., 2008). Interestingly, at this time, only
neurogenesis of the SCN is completed, but the faint rhythms were detected in the SCN in vivo
multilevel intercellular coupling may not yet be (Kovacikova et al., 2006). It is possible that these
functional due to the lack of reasonable number luminescence recordings from the SCN slices are
87

more sensitive than the detection of gene expres- vasoactive intestinal peptide (Vip) (Ginty et al.,
sion by in situ hybridization. However, it may also 1993; Hahm and Eiden, 1998; Iwasaki et al.,
imply that a coherent signal synchronizing the indi- 1997; Obrietan et al., 1999). In this context, it is
vidual oscillators is missing in vivo, but when the interesting to point out that maternal melatonin,
cells are explanted, the culture medium and/or whose role in the entrainment of the fetal SCN
the dissection procedure might be sufficient to syn- is still being questioned, seems to be a likely can-
chronize them. This hypothesis was supported by a didate of such rhythmic cue; it acts through this
recent observation that at E18, various mouse pathway by inhibiting production of cAMP (for
peripheral tissues exhibit rhythmicity in clock gene review, see Hastings, 1997; Vanecek, 1998) and
expression in culture but not in vivo (Dolatshad no direct immediate effect of melatonin on clock
et al., 2010). Similarly, the time when the gene expression has been detected thus far
organotypic SCN slice is dissected can influence (Poirel et al., 2003). Similarly, the cAMP pathway
the phase of the resulting in vitro oscillations is employed by dopamine, which is also a likely
(Yoshikawa et al., 2005). Importantly, placing the maternal signal (Viswanathan and Davis, 1997;
rat SCN tissue from arrhythmic animals into culture Viswanathan et al., 1994). If this scenario is true,
can even induce a rhythm in vitro (Yoshikawa et al., it would mean that even though the SCN cells lack
2005). These data underline the importance of the mutually synchronized molecular clocks, they may
presence of a synchronizing cue for existence of cir- respond synchronously to melatonin or any other
cadian rhythmicity at early developmental stages. mediator using a cAMP-signaling pathway. This
Although detectable rhythms in clock gene response would appear at any time and would be
expression are absent in the fetal SCN cells independent of the phase of the desynchronized
in vivo, the population of cells likely oscillates in individual cellular oscillators. In contrast, once
its metabolic activity (Reppert and Schwartz, the SCN cells establish their communication web
1983, 1984), in spite of the fact that an absence later in development, the response becomes highly
of the rhythm in the fetal SCN was reported in dependent on the circadian phase (Ginty et al.,
an earlier report (Fuchs and Moore, 1980). More- 1993; Shearman et al., 1997; Shigeyoshi et al.,
over, the neurotransmitter arginine vasopressin 1997). Therefore, it is plausible to hypothesize that
(Avp) (Reppert and Uhl, 1987) was also found the cAMP signaling might play a role in entraining
to be rhythmic in a population of SCN cells dur- the phases of various rhythmical processes in the
ing the prenatal stage. It means that these individual cells even before their clocks are mutu-
functions must be driven rhythmically in a syn- ally synchronized. Interestingly, once functional
chronous manner even though the population of synapses are formed during postnatal stages, the
SCN cells does not evince rhythmicity in clock cAMP-signaling pathway is likely to be involved
gene expression. It is difficult to understand how in coupling and strengthening the oscillation of
this situation arises, especially because at least molecular core clock mechanisms (O’Neill et al.,
some of these rhythms are known to be driven 2008). It remains unclear why the molecular clock
by the molecular clock mechanism in adults (Jin mechanism lacks this property during the syn-
et al., 1999). There is a possibility that during apse-lacking, fetal stage.
the fetal stage, these rhythms are directly driven The development of the SCN clock gene rhyth-
by rhythmic maternal signals independent of the micity may be dependent on the species-specific
molecular clockwork. These signals may be medi- degree of fetal maturation at the time of delivery.
ated by a common pathway, which likely involves In the capuchin monkey, a rhythm in the expres-
activation of cAMP and phosphorylation of sion of Per2 and Bmal1 was reported in the
CREB, because these steps are necessary for fetal SCN at the time corresponding to 90% of
activation of transcription of c-fos, Avp, and gestation (Torres-Farfan et al., 2006), although
88

these rhythms could only be fitted on the basis of developmental stages because the undeveloped
the levels at three time points when the animals SCN was not able to synchronize them. Indeed,
were sampled during the 24-h cycle. For obvious in adult animals, ablation of the signaling from
reasons, using nonrodent models cannot provide the SCN to the peripheral clocks disrupts the
temporal resolution sufficient for reliable deter- synchrony among the individual cells, which leads
mination of the rhythmicity and phase of the fetal to gradual loss of the rhythm at the level of the
clock in vivo. entire structure (Akhtar et al., 2002). However,
this simple explanation was seemingly opposed
Development of the fetal peripheral clocks by the finding that one of the clock genes, namely,
Rev-erba, exhibited high-amplitude rhythms in the
Only a few studies have examined the develop- fetal rat liver (Sladek et al., 2007). Theoretically,
ment of peripheral clocks during the prenatal there might be two explanations: (i) the fetal
stage and those have mostly focused on the liver. hepatocytes are already synchronized at the fetal
In mice, genetic markers of liver cells are first stage but their molecular clockwork is not yet fully
detected in tissue arising from the primitive gut functional or (ii) the fetal hepatocytes are not syn-
between E8.5 and E9 (Cascio and Zaret, 1991). chronized and Rev-erba expression is selectively
The early liver primordium originates from endo- driven rhythmically and synchronously by putative
dermal tissue and contains mostly bipotential external signals derived from the mother; this pro-
liver progenitor cells capable of differentiating cess would be analogous to the cAMP-signaling
into either hepatocytes or biliary epithelium cues in the fetal SCN (see above). The Rev-erba
(Dabeva et al., 2000; Shiojiri and Mizuno, 1993). gene is likely sensitive to various maternal sig-
The fetal liver is also the anatomic site of hemato- naling molecules such as glucocorticoids (Torra
poiesis. Hematopoietic progenitor cells, which are et al., 2000), which can pass from mother to fetuses
distinct from the hepatoblast lineage, occur in the despite their partial metabolic degradation in the
liver beginning at E11 (Dzierzak and Medvinsky, placenta (Staud et al., 2006). The latter possibility
1995). The later phase of fetal development seems to be more plausible because, at the cellular
is characterized by extensive growth of the level, a functional clock mechanism has been
organ mass, differentiation of the bipotential shown to exist in peripheral tissues at early prena-
hepatoblasts into mature parenchymal cell types, tal stages (Pando et al., 2002; Yagita et al., 2001).
and cessation of hematopoiesis (Walker et al., In the rat liver, the expression of Rev-erba peaked
2001). At E16, the rat liver is mainly composed in the subjective night at E20, which was opposite
of relatively well-differentiated hepatoblasts. As in phase to the expression in the adult liver
liver development progresses, the liver structure (Sladek et al., 2007). Therefore, while the SCN
starts to resemble that of the adult (Van Eyken clock is born in proper phase with the external
et al., 1988). LD cycle, the liver clock is born with a clock gene
Studies on the development of rhythmicity of cycling in the opposite phase (Fig. 2). This differ-
clock gene expression in the peripheral tissue rev- ence might be an adaptation of the peripheral
ealed results similar to those seen in the SCN; clock to the diurnal pattern of breast feeding in
the explanted fetal liver exhibited circadian rhyth- nocturnal animals (see below). It would be inter-
micity in vitro (Dolatshad et al., 2010), but in vivo esting to determine whether peripheral clocks also
expression of most clock genes detected by develop at an opposite phase in diurnal animals
RT-PCR did not exhibit significant rhythms and to compare the development of the circadian
(Dolatshad et al., 2010; Sladek et al., 2007). The clock in the liver with a clock located in another
most likely explanation is that the individual peripheral tissue that is not directly related with
hepatocytes were desynchronized during early feeding, such as the lung or heart. Currently
89

SCN Liver
Per1 Bmal1 Per1 Rev-erba

Fetal

P2

P10

P30

AD

Fig. 2. Schematic drawings of the development of circadian clocks in the rat. Daily profiles of clock genes Per1 and Bmal1 mRNA
in the SCN and Per1 and Rev-erba in the liver are depicted in fetuses; pups at postnatal day 2 (P2), P10, and P30; and adult rats.
X-axis represents daytime with the shaded area defining night hours. Y-axis represents relative mRNA levels. Drawings are
based on results published previously (Kovacikova et al., 2006; Sladek et al., 2004, 2007) and are adapted according to Sumova
et al. (2008) (with permission).

available data on the development of the clock the prenatal stage is still not fully understood.
gene expression rhythms also lead to speculation Specifically, a possibility exists that detection of
that, in the liver, the rhythmic expression of a rhythm at the cell population level in the unde-
Rev-erba might trigger the newly appearing veloped central and peripheral clocks might not
rhythms in clock gene expression during the post- a priori mean that the molecular oscillations in
natal stage because a constant phase relationship the individual cells of the clocks are mutually syn-
between the rhythms in Rev-erba and other clock chronized. A putative external rhythmical cue
genes is maintained during all the developmental derived from the maternal environment might
stages studied (Sladek et al., 2007). impinge on the desynchronized clock cells and
In addition to the liver, circadian clock gene induce or reduce non-clock gene expression in a
expression has also been studied in the fetal heart rhythmical manner according to its appearance
and kidney (Dolatshad et al., 2010). Similar to the without a change in the phase of the cycling cells
liver, both tissues exhibited a circadian rhythm in (Fig. 3). If this hypothesis is true, one might fur-
Per2:luciferace in vitro, but no rhythm in Per2 and ther speculate that, in the SCN, this imposed
Bmal1 expression was detected in vivo (Dolatshad rhythmicity may help to set the phase to which
et al., 2010). the whole population of cellular oscillators fixes
during development once the intercellular web is
formed. This scenario could also explain why the
Entrainment of the clocks during the fetal stage clock may shift due to prenatally imposing cues
even though the individual cells are not synchro-
The above summarized data suggest that develop- nized. Under such conditions, the maternal cues
ment of the molecular clock mechanism during likely affect the desynchronized cells at any time
90

Fig. 3. Hypothesis on development of the synchronized circadian rhythms in the SCN during ontogenesis. During the fetal stage, the
individual oscillators are not interconnected, and therefore, no circadian rhythm in clock gene expression can be detected in
population of the SCN cells. Putative external rhythmic cues derived from the maternal environment might impinge on the
desynchronized clock cells and induce or reduce a non-clock gene expression (e.g., Avp, c-fos) in a rhythmical manner according
to their appearance without a change in the phase of the cycling cells. Apparently, such effects may appear any time of the day.
During the late prenatal and early postnatal stage, the SCN oscillators become gradually interconnected via a web of synapses.
Due to the gradual synchronization, the clock gene expression begins to exhibit low-amplitude rhythms, which may fortify
expression of clock-controlled genes (e.g., Avp). The maternal or external cues may entrain the oscillators only during a specific
window of time according to the development of the gating mechanism. During the late postnatal period, both the intercellular
communication and the gating mechanism are fully developed and achieve the adult-like stage, resulting in high-amplitude
oscillations of the clock and clock-controlled gene expression (for details, see text).
91

of the day, while in the case of a highly synchroni- mechanism supports the hypothesis mentioned
zed population, the effect would be gated to a above regarding imposed rhythmicity on the
specific time window, as happens in adults. This desynchronized fetal SCN cells.
might be the reason of a lack of evolutionary Although the participation of the melatonin/
pressure to develop intercellular communication dopamine couple in fetal entrainment seems
among the fetal SCN neurons. There are numer- likely, it is not the only possible mechanism by
ous pieces of experimental evidence that the which the fetal clocks perceive temporal cues.
phase of the developing SCN clock may already Other likely candidates are related to maternal
be set prenatally. The evidence comes from ear- behavior and may involve temperature or feeding
lier studies in which the phase of the fetal clock cycles. Exposure of pregnant rats with surgically
was extrapolated from the phase of behavioral removed SCN to a restricted feeding regime was
activity when it developed in pups postnatally able to shift the phase of the SCN clocks in their
(Davis and Gorski, 1988; Reppert and Schwartz, fetuses as measured by a shift in the phase of
1984; Viswanathan and Davis, 1997; Weaver and the drinking activity of pups after weaning
Reppert, 1987). More recently, a shift of the fetal (Weaver and Reppert, 1989). Recently, another
SCN by maternal cues has been demonstrated by study demonstrated that the fetal clock is
the detection of the phase of gene expression entrained by a maternal-restricted feeding regime
rhythms in the SCN of rat pups sampled immedi- by measuring the phase of gene expression in the
ately after birth, which more closely reflects the SCN of pups immediately after birth; however,
phase of the fetal clock (El-Hennamy et al., this entrainment only occurs when the maternal
2008). This study explored the dynamics of fetal SCN clock is disturbed due to prolonged expo-
entrainment and demonstrated that several days sure to constant light (Novakova et al., 2010). In
elapse between the time when the maternal clock contrast, when the pregnant rats were maintained
sends its resetting signal and the time when the under a standard LD cycle, the restricted feeding
fetal clock entrains accordingly. As mentioned had no effect on the phase of gene expression
above, clock gene expression profiles exhibit profiles in the SCN of newborn rats. This result
low-amplitude rhythms at this developmental did not support the previous finding of Ohta
stage; therefore, rhythms in c-fos and Avp expres- et al. (2008), who demonstrated the effect of
sion were used as reliable markers of the activity maternal-restricted feeding on the phase of the
of the fetal SCN cells. rhythms in Per1:luciferase in the SCN slices from
The nature of the maternal cue(s) that entrain fetuses of transgenic rats maintained under the
the fetal clock has been studied for decades, LD cycle. It remains unclear whether differences
and melatonin and dopamine have been the ever- in the experimental design of these two studies
green candidates as previously mentioned. Mela- accounted for the different results. Nevertheless,
tonin is a plausible molecule for many reasons, as the restricted feeding entrained the in vivo
and studies that provide evidence supporting fetal clock only when the maternal SCN was
or excluding its role have been previously absent or nonfunctional, the outcome of these
summarized elsewhere (Bellavia et al., 2006; studies questioned the relative entraining potency
Davis, 1997; Seron-Ferre et al., 2007). Dopamine of the signaling from the maternal SCN and
likely has the opposite effect than melatonin on restricted feeding to the fetal SCN. Apparently,
signaling pathways in the fetal SCN (Weaver the signaling from the feeding regime did not
et al., 1992). Thus, melatonin is thought to medi- override the signaling from the maternal SCN
ate the subjective night and dopamine the subjec- and may thus be redundant. Moreover, the results
tive day, regardless of when they are present (for of Novakova et al. (2010) provided evidence that
a review, see Weinert, 2005). In fact, this putative the entrainment of the fetal SCN is possible even
92

without melatonin. Because the pregnant rats undergoes substantial reorganization and func-
exposed to a restricted feeding regime were tional specialization that results in formation of
maintained under constant light during the entire various cell subpopulations. Because intercellular
pregnancy, their endogenous melatonin was likely coupling has been shown to have a significant role
suppressed, and melatonin could not serve as a in the clock mechanism (reviewed by Mohawk
messenger. However, this redundant mechanism and Takahashi, 2011), it would not be surprising
of fetal SCN entrainment by food remains if the rhythms in clock gene expression as
unknown. The redundancy of the mechanism also measured in the population of the SCN cells grad-
suggests that formal properties of the fetal SCN ually gain their robustness in parallel with the for-
clock more closely resemble the adult SCN than mation of the intercellular web. Indeed, the
the adult peripheral clocks. The only difference amplitudes of the rhythms in clock gene expres-
is in the way how information about the LD cycle sion in the rat SCN gradually increased up to
is conveyed to the adult or fetal SCN. While pho- P10, when they attained an adult-like state
tic cues directly delivered from the retina are (Kovacikova et al., 2006; Sladek et al., 2004).
dominant for the entrainment of the adult SCN, Interestingly also in vitro, the amplitudes of
the information about the LD cycle is transduced Bmal1:luciferase rhythms in SCN slices from
to the fetal SCN by the putative signaling from mouse pups at P6 were much smaller than those
the maternal SCN. in SCN slices from adult mice (Nishide et al.,
2008). Although both of these processes are
remarkably correlated, there is no direct evidence
Development of the circadian system under thus far that they are mutually conditional.
maternal care

SCN develops postnatally Maternal entrainment of the SCN clock


after birth
With the postnatal progression of synaptogenesis
in the SCN (Moore, 1991), the coupling among Entrainment of a pup’s SCN with the external LD
individual cellular oscillators is strengthened and cycle already occurs during the prenatal period
reaches an adult-like state at P10. At the same through maternal cues (see above). The pups
time, the number of astrocytes in the SCN rapidly are thus born with their SCN clocks already
increases (Munekawa et al., 2000). Moreover, the set to the proper phase, and during the early post-
intrinsically photosensitive retinal ganglion cells natal period, they only need to maintain the
that mediate photic information from the retina phase until their own SCN clock becomes fully
begin to innervate the SCN only during this entrainable by light. The postnatal maternal
period (McNeill et al., 2011). Although this inner- entrainment gradually loses its significance
vation occurs postnatally, the axons of those cells (Reppert and Schwartz, 1986; Shimoda et al.,
are already present in the optic chiasm during the 1986; Viswanathan, 1999) with the increasing
fetal stage. The terminals of the intrinsically pho- probability of incidental exposure of pups to
tosensitive retinal ganglion cells gradually spread external light. In rats, this might happen after
throughout the retinorecipient zone of the SCN, about P10 when they begin to move around the
and the dynamics of this process spatially and nest in the dark burrow even though their eyes
temporally correlate with the pattern of c-fos pho- are opened only around P15 (see below). During
toinduction in the mouse and rat SCN (Mateju the period when the pups are in darkness in the
et al., 2009; Sekaran et al., 2005). Therefore, dur- nest, they are synchronized exclusively by the
ing the early postnatal period, the rodent SCN maternal circadian system. The mechanism of
93

how the mother entrains her pups is not yet related to the fact that during the early postnatal
understood, although various candidate pathways period, the SCN is under intensive structural
have been suggested. development when intercellular communication
During the period of lactation, the mother is formed and the cells become mutually coupled.
keeps behaving like a nocturnal animal and With creation of this web, the SCN cells might
leaves the nest during most of the night to forage lose their round-the-clock responsiveness, and
(Viswanathan, 1999). During the daytime, she the rhythmical maternal cues might become effec-
stays in the nest and takes care of her pups by tive only during a specific window of time similar
making contact with them and keeping them to mature animals.
warm and satiated. Thus, all these mutually cou- A more potent effect on the pup’s circadian
pled factors, such as social contact, temperature, system results from the deprivation of maternal
and food, are presented to the pups rhythmically care (Ohta et al., 2003). It has been shown in
and may entrain their circadian system. Melato- blind pups that are subjected to a complete
nin which is present in milk (Illnerova et al., absence of their mothers from their home cages
1993) and may thus theoretically provide the during the daytime from birth until P6. Due to
pup’s SCN with temporal information likely does this maternal deprivation, circadian rhythms in
not play such a role (Rowe and Kennaway, 2002). Per1 and Per2 expressions were completely
Only few studies have focused on the effect of phase-reversed in the SCN of these pups. This
these maternal factors on the molecular clock in experimental arrangement appears to have been
the SCN. It has been demonstrated that if a foster a much stronger entraining cue to the neonatal
mother rears pups that are blind since birth under SCN than the previously described cross-
the opposite LD cycle than the original mother, fostering. Under the cross-fostering experiment,
this has only marginal effect on the rhythms of a mother still provided care but it was reversed
Per1 and Per2 expression in the pups SCN at P6 in phase when compared with the pup’s clock.
(Ohta et al., 2002). This implies that the opposite However, under the deprivation experiment, the
phase of maternal care since birth was not able to pups were more extensively deprived of their
reprogram the pup’s SCN rhythmicity. These mother’s care because the mother was removed
results suggest that the role of maternal entrain- during the daytime when she would take care of
ment during the early postnatal period is only to the pups and returned at the nighttime when she
keep the phase of the clock that was already set would normally not take care of them. It can be
prenatally. However, it is not clear whether the expected that the nighttime care of pups was not
same result would be obtained under a natural sit- appropriate. Therefore, this latter condition was
uation where the mother needs to go a long dis- life-threatening for the pups which were thus
tance from her pups to chase for food and likely deprived from food, warmth, and social
would therefore need to leave them alone for a contact. Thus, it is not surprising that such a situ-
much longer time than in laboratory conditions ation activated stress-related gene expression in
where they remain in the same cage and the food the pups (Ohta et al., 2003). Whether the activa-
is available ad libitum. Nevertheless, the maternal tion of this pathway was responsible for the phase
entrainment, though likely of low potency, is reversal of the pup’s clock remains to be deter-
present during the early postnatal period because mined, but data from studies in adults suggest
the pups do not free-run but remain synchronized the feasibility of such a possibility (Reddy et al.,
with their mothers. The inability of the circadian 2007; Yamamoto et al., 2005).
clock to reverse its phase due to maternal care The effect of experimentally induced temporal
provided in the opposite phase, as shown in the restriction of maternal care on the pup’s circadian
above-mentioned foster experiment, is likely system (as described above) resembles the situation
94

that naturally occurs in rabbits. During early (Cai et al., 2009). The data demonstrated that at
development, the rabbit pups live in complete P3, none of the measured clock genes (Bmal1,
darkness in subterranean burrows. In contrast to Clock, Npas2, Cry1, Per1, and Rev-erba)
rats, rabbit mothers visit their pups only once a exhibited a circadian rhythm. At P14, only Rev-
day to nurse them for a brief period of < 5 min erba and Npas2 expression was driven rhythmi-
(Zarrow et al., 1965). It is vitally important for cally, and the rhythms of all these clock genes
the pups not to miss their chance to be fed, and were at least borderline significant only at P60.
therefore, they present a behavioral arousal in In peripheral organs, development of clock gene
anticipation of the feeding. The time of feeding expression profiles was studied in detail in both
is a strong entraining cue for the rabbit pups, the liver and the heart; information about other
and it can entrain many physiological variables tissues is sparse. In the rat heart and liver, the
such as core body temperature and hormonal, rhythms of all clock genes studied were expressed
metabolic, and neural functions independent of only around P10 and P20, respectively (Sakamoto
the time of the day (Caba and Gonzalez-Mariscal, et al., 2002; Sladek et al., 2007). The only excep-
2009). However, similar to the previously de- tion was the rhythm in Rev-erba expression which
scribed foster mother experiment in rats (Ohta was present in the liver since the embryonic stage
et al., 2002), only a small shift in the rhythm in (Sladek et al., 2007). However, the adult-like
the SCN (as measured by PER1 immunoreactiv- waveform and amplitude of the rhythms were
ity) was found in pups that were fed during present in most of the clock genes in the liver only
the day when compared with those that were fed around weaning (Sladek et al., 2007). Studies
during the night (Caba et al., 2008). Therefore, employing transgenic rats and measuring the
the feeding-induced arousal in rabbit pups was Per1 expression in vitro revealed that circadian
entrained by a food-entrainable oscillator rather rhythms in the expression of this gene were pres-
than by the central clock in the SCN. ent in pineal, liver, and thyroid tissues since birth.
It seems that maternal care attains a significant In the adrenals, the rhythm was present but
efficiency as an entraining cue for the pup’s circa- increased its amplitude until P25, and in the lungs,
dian system only under such conditions when it the rhythm was absent until P12 (Yamazaki et al.,
becomes restricted and chancy. Therefore, it is 2009). Thus, the lung does not seem to be capable
highly probable that more than one pathway is of expressing an autonomous rhythm in Per1:
involved in maternal entrainment and the mecha- luciferase until P10, although the expression is
nism is likely very complex. transiently inducible by addition of forskolin or
serum to the incubation medium, and all other
genes of the clockwork are present (Yamazaki
Maternal entrainment of the peripheral clocks et al., 2009). However, the authors do not report
after birth whether the other clock genes were expressed
rhythmically in the “arrhythmic” lungs. This pos-
During the postnatal period, molecular clock sibility arises from in vivo data published so far
mechanism seems to mature at different rates in by several groups that clearly demonstrates that
various peripheral tissues. Evidence for this the- the rhythmic expression of each individual clock
ory comes from both in vivo and in vitro studies genes may not occur at the same time during
that share some similar, but not identical, out- development (Sakamoto et al., 2002; Sladek
comes. In studies using an in vivo approach, it et al., 2007). These results are difficult to explain
was shown that the rhythms in most clock in the context of the current view on the mecha-
genes started gradually during postnatal develop- nism of how the molecular clockwork operates
ment. In the brain, ontogenesis of the mouse and could merely reflect differences in assay sen-
striatal molecular clock machinery was studied sitivity for various genes.
95

Both in vivo and in vitro studies confirmed that, pups begin to move independently in the burrow
during development, the phase of the clock gene and the probability of their light exposure in-
expression rhythms in some tissues gradually creases. In rats, photic entrainment was indi-
shifted until the adult stage was achieved. The cated already at P6 (Duncan et al., 1986; Weaver
shifts in the phases of the rhythms were detected and Reppert, 1995). The development of this
in the heart (Sakamoto et al., 2002), liver (Sladek mechanism is not completely understood. Mor-
et al., 2007; Yamazaki et al., 2009), and pineal phologically, the neuronal projection from the ret-
gland (Yamazaki et al., 2009). These data indicate ina to the SCN develops gradually during the
that the circadian clocks in these tissues are postnatal period. At P1, the innervation termina-
entrained by maternal cues likely related to tes close to the ventral border of the SCN;
maternal behavior and breast feeding. They likely later, the area gradually extends and achieves an
reflect the fact that during lactation, the newborn adult-like pattern at around P10 (Speh and Moore,
pup of a nocturnal animal is fed as a diurnal ani- 1993). Although it is undeveloped, this innervation
mal (see above), and the “nocturnality” appears is already functional because both the retina and
with increasing independence from maternal milk the SCN are sensitive to light immediately after
during gradual weaning. Indeed, the circadian birth (Hannibal and Fahrenkrug, 2004; Leard
clock in the liver was completely reversed by et al., 1994; Mateju et al., 2009, 2010), despite the
P30 when weaning was complete. Importantly, fact that the eyes are closed until about P15 and
the SCN clock remains in the same phase during the light is thus perceived only through the eyelids.
this developmental period (Fig. 2). Thus, the The SCN is connected to the retina via ret-
selective sensitivity of peripheral tissues to food inohypothalamic tract, which is formed by pro-
that is observed in the adult (Damiola et al., cesses of a small subset of retinal ganglion cells
2000; Stokkan et al., 2001) is already present containing the photopigment melanopsin (Hattar
immediately after birth and likely represents the et al., 2002; Provencio et al., 2002), and projects
adaptation of the organism to the developmen- to the retinorecipient zone of the SCN. These cells
tally dependent timing of food availability. are already present within the retina at prenatal
In summary, the available data are in favor of stages in mice and rats; thus, they may transmit
the view that the synchronized rhythms of clock information about external lighting conditions dur-
gene expression develop earlier in the central ing the early postnatal period, that is, well before
SCN clock than in the peripheral oscillator. the visual photopigments maturate (Fahrenkrug
A putative role of the developing SCN in the syn- et al., 2004; Hannibal and Fahrenkrug, 2004;
chronization of the peripheral clocks during the Sekaran et al., 2005). Recently, it has been shown
early postnatal period remains to be determined. that while the image-forming brain targets are
Tissue-specific differences in the programming of innervated from the retinal ganglion cells prena-
the development of molecular oscillations have tally, the innervation from the intrinsically photo-
also been suggested. Such differences might help sensitive subset of the cells to the SCN is delayed
to elucidate specific roles of the individual compo- and proceeds into postnatal period (McNeill et al.,
nents of the molecular core clockwork in the SCN 2011). The dynamics of the innervation remarkably
and in peripheral organs during development. correlate with the spread of the retinorecipient area
within the rat SCN during the postnatal period
(Mateju et al., 2009) (Fig. 4). However, despite
LD cycle entrains the developing the high sensitivity to light, the rodent SCN is likely
clock in the SCN not able to be entrained by light during the first
days of life. The exact mechanism of how the func-
As previously mentioned, the mechanism of photic tionality develops is not yet known; although it
entrainment of the SCN likely develops once the likely depends on the development of the SCN, a
96

P1 P3 P5 P10

c-fos

Per1

Shell
Per2 Core
OC

Fig. 4. The development of photosensitivity in the rat SCN. In situ hybridization emulsion autoradiographs demonstrating spatio-
temporal distribution of c-fos, Per1, and Per2 expression in SCN sections of rat pups exposed to a 30-min light pulse during
subjective night at postnatal day (P)1, P3, P5, and P10 (for more data, see Mateju et al., 2009).

role for retinal development cannot be excluded. day by induction of Per1 gene expression; thus,
The development of photic entrainment is difficult the response in the SCN is not yet gated by the
to follow in vivo because the ability to read a phase clock. The gating mechanism develops gradually
shift in response to light exposure requires using a between P3 and P5 but was only fully completed
reliable marker that reflects the functional state of at P10 (Mateju et al., 2009). Interestingly, the devel-
the clock both at the time of the light exposure opment of photoinducibility of Per2, its spatial dis-
and after the phase shift. However, most overt tribution, and its gating mechanism differed from
rhythms are not yet present at these early develop- Per1 (Mateju et al., 2009) and might thus be con-
mental stages. Also, dissection of the immediate trolled by a different mechanism from the begin-
response from the developmental change of the ning. The development of the mechanism that
variable is experimentally difficult. An insight into gates the photoinducibility of Per1 and Per2 to sub-
this process was provided in a study where develop- jective night depends on the development of the
ment of a mechanism gating the SCN photosensitiv- retinal innervation of the SCN, but it also likely
ity to a time-specific window was used as a marker depends on development of intercellular communi-
of photic entrainment (Mateju et al., 2009). In cation among the individual SCN oscillators (see
adults, light entrainment is accomplished so that above). It appears that this gating mechanism
light induces signaling pathways that are responsi- develops in parallel with the developmental stage
ble for induction of expression of the clock genes when population of SCN cells begins to express
that are involved in entrainment of the molecular Per1 and Per2 rhythmically, which likely occurred
clockwork, that is, Per1 and Per2 (for a review, because these cells attained synchrony (Mateju
see Meijer and Schwartz, 2003). Importantly, this et al., 2009). It seems that the development of the
response is possible only during a restricted window time-specific response to light goes hand-in-hand
of time when the endogenous expression of these with the decline in spontaneous levels of expression
genes is low, that is, during the subjective night during the night in control animals. Again, these
(Albrecht et al., 1997; Shearman et al., 1997; data suggest that for the SCN to fulfill the role of
Shigeyoshi et al., 1997). Mateju et al. (2009) demon- a central clock, postnatal maturation must extend
strated that immediately after birth, the SCN far beyond the development of the rhythmical con-
responds to light pulses delivered any time of the trol of activation of a clock gene promoter within
97

one cell. This view is even more obvious when we in an increased vulnerability to disruption by con-
consider the development of the mechanism by stant light (Cambras et al., 1997; Canal-Corretger
which the SCN codes the day length, that is, photo- et al., 2000; Ohta et al., 2006; Smith and Canal,
period. Apparently, photoperiodic modulation of 2009). However, the disrupted rhythms may be
the SCN is highly dependent on communication restored by exposure to LD cycles (Ohta et al.,
among individual cells and among populations of 2006). Similarly, the effect of pharmacological inter-
cells (Schaap et al., 2003; VanderLeest et al., vention of the circadian clock during the fetal/early
2007). In the adult rat SCN, the photoperiod neonatal stage was completely eliminated by
modulates markedly the rhythm of photosensitivity subsequent entrainment to a LD cycle (Lorelli
in the ventrolateral SCN (Sumova et al., 1995) and et al., 2011). These data suggest a remarkable resil-
rhythm of spontaneous neuronal activity in the ience of the central clock in the SCN to disturbances
dorsomedial SCN (Sumova et al., 2000). The photo- in the external environment during early develop-
period also modulates the waveform of the clock mental stages. In contrast, a recent report suggested
gene expression profiles as measured in a popula- that although the presence of seasonal photoperiods
tion of SCN cells within the medial part of the during perinatal period does not have a long-lasting
rostro-caudal axis (Sumova et al., 2003), and the effect on the network-level timing of neuronal
mutual phasing of these profiles measured in rhythms in the SCN, it might have a persistent effect
the rostral and caudal parts of the SCN (Hazlerigg on waveform and periods of individual SCN
et al., 2005; Sosniyenko et al., 2009, 2010). The neurons (Ciarleglio et al., 2007).
photoperiod begins to gradually modulate the Recent data suggest that changes in perinatal
waveforms of the clock gene expression profiles in nutrition may induce long-lasting alterations of
the medial part of the SCN from P10 until weaning the circadian control of food intake and metabo-
(Kovacikova et al., 2005). The development of pho- lism. The effect of nutrient restriction during this
toperiodic modulation of the mutual phasing period had a significant effect on the circadian
among the rostral and caudal parts of the SCN has expression of genes regulating food intake in the
not been studied. hypothalamus and on energy metabolism in
the liver, which persisted in rats after weaning
(Orozco-Solis et al., 2010). These data indicate that
Effect of disturbing the circadian system during the circadian system undergoes nutritional pro-
early ontogenesis may persist through adulthood gramming during perinatal development and that
such programming may contribute to the develop-
From data summarized above, it appears that the ment of metabolic disorders in adulthood. Vice
circadian system is sensitive to disturbances in the versa, disruption of circadian rhythmicity during
external environment during early developmental gestation resulted in metabolic consequences for
stage; no matter whether they are related to adult offspring (Varcoe et al., 2011). Because these
changes in LD cycles or to the feeding regime. data are likely translatable to human medicine,
It is still unclear whether these changes may have they reinforce the importance of maternal nutri-
long-lasting effects on the organism and whether tion together with integrity of the circadian system
they may persist till adulthood. during the perinatal period.
Some results indicate that light experience during
the early postnatal period may affect clock function
and clock output in adulthood. For example, expos- Conclusion
ing pups to abnormal lighting conditions, such as
constant light during lactation, has both acute and Studies focused on unraveling the basic prin-
lasting effects on circadian organization that result ciples of circadian clock development during
98

early ontogenesis have uncovered an enormous Cry2 clock gene Cryptochrome 2


plasticity of the whole system, which seems E embryonic day
dependent on the development of its individual LD light/dark
bricks of which it is built. These studies also point SCN suprachiasmatic nucleus
out the necessity of a detailed understanding of P postnatal day
the processes by which the circadian system Per1 clock gene Period 1
develops. This importance is underscored by the Per2 clock gene Period 2
findings that intervention of the system by various Rev- clock gene Reverse erb A alpha
cues during development may permanently impact erba
the temporal control of physiological functions Rora clock gene RAR-related orphan
in adulthood. Such interventions become more receptor A
and more frequent in the modern lifestyle that Vip vasoactive intestinal peptide
has been adopted during the past decades. More-
over, pregnant women often are not excluded from
the socio-economic pressure that underlies some of References
these interventions. It has been widely recognized
Akhtar, R. A., Reddy, A. B., Maywood, E. S., Clayton, J. D.,
that the circadian system is involved in the tempo- King, V. M., Smith, A. G., et al. (2002). Circadian cycling of
ral control of most physiological functions and the mouse liver transcriptome, as revealed by cDNA micro-
that distortion of the temporal regulation may array, is driven by the suprachiasmatic nucleus. Current
promote various diseases including cancer and psy- Biology, 12, 540–550.
chiatric, metabolic, and cardiovascular disorders. Albrecht, U., Sun, Z. S., Eichele, G., & Lee, C. C. (1997).
A differential response of two putative mammalian circa-
Therefore, understanding how interventions dur- dian regulators, mper1 and mper2, to light. Cell, 91,
ing the perinatal period impact on the circadian 1055–1064.
system may help in the prevention and treatment Amano, T., Matsushita, A., Hatanaka, Y., Watanabe, T.,
of many disorders. Oishi, K., Ishida, N., et al. (2009). Expression and functional
analyses of circadian genes in mouse oocytes and preimplan-
tation embryos: Cry1 is involved in the meiotic process inde-
Acknowledgments pendently of circadian clock regulation. Biology of
Reproduction, 80, 473–483.
Ansari, N., Agathagelidis, M., Lee, C., Korf, H. W., & von
The study was supported by grant nos. 305/09/ Gall, C. (2009). Differential maturation of circadian rhythms
0321, P303/11/0668, P303/12/1108, NT11474-4/ in clock gene proteins in the suprachiasmatic nucleus and
2010, and LC554, and by Research Project the pars tuberalis during mouse ontogeny. The European
AV0Z 50110509. Journal of Neuroscience, 29, 477–489.
Bellavia, S. L., Carpentieri, A. R., Vaque, A. M.,
Macchione, A. F., & Vermouth, N. T. (2006). Pup circadian
Abbreviations rhythm entrainment—Effect of maternal ganglionectomy or
pinealectomy. Physiology and Behavior, 89, 342–349.
Brown, T. M., & Piggins, H. D. (2009). Spatiotemporal hetero-
Avp arginine vasopressin geneity in the electrical activity of suprachiasmatic nuclei
Bmal1 clock gene Brain and muscle Arnt- neurons and their response to photoperiod. Journal of
like protein Biological Rhythms, 24, 44–54.
Clock clock gene Circadian Locomotor Caba, M., & Gonzalez-Mariscal, G. (2009). The rabbit pup, a
natural model of nursing-anticipatory activity. The Euro-
Output Cycles Kaput pean Journal of Neuroscience, 30, 1697–1706.
CREB cyclic AMP response element-binding Caba, M., Tovar, A., Silver, R., Mogado, E., Meza, E.,
protein Zavaleta, Y., et al. (2008). Nature’s food anticipatory exper-
Cry1 clock gene Cryptochrome 1 iment: Entrainment of locomotor behavior, suprachiasmatic
99

and dorsomedial hypothalamic nuclei by suckling in rabbit suprachiasmatic nucleus: A study using the 2-deoxy[1-14C]
pups. The European Journal of Neuroscience, 27, 432–443. glucose method. Proceedings of the National Academy of
Cai, Y., Liu, S., Li, N., Xu, S., Zhang, Y., & Chan, P. (2009). Sciences of the United States of America, 77, 1204–1208.
Postnatal ontogenesis of molecular clock in mouse striatum. Ginty, D. D., Kornhauser, J. M., Thompson, M. A.,
Brain Research, 1264, 33–38. Bading, H., Mayo, K. E., Takahashi, J. S., et al. (1993). Reg-
Cambras, T., Canal, M. M., Torres, A., Vilaplana, J., & Diez- ulation of CREB phosphorylation in the suprachiasmatic
Noguera, A. (1997). Manifestation of circadian rhythm under nucleus by light and a circadian clock. Science, 260, 238–241.
constant light depends on lighting conditions during lactation. Gotz, M., & Sommer, L. (2005). Cortical development: The art
The American Journal of Physiology, 272, R1039–R1046. of generating cell diversity. Development, 132, 3327–3332.
Canal-Corretger, M. M., Cambras, T., Vilaplana, J., & Diez- Hahm, S. H., & Eiden, L. E. (1998). Five discrete cis-active
Noguera, A. (2000). Bright light during lactation alters the domains direct cell type-specific transcription of the vasoac-
functioning of the circadian system of adult rats. American tive intestinal peptide (VIP) gene. The Journal of Biological
Journal of Physiology. Regulatory, Integrative and Compara- Chemistry, 273, 17086–17094.
tive Physiology, 278, R201–R208. Hannibal, J., & Fahrenkrug, J. (2004). Melanopsin containing
Cascio, S., & Zaret, K. S. (1991). Hepatocyte differentiation retinal ganglion cells are light responsive from birth.
initiates during endodermal-mesenchymal interactions prior Neuroreport, 15, 2317–2320.
to liver formation. Development, 113, 217–225. Hastings, M. H. (1997). Circadian clocks. Current Biology, 7,
Ciarleglio, C. M., Axley, J. C., Strauss, B. R., Gamble, K. L., & R670–R672.
McMahon, D. G. (2007). Perinatal photoperiod imprints the Hattar, S., Liao, H. W., Takao, M., Berson, D. M., &
circadian clock. Nature Neuroscience, 14, 25–27. Yau, K. W. (2002). Melanopsin-containing retinal ganglion
Dabeva, M. D., Petkov, P. M., Sandhu, J., Oren, R., cells: Architecture, projections, and intrinsic photosensitiv-
Laconi, E., Hurston, E., et al. (2000). Proliferation and dif- ity. Science, 295, 1065–1070.
ferentiation of fetal liver epithelial progenitor cells after Hazlerigg, D. G., Ebling, F. J., & Johnston, J. D. (2005).
transplantation into adult rat liver. American Journal of Photoperiod differentially regulates gene expression rhythms
Pathology, 156, 2017–2031. in the rostral and caudal SCN. Current Biology, 15,
Damiola, F., Le Minh, N., Preitner, N., Kornmann, B., Fleury- R449–R450.
Olela, F., & Schibler, U. (2000). Restricted feeding Illnerova, H., Buresova, M., & Presl, J. (1993). Melatonin
uncouples circadian oscillators in peripheral tissues from rhythm in human milk. The Journal of Clinical Endocrinol-
the central pacemaker in the suprachiasmatic nucleus. Genes ogy and Metabolism, 77, 838–841.
& Development, 14, 2950–2961. Iwasaki, Y., Oiso, Y., Saito, H., & Majzoub, J. A. (1997). Posi-
Davis, F. C. (1997). Melatonin: Role in development. Journal tive and negative regulation of the rat vasopressin gene pro-
of Biological Rhythms, 12, 498–508. moter. Endocrinology, 138, 5266–5274.
Davis, F. C., & Gorski, R. A. (1988). Development of hamster Jin, X., Shearman, L. P., Weaver, D. R., Zylka, M. J.,
circadian rhythms: Role of the maternal suprachiasmatic de Vries, G. J., & Reppert, S. M. (1999). A molecular mech-
nucleus. Journal of Comparative Physiology. A, 162, anism regulating rhythmic output from the suprachiasmatic
601–610. circadian clock. Cell, 96, 57–68.
Dolatshad, H., Cary, A. J., & Davis, F. C. (2010). Differential Johnson, M. H., Lim, A., Fernando, D., & Day, M. L. (2002).
expression of the circadian clock in maternal and embryonic Circadian clockwork genes are expressed in the reproduc-
tissues of mice. PLoS One, 5, e9855. tive tract and conceptus of the early pregnant mouse. Repro-
Duncan, M. J., Banister, M. J., & Reppert, S. M. (1986). ductive Biomedicine Online, 4, 140–145.
Developmental appearance of light–dark entrainment in Jud, C., & Albrecht, U. (2006). Circadian rhythms in murine
the rat. Brain Research, 369, 326–330. pups develop in absence of a functional maternal circadian
Dzierzak, E., & Medvinsky, A. (1995). Mouse embryonic clock. Journal of Biological Rhythms, 21, 149–154.
hematopoiesis. Trends in Genetics, 11, 359–366. Klein, D. C., Moore, R. Y., & Reppert, S. M. (1991).
El-Hennamy, R., Mateju, K., Bendova, Z., Sosniyenko, S., & Suprachiasmatic nucleus: The mind’s clock. New York, NY:
Sumova, A. (2008). Maternal control of the fetal and neona- Oxford University Press.
tal rat suprachiasmatic nucleus. Journal of Biological Ko, M. S., Kitchen, J. R., Wang, X., Threat, T. A.,
Rhythms, 23, 435–444. Hasegawa, A., Sun, T., et al. (2000). Large-scale cDNA
Fahrenkrug, J., Nielsen, H. S., & Hannibal, J. (2004). Expres- analysis reveals phased gene expression patterns during pre-
sion of melanopsin during development of the rat retina. implantation mouse development. Development, 127,
Neuroreport, 15, 781–784. 1737–1749.
Fuchs, J. L., & Moore, R. Y. (1980). Development of circadian Ko, C. H., Yamada, Y. R., Welsh, D. K., Buhr, E. D.,
rhythmicity and light responsiveness in the rat Liu, A. C., Zhang, E. E., et al. (2010). Emergence of
100

noise-induced oscillations in the central circadian pace- Moore, R. Y., & Bernstein, M. E. (1989). Synaptogenesis in
maker. PLoS Biology, 8, e1000513. the rat suprachiasmatic nucleus demonstrated by electron
Kovacikova, Z., Sladek, M., Bendova, Z., Illnerova, H., & microscopy and synapsin I immunoreactivity. The Journal
Sumova, A. (2006). Expression of clock and clock-driven of Neuroscience, 9, 2151–2162.
genes in the rat suprachiasmatic nucleus during late fetal Morin, L. P. (2007). SCN organization reconsidered. Journal
and early postnatal development. Journal of Biological of Biological Rhythms, 22, 3–13.
Rhythms, 21, 140–148. Munekawa, K., Tamada, Y., Iijima, N., Hayashi, S.,
Kovacikova, Z., Sladek, M., Laurinova, K., Bendova, Z., Ishihara, A., Inoue, K., et al. (2000). Development of
Illnerova, H., & Sumova, A. (2005). Ontogenesis of photope- astroglial elements in the suprachiasmatic nucleus of the
riodic entrainment of the molecular core clockwork in the rat rat: With special reference to the involvement of the optic
suprachiasmatic nucleus. Brain Research, 1064, 83–89. nerve. Experimental Neurology, 166, 44–51.
Kowalska, E., Moriggi, E., Bauer, C., Dibner, C., & Nishide, S. Y., Honma, S., & Honma, K. (2008). The circadian
Brown, S. A. (2010). The circadian clock starts ticking at a pacemaker in the cultured suprachiasmatic nucleus from
developmentally early stage. Journal of Biological Rhythms, pup mice is highly sensitive to external perturbation. The
25, 442–449. European Journal of Neuroscience, 27, 2686–2690.
Leard, L. E., Macdonald, E. S., Heller, H. C., & Kilduff, T. S. Novakova, M., Sladek, M., & Sumova, A. (2010). Exposure of
(1994). Ontogeny of photic-induced c-fos mRNA expression pregnant rats to restricted feeding schedule synchronizes the
in rat suprachiasmatic nuclei. Neuroreport, 5, 2683–2687. SCN clocks of their fetuses under constant light but not
Li, X., & Davis, F. C. (2005). Developmental expression of under a light–dark regime. Journal of Biological Rhythms,
clock genes in the Syrian hamster. Brain Research. Develop- 25, 350–360.
mental Brain Research, 158, 31–40. O’Neill, J. S., Maywood, E. S., Chesham, J. E., Takahashi, J. S.,
Liu, A. C., Welsh, D. K., Ko, C. H., Tran, H. G., Zhang, E. E., & Hastings, M. H. (2008). cAMP-dependent signaling as
Priest, A. A., et al. (2007). Intercellular coupling confers a core component of the mammalian circadian pacemaker.
robustness against mutations in the SCN circadian clock net- Science, 320, 949–953.
work. Cell, 129, 605–616. Obrietan, K., Impey, S., Smith, D., Athos, J., & Storm, D. R.
Lorelli, C. J., Wreschnig, D., & Davis, F. C. (2011). Resilience (1999). Circadian regulation of cAMP response element-
of circadian pacemaker development in hamsters. Journal of mediated gene expression in the suprachiasmatic nuclei.
Biological Rhythms, 26, 221–229. The Journal of Biological Chemistry, 274, 17748–17756.
Mateju, K., Bendova, Z., El-Hennamy, R., Sladek, M., Ohta, H., Honma, S., Abe, H., & Honma, K. (2002). Effects of
Sosniyenko, S., & Sumova, A. (2009). Development of the nursing mothers on rPer1 and rPer2 circadian expressions in
light sensitivity of the clock genes Period1 and Period2, the neonatal rat suprachiasmatic nuclei vary with develop-
and immediate-early gene c-fos within the rat mental stage. The European Journal of Neuroscience, 15,
suprachiasmatic nucleus. The European Journal of Neurosci- 1953–1960.
ence, 29, 490–501. Ohta, H., Honma, S., Abe, H., & Honma, K. (2003). Periodic
Mateju, K., Sumova, A., & Bendova, Z. (2010). Expression absence of nursing mothers phase-shifts circadian rhythms
and light sensitivity of clock genes Per1 and Per2 and of clock genes in the suprachiasmatic nucleus of rat pups.
immediate-early gene c-fos within the retina of early postna- The European Journal of Neuroscience, 17, 1628–1634.
tal Wistar rats. The Journal of Comparative Neurology, 518, Ohta, H., Mitchell, A. C., & McMahon, D. G. (2006). Constant
3630–3644. light disrupts the developing mouse biological clock. Pediat-
McNeill, D. S., Sheely, C. J., Ecker, J. L., Badea, T. C., ric Research, 60, 304–308.
Morhardt, D., Guido, W., et al. (2011). Development of Ohta, H., Xu, S., Moriya, T., Iigo, M., Watanabe, T.,
melanopsin-based irradiance detecting circuitry. Neural Nakahata, N., et al. (2008). Maternal feeding controls fetal
Development, 6, 8–18. biological clock. PLoS One, 3, e2601.
Meijer, J. H., & Schwartz, W. J. (2003). In search of the pathways Orozco-Solis, R., Matos, R. J., Guzman-Quevedo, O., Lopes
for light-induced pacemaker resetting in the suprachiasmatic de Souza, S., Bihouee, A., Houlgatte, R., et al. (2010).
nucleus. Journal of Biological Rhythms, 18, 235–249. Nutritional programming in the rat is linked to long-lasting
Mohawk, J. A., & Takahashi, J. S. (2011). Cell autonomy and changes in nutrient sensing and energy homeostasis in the
synchrony of suprachiasmatic nucleus circadian oscillators. hypothalamus. PLoS One, 5, e13537.
Trends in Neurosciences, 34, 349–358. Pando, M. P., Morse, D., Cermakian, N., & Sassone-Corsi, P.
Moore, R. Y. (1991). Development of the suprachiasmatic (2002). Phenotypic rescue of a peripheral clock genetic defect
nucleus. In D. C. Klein, R. Y. Moore & S. M. Reppert via SCN hierarchical dominance. Cell, 110, 107–117.
(Eds.), Suprachiasmatic nucleus: The mind’s clock Poirel, V. J., Boggio, V., Dardente, H., Pevet, P., Masson-
(pp. 197–216). New York, NY: Oxford University Press. Pevet, M., & Gauer, F. (2003). Contrary to other non-photic
101

cues, acute melatonin injection does not induce immediate homologs: Circadian expression and photic regulation in the
changes of clock gene mRNA expression in the rat suprachiasmatic nuclei. Neuron, 19, 1261–1269.
suprachiasmatic nuclei. Neuroscience, 120, 745–755. Shibata, S., & Moore, R. Y. (1988). Development of a fetal cir-
Provencio, I., Rollag, M. D., & Castrucci, A. M. (2002). Photo- cadian rhythm after disruption of the maternal circadian sys-
receptive net in the mammalian retina. This mesh of cells tem. Brain Research, 469, 313–317.
may explain how some blind mice can still tell day from Shigeyoshi, Y., Taguchi, K., Yamamoto, S., Takekida, S.,
night. Nature, 415, 493. Yan, L., Tei, H., et al. (1997). Light-induced resetting of a
Reddy, A. B., Maywood, E. S., Karp, N. A., King, V. M., mammalian circadian clock is associated with rapid induc-
Inoue, Y., Gonzalez, F. J., et al. (2007). Glucocorticoid tion of the mPer1 transcript. Cell, 91, 1043–1053.
signaling synchronizes the liver circadian transcriptome. Shimoda, K., Hanada, K., Yamada, N., Takahashi, K., &
Hepatology, 45, 1478–1488. Takahashi, S. (1986). Periodic exposure to mother is potent
Reppert, S. M., & Schwartz, W. J. (1983). Maternal coordina- zeitgeber of rat pups’ rhythm. Physiology and Behavior, 36,
tion of the fetal biological clock in utero. Science, 220, 723–730.
969–971. Shiojiri, N., & Mizuno, T. (1993). Differentiation of functional
Reppert, S. M., & Schwartz, W. J. (1984). The suprachiasmatic hepatocytes and biliary epithelial cells from immature
nuclei of the fetal rat: Characterization of a functional circa- hepatocytes of the fetal mouse in vitro. Anatomy and
dian clock using 14C-labeled deoxyglucose. The Journal of Embryology (Berlin), 187, 221–229.
Neuroscience, 4, 1677–1682. Sladek, M., Jindrakova, Z., Bendova, Z., & Sumova, A.
Reppert, S. M., & Schwartz, W. J. (1986). Maternal (2007). Postnatal ontogenesis of the circadian clock within
suprachiasmatic nuclei are necessary for maternal coordina- the rat liver. American Journal of Physiology. Regulatory,
tion of the developing circadian system. The Journal of Integrative and Comparative Physiology, 292,
Neuroscience, 6, 2724–2729. R1224–R1229.
Reppert, S. M., & Uhl, G. R. (1987). Vasopressin messenger Sladek, M., Sumova, A., Kovacikova, Z., Bendova, Z.,
ribonucleic acid in supraoptic and suprachiasmatic nuclei: Laurinova, K., & Illnerova, H. (2004). Insight into molecular
Appearance and circadian regulation during development. core clock mechanism of embryonic and early postnatal
Endocrinology, 120, 2483–2487. rat suprachiasmatic nucleus. Proceedings of the National Acad-
Rowe, S. A., & Kennaway, D. J. (2002). Melatonin in rat milk emy of Sciences of the United States of America, 101, 6231–6236.
and the likelihood of its role in postnatal maternal entrain- Smith, L., & Canal, M. M. (2009). Expression of circadian
ment of rhythms. American Journal of Physiology. Regu- neuropeptides in the hypothalamus of adult mice is affected
latory, Integrative and Comparative Physiology, 282, by postnatal light experience. Journal of Neuroendocrinology,
R797–R804. 21, 946–953.
Sakamoto, K., Oishi, K., Nagase, T., Miyazaki, K., & Ishida, N. Sosniyenko, S., Hut, R. A., Daan, S., & Sumová, A. (2009).
(2002). Circadian expression of clock genes during ontogeny Influence of photoperiod duration and light–dark trans-
in the rat heart. Neuroreport, 13, 1239–1242. itions on entrainment of Per1 and Per2 gene and protein
Schaap, J., Albus, H., VanderLeest, H. T., Eilers, P. H., expression in subdivisions of the mouse suprachiasmatic
Detari, L., & Meijer, J. H. (2003). Heterogeneity of nucleus. The European Journal of Neuroscience, 30,
rhythmic suprachiasmatic nucleus neurons: Implications 1802–1814.
for circadian waveform and photoperiodic encoding. Pro- Sosniyenko, S., Parkanova, D., Illnerova, H., Sladek, M., &
ceedings of the National Academy of Sciences of the United Sumova, A. (2010). Different mechanisms of adjustment to
States of America, 100, 15994–15999. a change of the photoperiod in the suprachiasmatic and liver
Schibler, U., Ripperger, J., & Brown, S. A. (2003). Peripheral circadian clocks. American Journal of Physiology. Regu-
circadian oscillators in mammals: Time and food. Journal of latory, Integrative and Comparative Physiology, 298,
Biological Rhythms, 18, 250–260. R959–R971.
Sekaran, S., Lupi, D., Jones, S. L., Sheely, C. J., Hattar, S., Speh, J. C., & Moore, R. Y. (1993). Retinohypothalamic tract
Yau, K. W., et al. (2005). Melanopsin-dependent photore- development in the hamster and rat. Brain Research. Devel-
ception provides earliest light detection in the mammalian opmental Brain Research, 76, 171–181.
retina. Current Biology, 15, 1099–1107. Staud, F., Mazancova, K., Miksik, I., Pavek, P., Fendrich, Z., &
Seron-Ferre, M., Valenzuela, G. J., & Torres-Farfan, C. Pacha, J. (2006). Corticosterone transfer and metabolism in
(2007). Circadian clocks during embryonic and fetal devel- the dually perfused rat placenta: Effect of 11beta-
opment. Birth Defects Research. Part C, Embryo Today: hydroxysteroid dehydrogenase type 2. Placenta, 27, 171–180.
Reviews, 81, 204–214. Stokkan, K. A., Yamazaki, S., Tei, H., Sakaki, Y., &
Shearman, L. P., Zylka, M. J., Weaver, D. R., Menaker, M. (2001). Entrainment of the circadian clock in
Kolakowski, L. F., Jr., & Reppert, S. M. (1997). Two period the liver by feeding. Science, 291, 490–493.
102

Sumova, A., Bendova, Z., Sladek, M., El-Hennamy, R., Viswanathan, N. (1999). Maternal entrainment in the circadian
Laurinova, K., Jindrakova, Z., et al. (2006a). Setting the activity rhythm of laboratory mouse (C57BL/6J). Physiology
biological time in central and peripheral clocks during onto- and Behavior, 68, 157–162.
genesis. FEBS Letters, 580, 2836–2842. Viswanathan, N., & Davis, F. C. (1997). Single prenatal
Sumova, A., Bendova, Z., Sladek, M., El-Hennamy, R., injections of melatonin or the D1-dopamine receptor agonist
Mateju, K., Polidarova, L., et al. (2008). Circadian molecular SKF 38393 to pregnant hamsters sets the offsprings’ circa-
clocks tick along ontogenesis. Physiological Research, 57 dian rhythms to phases 180 degrees apart. Journal of Com-
(Suppl. 3), S139–S148. parative Physiology. A, 180, 339–346.
Sumova, A., Bendova, Z., Sladek, M., Kovacikova, Z., Viswanathan, N., Weaver, D. R., Reppert, S. M., &
El-Hennamy, R., Laurinova, K., et al. (2006b). The rat circa- Davis, F. C. (1994). Entrainment of the fetal hamster circa-
dian clockwork and its photoperiodic entrainment during dian pacemaker by prenatal injections of the dopamine
development. Chronobiology International, 23, 237–243. agonist SKF 38393. The Journal of Neuroscience, 14,
Sumova, A., Jac, M., Sladek, M., Sauman, I., & Illnerova, H. 5393–5398.
(2003). Clock gene daily profiles and their phase relation- Walker, L., Carlson, A., Tan-Pertel, H. T., Weinmaster, G., &
ship in the rat suprachiasmatic nucleus are affected by pho- Gasson, J. (2001). The notch receptor and its ligands are
toperiod. Journal of Biological Rhythms, 18, 134–144. selectively expressed during hematopoietic development in
Sumova, A., Travnickova, Z., & Illnerova, H. (2000). Sponta- the mouse. Stem Cells, 19, 543–552.
neous c-Fos rhythm in the rat suprachiasmatic nucleus: Weaver, D. R., & Reppert, S. M. (1987). Maternal-fetal com-
Location and effect of photoperiod. American Journal of munication of circadian phase in a precocious rodent, the
Physiology. Regulatory, Integrative and Comparative Physi- spiny mouse. The American Journal of Physiology, 253,
ology, 279, R2262–R2269. E401–E409.
Sumova, A., Travnickova, Z., Peters, R., Schwartz, W. J., & Weaver, D. R., & Reppert, S. M. (1989). Periodic feeding of
Illnerova, H. (1995). The rat suprachiasmatic nucleus is a SCN-lesioned pregnant rats entrains the fetal biological clock.
clock for all seasons. Proceedings of the National Academy Brain Research. Developmental Brain Research, 46, 291–296.
of Sciences of the United States of America, 92, 7754–7758. Weaver, D. R., & Reppert, S. M. (1995). Definition of the
Takahashi, J. S., Hong, H. K., Ko, C. H., & McDearmon, E. L. developmental transition from dopaminergic to photic regu-
(2008). The genetics of mammalian circadian order and dis- lation of c-fos gene expression in the rat suprachiasmatic
order: Implications for physiology and disease. Nature nucleus. Brain Research. Molecular Brain Research, 33,
Reviews. Genetics, 9, 764–775. 136–148.
Torra, I. P., Tsibulsky, V., Delaunay, F., Saladin, R., Weaver, D. R., Rivkees, S. A., & Reppert, S. M. (1992). D1-
Laudet, V., Fruchart, J. C., et al. (2000). Circadian and glu- dopamine receptors activate c-fos expression in the fetal
cocorticoid regulation of Rev-erb alpha expression in liver. suprachiasmatic nuclei. Proceedings of the National Acad-
Endocrinology, 141, 3799–3806. emy of Sciences of the United States of America, 89,
Torres-Farfan, C., Rocco, V., Monso, C., Valenzuela, F. J., 9201–9204.
Campino, C., Germain, A., et al. (2006). Maternal melatonin Weinert, D. (2005). Ontogenetic development of the mamma-
effects on clock gene expression in a nonhuman primate lian circadian system. Chronobiology International, 22,
fetus. Endocrinology, 147, 4618–4626. 179–205.
Van Eyken, P., Sciot, R., Callea, F., Van der Steen, K., Welsh, D. K., Logothetis, D. E., Meister, M., & Reppert, S. M.
Moerman, P., & Desmet, V. J. (1988). The development of (1995). Individual neurons dissociated from rat
the intrahepatic bile ducts in man: A keratin-immunohisto- suprachiasmatic nucleus express independently phased cir-
chemical study. Hepatology, 8, 1586–1595. cadian firing rhythms. Neuron, 14, 697–706.
VanderLeest, H. T., Houben, T., Michel, S., Deboer, T., Welsh, D. K., Takahashi, J. S., & Kay, S. A. (2010).
Albus, H., Vansteensel, M. J., et al. (2007). Seasonal Suprachiasmatic nucleus: Cell autonomy and network pro-
encoding by the circadian pacemaker of the SCN. Current perties. Annual Review of Physiology, 72, 551–577.
Biology, 17, 468–473. Yagita, K., Horie, K., Koinuma, S., Nakamura, W.,
VanDunk, C., Hunter, L. A., & Gray, P. A. (2011). Develop- Yamanaka, I., Urasaki, A., et al. (2010). Development of
ment, maturation, and necessity of transcription factors in the circadian oscillator during differentiation of mouse
the mouse suprachiasmatic nucleus. The Journal of Neuro- embryonic stem cells in vitro. Proceedings of the National
science, 31, 6457–6467. Academy of Sciences of the United States of America, 107,
Vanecek, J. (1998). Cellular mechanisms of melatonin action. 3846–3851.
Physiological Reviews, 78, 687–721. Yagita, K., Tamanini, F., van Der Horst, G. T., & Okamura, H.
Varcoe, T. J., Wight, N., Voultsios, A., Salkeld, M. D., & (2001). Molecular mechanisms of the biological clock in
Kennaway, D. J. (2011). Chronic phase shifts of the photo- cultured fibroblasts. Science, 292, 278–281.
period throughout pregnancy programs glucose intolerance Yamamoto, T., Nakahata, Y., Tanaka, M., Yoshida, M.,
and insulin resistance in the rat. PLoS One, 6, e18504. Soma, H., Shinohara, K., et al. (2005). Acute physical stress
103

elevates mouse period1 mRNA expression in mouse periph- Yoshikawa, T., Yamazaki, S., & Menaker, M. (2005). Effects
eral tissues via a glucocorticoid-responsive element. The of preparation time on phase of cultured tissues reveal com-
Journal of Biological Chemistry, 280, 42036–42043. plexity of circadian organization. Journal of Biological
Yamazaki, S., Yoshikawa, T., Biscoe, E. W., Numano, R., Rhythms, 20, 500–512.
Gallaspy, L. M., Soulsby, S., et al. (2009). Ontogeny of circa- Zarrow, M. X., Denenberg, V. H., & Anderson, C. O. (1965).
dian organization in the rat. Journal of Biological Rhythms, Rabbit: Frequency of suckling in the pup. Science, 150,
24, 55–63. 1835–1836.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 6

When does it start ticking? Ontogenetic


development of the mammalian circadian system

Elmar Christ{,*, Horst-Werner Korf{,{ and Charlotte von Gall{,{,1

{
Dr. Senckenbergische Anatomie II, Fachbereich Medizin, Goethe-Universität Frankfurt,
Frankfurt am Main, Germany
{
Dr. Senckenbergisches Chronomedizinisches Institut, Goethe-Universität Frankfurt,
Frankfurt am Main, Germany

Abstract: Circadian rhythms in physiology and behavior ensure that vital functions are temporally
synchronized with cyclic environmental changes. In mammals, the circadian system is conducted by a
central circadian rhythm generator that resides in the hypothalamic suprachiasmatic nucleus (SCN)
and controls multiple subsidiary circadian oscillators in the periphery. The molecular clockwork in
SCN and peripheral oscillators consists of autoregulatory transcriptional/translational feedback loops
of clock genes. The adult circadian system is synchronized to the astrophysical day by light whereas
the fetal and neonatal circadian system entrains to nonphotic rhythmic maternal signals. This chapter
reviews maturation and entrainment of the central circadian rhythm generator in the SCN and of
peripheral oscillators during ontogenetic development.

Keywords: circadian clock; ontogenesis; development; suprachiasmatic nucleus; pars tuberalis; adrenal
gland; peripheral clock.

The circadian system and its molecular clockwork

Based on the rotation of the Earth around its axis,


*Corresponding author. the light conditions change rhythmically with a
Tel.: þ49-069-6301-83156; Fax: þ49-069-6301-6017 period length of 24 h (day). During phylogeny,
E-mail: E.Christ@med.uni-frankfurt.de circadian systems have evolved in nearly all living
organisms to anticipate these rhythmic changes in
1
Current address: Zentrum für Anatomie und Hirnforschung,
environmental light conditions (Dunlap, 1999;
Institut für Anatomie II, Universitätsklinikum Düsseldorf, Life Dvornyk et al., 2003; Johnson and Golden,
Science Center, Düsseldorf, Germany 1999) and to drive rhythms in physiology and

http://dx.doi.org/10.1016/B978-0-444-59427-3.00006-X 105
106

behavior. In mammals, the circadian system is Syrian hamster it is 22.5 h (Ralph and Menaker,
organized in a hierarchy of multiple oscillators. 1988), in laboratory mouse strains 23.6 h
The central circadian pacemaker is located in the (Schwartz and Zimmerman, 1990; von Gall et al.,
suprachiasmatic nucleus (SCN) of the anterior 1998), and in humans 25 h (Aschoff, 1965). Period
hypothalamus. The SCN integrates light informa- and phase of the circadian oscillation is entrained
tion conveyed by the retinohypothalamic tract to the external rhythm by so-called zeitgebers
(RHT) and coordinates peripheral oscillators (see below). The molecular clockwork controls
distributed throughout the body. In each SCN the rhythmic expression of clock-controlled genes
neuron, a molecular clockwork consisting of such as arginine vasopressin (AVP) (Jin et al.,
autoregulatory transcriptional/translational feed- 1999), an important rhythmic SCN output signal
back loops of clock genes drives rhythmic cellular (see below). Importantly, molecular clocks are
properties and rhythmic output signals with a not restricted to the SCN, but are present in
circadian period length ( 24 h) (Okamura et al., peripheral oscillators such as other brain regions
2002; Reppert and Weaver, 2002). The transcrip- and peripheral organs (Balsalobre, 2002;
tional activators in the core molecular clockwork, McNamara et al., 2001; Oishi et al., 1998; Sun
CLOCK/NPAS2 and BMAL1 are characterized et al., 1997; Yoo et al., 2004; Zylka et al., 1998).
by a basic helix–loop–helix DNA-binding domain Robust sustained oscillation of clock gene expres-
and two Per-Arnt-Sim protein interaction domains sion in the SCN and in peripheral oscillators
that allow for the formation of heterodimers. depends on intercellular coupling (Liu et al.,
These heterodimers activate gene expression 2007). Within the SCN paracrine signaling via
through E-box (like) enhancer elements located neuropeptides such as vasoactive intestinal peptide,
in the promoter region of the clock genes Per1, AVP, and gastrin-releasing peptide provides pow-
Per2, Cry1, and Cry2 and of so-called clock- erful interneuronal communication (Maywood
controlled genes. The protein products of the et al., 2011).
Per and Cry genes form repressor complexes
that comprise additional proteins such as casein
kinase 1 e and d. After translocation into the Synchronization mechanisms within the adult
nucleus, the repressor complex inhibits CLOCK/ circadian system
NPAS2:BMAL1-mediated transcription and thus
suppresses its own transcription. A new cycle In the adult circadian system, phase and period
starts after hyperphosphorylation, ubiquitination, length of the SCN molecular clockwork is
and proteasomal degradation of the repressor entrained to the environmental day/night cycle
complex. This core negative feedback loop is by light. This entrainment is essential for normal
modulated by accessory feedback loops that physiology because long-term internal temporal
involve the orphan nuclear receptors REV-ERBa desynchronization causes sleep disorders and
and RORa. By binding to ROR enhancer chronic illnesses, such as cardiovascular disease,
elements REV-ERBa and RORa control expres- metabolic syndrome, and cancer (Hastings et al.,
sion of Bmal1. As CLOCK and BMAL1 are con- 2003). Light during early or late night is a strong
stitutively present and bound to the E-box stimulus for delaying or advancing the phase of
element throughout the circadian cycle, the circa- the molecular clockwork, respectively (Reppert
dian rhythm in gene expression depends mainly and Weaver, 2002). Light information is transmit-
on the presence of the repressor complex (von ted to the SCN by glutamate and the neuropep-
Gall et al., 2003). The molecular clockwork genet- tide PACAP released from specialized retinal
ically determines the endogenous circadian period ganglion cells projecting into the SCN (Reppert
length that varies with the species, for example, in and Weaver, 2002). The activation of glutamate
107

receptors during night activates the transcription release by light as well as in the control of the circa-
factor Ca2 þ/cAMP-responsive element-binding dian rhythm in melatonin synthesis (Kalsbeek
protein (Gau et al., 2002), which activates the et al., 2006). Melatonin is an important rhythmic
expression of mPer1 and mPer2 in the SCN endocrine signal within the circadian system
(Albrecht et al., 1997; Gau et al., 2002; Shearman and can adjust the phase of the SCN rhythm gener-
et al., 1997). The photic induction of Per expres- ator (reviewed by von Gall et al., 2002b). Melato-
sion is crucial for the daily adjustment of the nin influences insulin production (reviewed by
endogenous rhythm of the molecular clockwork Peschke, 2008), modulates pain (reviewed by
because light during early night delays the rhythm Ambriz-Tututi et al., 2009), synchronizes slave
by activating Per in its declining phase, whereas oscillators such as the pars tuberalis (PT) of the
light during late night leads to a precocious anterior pituitary (Jilg et al., 2005; Messager et al.,
increase in Per expression, resulting in a phase 2000; Unfried et al., 2009, 2010; von Gall et al.,
advance (Hastings et al., 2003). 2001, 2005), and is believed to play a role in
The SCN transmits its phase and period to maintaining sleep throughout the night (reviewed
peripheral oscillators via paracrine, neuronal, and by Cajochen et al., 2003). In addition, the SCN con-
neuroendocrine output pathways. Behavioral trols sensitivity of peripheral glands to pituitary
rhythms such as circadian locomotor activity hormones (Buijs et al., 2003) and the circadian
depend on diffusible signals from the SCN such as rhythm in plasma glucose levels (Kalsbeek et al.,
transforming growth factor (Kramer et al., 2001), 2004) via the autonomic nervous system.
cardiotrophin-like cytokine (Kraves and Weitz, Neurons in the medial parvocellular PVH syn-
2006), and prokineticin 2 (Cheng et al., 2002; thesize corticotropin-releasing hormone (CRH)
LeSauter et al., 1996; Ralph et al., 1990; Silver which is secreted into the portal vasculature
et al., 1996). In contrast, the circadian control of of the hypophysis (Kalsbeek et al., 2004) and
neuroendocrine rhythms requires intact neuronal regulates the secretion of adrenal corticosteroids
projections from the SCN (Kalsbeek et al., 2000; via the hypothalamo-pituitary-adrenal (HPA)
LeSauter et al., 1996; Perreau-Lenz et al., 2003). gland axis (Buijs and Kalsbeek, 2001). Vaso-
Vasopressinergic SCN neurons provide an impor- pressinergic SCN output signals inhibit CRH
tant output pathway of the SCN: they project to release from the PVN and thus control the diur-
the paraventricular nucleus (PVN) of the hypothal- nal glucocorticoid (GC) rhythm (Kalsbeek et al.,
amus. The PVN is a prime center for the control of 1992; Tousson and Meissl, 2004). In mice with a
both the autonomic nervous system and the neuro- corrupted molecular clockwork, the regulation of
endocrine system. The autonomic nervous system the HPA axis is defective (Oster et al., 2006b).
that controls body homoeostasis originates from Light can affect GC release from the adrenal
the dorsal parvocellular portion of the PVN which by influencing either the HPA or the SCN-
contains presympathetic and preparasympathetic sympathetic nervous system (Ishida et al., 2005).
neurons (Kalsbeek et al., 2006; Vrang et al., Moreover, the molecular clockwork in the
1995). The presympathetic neurons send their adrenal gland gates GC production in response
axons to the intermediolateral column of the upper to adrenocorticotropin (ACTH) and controls
thoracic spinal cord, where they contact sympa- rhythmic expression of a variety of genes involved
thetic preganglionic neurons. Postganglionic nerve in corticosterone biosynthesis (Oster et al., 2006a,
fibers originating from the superior cervical ganglia b). GC levels show a strong daily oscillation in
regulate the nocturnal secretion of melatonin via both laboratory rodents and humans, and repre-
the rhythmic release of norepinephrine (reviewed sent an important phase entrainment signal within
by Klein et al., 1991). GABAergic neurons in the the circadian system (Balsalobre et al., 2000). In
SCN are involved in the inhibition of melatonin addition to rhythmic cues from the SCN, external
108

rhythmic cues such as food intake also provide a period: 20 days), a circadian rhythm in Avp
strong entrainment signal for peripheral mRNA expression is evident as early as E18
oscillators (Bur et al., 2010; Damiola et al., 2000; (Ansari et al., 2009). However, it is still a matter
Schibler et al., 2003; Shibata et al., 2010; Stephan, of debate whether Avp expression is rhythmic in
2002; Yamazaki et al., 2009). the fetal SCN as heteronuclear Avp RNA, which
is supposedly a more reliable marker of transcrip-
tion rate than mRNA, is not rhythmic until P1 in
Development of the central endogenous rat SCN (Kovacikova et al., 2006).
rhythm generator The expression of clock genes in the fetal SCN
is first detected at E13.5 in hamsters (mean gesta-
A complex and dynamic process of development tion period: 16 days), at E17 in mice, at E19 in
prepares the embryo for the living conditions rats, and at E142 in capuchin monkeys (gestation
which it will experience after birth. In mammals, period: 180 days). Thus, the molecular clockwork
the embryos and fetuses receive important mater- in the SCN starts to develop at around 90% of
nal signals for maturation and synchronization of gestation in these species. In hamster SCN, circa-
the fetal circadian system via the placental sys- dian rhythms in clock genes (Bmal1, Per1, and
tem. In the early postnatal period, the intense Cry1) occur for the first time around P0 and clear
social interaction with the mother as well as rhythms can be detected at P2 (Li and Davis, 2005).
hormones in the milk still provides rhythmic In rat SCN, a circadian rhythm in Per1 occurs
maternal signals for the developing circadian sys- already at E20, while circadian rhythms of Bmal1
tem. These maternal signals might directly drive and Per2 start at P1 and of Cry1 at P2 (Kovacikova
neonatal oscillators until the light input pathway et al., 2006; Sumova et al., 2004, 2006). In acute
and the SCN endogenous rhythm generator have SCN slice preparations from genetically modified
fully developed. rats bearing a Per1 promoter-driven luciferase
The SCN is formed during the third wave of (luc) reporter gene, a circadian rhythm in Per1
hypothalamic neuron generation and develops as expression was found to persist for at least 4 days
a component of the periventricular cell group. in vitro (Ohta et al., 2008). Clock gene proteins
In rats (mean gestation period: 22 days), SCN PER1, PER2, and CRY1 are undetectable in rat
neurons start to develop and differentiate SCN at E19 (Sumova et al., 2004). In the SCN of
between embryonic days 14–17 (E14–E17) and capuchin monkey, a circadian rhythm in Bmal1
the adult morphology of the SCN is reached and Per2 becomes apparent at E142 (Torres-
within postnatal day 10 (P10). The ventrolateral Farfan et al., 2006a). In mouse SCN, mPer1, and
subdivision of the SCN develops around mPer2 mRNAs are detectable at E17 (Shearman
E14–E16, prior to the dorsomedial subdivision et al., 1997), but only mPer1 mRNA levels show a
which develops around E16–E17 (Moore, 1991; significant circadian rhythm during this fetal stage
Moore and Leak, 2001; Weinert, 2005). Circadian (Shimomura et al., 2001). There is a small but sig-
rhythms in SCN activity in rodents can be nificant circadian variation in mPER1 and mPER2
detected already in utero. The fetal rat SCN immunoreaction in the SCN of C3H mice at E18,
shows circadian rhythms in metabolic activity coincident with a low-amplitude circadian rhythm
monitored by a 2-deoxyglucose uptake (Reppert in Avp mRNA levels (Ansari et al., 2009). How-
and Schwartz, 1984a,b), Avp mRNA expression ever, the rhythm of mCRY2 and mCRY1 levels
(Ansari et al., 2009; Reppert and Uhl, 1987), starts only after birth at P2 and P10, respectively
and neuronal firing rate between E19 and E22 (Ansari et al., 2009; Shearman et al., 1997; Shim-
(Shibata and Moore, 1987; Weinert, 2005). In omura et al., 2001). This suggests that immature
melatonin-proficient C3H mice (mean gestation molecular clockwork driving rhythmic mPer
109

expression might already be present in individual E7 leads to the disruption of circadian rhythms
fetal mouse SCN neurons. Taken together, these in SCN metabolic activity at E21 and pineal
results show that the endogenous circadian AANAT activity at P10 within the litter (Reppert
rhythm generator in the SCN starts to develop and Schwartz, 1986). Synchronization within the
during late fetal stages and matures gradually dur- fetuses of SCN-lesioned dams can be restored by
ing the postnatal period. In rodents, a mature rhythmic feeding cues (Weaver and Reppert,
molecular clockwork clock is established around 1989). Thus, during fetal development, the circa-
P10. Coincidently, synapse formation, which is dian system is primed to prevailing lighting
sparse at the time of birth, increases dramatically conditions by the maternal circadian system via
between P4 and P10 (Moore and Bernstein, endocrine and/or metabolic substances crossing
1989). Thus, synapse formation and consequently, the placenta (Reppert et al., 1984). Rearing of
synchronization between SCN neurons accounts pubs born to SCN-lesioned dams by intact dams
for robust and coherent circadian rhythmicity in partially restores circadian rhythms in pineal
the developing SCN neuronal network (Moore, AANAT activity at P10 (Reppert and Schwartz,
1991). Circadian rhythms in AANAT activity 1986). This suggests that rhythmic cues from the
in the pineal gland can be observed from P10 mother are necessary and sufficient for entrain-
(Reppert et al., 1984) accurately reflecting rhyth- ment of the newborns. However, the nature of
mic output from the developing SCN (Deguchi, these rhythmic maternal cues is still enigmatic;
1982). Interestingly, the postnatal maturation of they can be provided by the social maternal behav-
the molecular clockwork seems to be independent ior and/or the milk. As far as chemical messengers
of rhythmic maternal cues as robust rhythms are concerned, both dopamine and melatonin have
in Per1 and Per2 expression can be observed in been shown to entrain the fetal circadian rhythms
organotypic SCN slice cultures derived from new- in metabolic activity (Davis and Mannion, 1988;
born (P4–P7) genetically modified mice and rats Weaver et al., 1995). The melatonin receptor 1
bearing luc reporter genes after 1 week up to sev- (MT1) is expressed in the fetal SCN of different
eral months in vitro (Yamaguchi et al., 2001; species (Thomas et al., 2002; Torres-Farfan et al.,
Yamazaki et al., 2000). 2006a; Weaver and Reppert, 1996) and melatonin
It is still not known whether circadian rhythms can readily cross the placenta (McMillen et al.,
in fetal SCN activity are driven by fetal SCN 1990; Yellon and Longo, 1988). In capuchin
molecular clockwork. Although few mouse SCN monkeys, suppression of maternal melatonin by
cells show circadian rhythms in mPER1 and constant light leads to low expression levels of
mPER2 levels (Ansari et al., 2009), there is no Per2 and MT1 and increased expression levels of
evidence that this is sufficient for driving rhythmic Bmal1 in the fetal SCN. The effects on clock gene
Avp expression. Moreover, the rhythms in clock expression levels by constant light condition could
gene expression in the fetal SCN might not be be reversed with melatonin injections (Torres-Far-
self-sustained as in the adult endogenous circa- fan et al., 2006a) indicating that maternal melato-
dian rhythm generator but rather driven by rhyth- nin is capable to influence clock gene expression
mic maternal cues in a peripheral oscillator-like in the fetal SCN. This assumption is further
manner. Importantly, an intact maternal SCN is supported by the observation that in the SCN of
indispensible for the synchronization of the melatonin-deficient C57BL mice, rhythmicity of
embryonic/fetal circadian system. In newborn clock gene expression becomes evident only sev-
Syrian hamsters, the synchrony in rest/activity eral days later (at P5; Huang et al., 2010) than in
within the litter is lost after SCN lesion of the the SCN of melatonin-proficient C3H mice
mother between E7 and E14 (Davis and Gorski, (Ansari et al., 2009). The hypothesis that maternal
1988). In rats, a lesion of the maternal SCN at melatonin drives rhythmic clock gene expression in
110

the fetal mouse SCN needs to be corroborated by Ontogenesis of peripheral circadian oscillators
comparative studies with melatonin (receptor)-
proficient and -deficient mice. Another rhythmic Peripheral oscillators have different capacities for
maternal cue that entrains rodent fetuses is dopa- autonomous rhythmic expression of clock genes
mine acting on D1 receptors in the SCN (Vis- in vitro (Yoo et al., 2004). Some peripheral
wanathan et al., 1994) and this pathway may oscillators such as the liver show robust rhythmic-
provide the daytime signal. In addition to melato- ity over several weeks while others damp out rap-
nin and dopamine, maternal behavior such as loco- idly in vitro demonstrating their dependency on
motor activity, body temperature, and periodic periodic rhythmic input for sustained rhythmic
food intake could entrain the fetal SCN molecular clock gene expression.
clockwork. In Per1-luc transgenic rats, maternal An excellent model for a strongly input-
feeding has been shown to entrain the fetal SCN dependent oscillator is the PT of the anterior pitu-
independent of both the maternal SCN and the itary. The PT is part of the anterior pituitary devel-
light–dark cycle (Ohta et al., 2008). oping from a distinct antero-ventral area of
Environmental lighting has no direct effect on Rathke’s pouch (Stoeckel et al., 1979). The adult
the phase of fetal SCN metabolic activity (Reppert PT is a major target of the hormone melatonin,
and Schwartz, 1983) and this is consistent with the which conveys photoperiodic information to the
observations that the light input pathway into the endocrine system (Hazlerigg, 2001; Lincoln, 2002;
SCN develops postnatally (Mateju et al., 2009; Morgan et al., 1994; Roca et al., 1996). Melatonin
Munoz Llamosas et al., 2000; Sumova et al., 2003; by acting on PT cells affects prolactin secretion
Weaver and Reppert, 1995). In rats, light stimuli from the anterior pituitary (Lincoln, 2002; von
were shown to induce the immediate early gene Gall et al., 2002a) presumably via PT-derived
c-fos in the SCN as early as P0 (Leard et al., 1994; paracrine factors (called tuberalins) that act
Weaver and Reppert, 1995). Per1 and Per2 mRNA anterogradely on lactotrophs in the pars distalis
expression in the SCN can be induced by light at P1 (Morgan and Williams, 1996; Yasuo and Korf,
and P3, respectively (Mateju et al., 2009). At 2011; Yasuo et al., 2010a,b). In addition, melatonin
variance, in hamster and mouse SCN, Fos/Fos-like controls PT-derived thyrotropin which acts retro-
immunoreaction is inducible by light at P4 gradely and activates Dio2 expression in the
(Kaufman and Menaker, 1994; Reppert et al., ependymal cell layer of the infundibular recess
1984). This species difference might be a con- (Yasuo et al., 2007, 2009, 2010b), thereby
sequence of developmental differences in the regulating local thyroid hormone levels in the
formation of the RHT. However, under natural mediobasal hypothalamus. The adult PT shows a
conditions, newborn rodents experience environ- rhythmic expression of clock genes which strongly
mental light only rarely as they are raised in dens. depends on melatonin signaling via the MT1
Feeding and social interaction with the mother receptor. In pinealectomized animals or in mice
might be the most important rhythmic signals for with a targeted deletion of the MT1, no rhythms
the newborns. Rodents feed their pups during in clock genes/proteins can be detected in the adult
the day while the mother takes food during the PT (Jilg et al., 2005; Messager et al., 2001; von Gall
night (Ohta et al., 2003). Most likely, melatonin et al., 2002a, 2005). Importantly, circadian rhythms
is absent in the milk during day (Illnerova et al., in clock proteins are already present in the fetal
1993) and thus provides a reliable rhythmic (E18) PT of melatonin-proficient mice. These
maternal signal for the newborns. After weaning, rhythms are in phase with and show the same
photic input becomes the most important rhyth- amplitude as those in the maternal PT (Ansari
mic cue for entrainment of the circadian system et al., 2009). This suggests that the PT circadian
in the juvenile animals. oscillator is already fully established in the
111

embryonic stage. As the PT circadian oscillator is synchronized by maternal melatonin (Torres-


driven by the rhythmic melatonin signal and fetal Farfan et al., 2006a), demonstrating species-specific
rodents are not capable to produce melatonin differences in the control of fetal peripheral
rhythmically (Deguchi, 1975; Kennaway et al., oscillators.
1992; Nowak et al., 1990), the oscillations in clock The liver has a very strong capacity for autono-
proteins in the fetal PT are presumably driven by mous rhythmic expression of clock genes in vitro
maternal melatonin. (Yoo et al., 2004). In the adult liver, a large num-
The cortex of the adult adrenal gland controls ber of genes (Akhtar et al., 2002; Panda et al.,
metabolic homeostasis and circadian rhythms in 2002; Reddy et al., 2007) and proteins (Reddy
energy balance via the hormone corticosterone. et al., 2006) involved in plasma protein synthesis,
This hormone also represents an important rhyth- glycogen metabolism, detoxification, and the core
mic cue for other peripheral oscillators like the molecular clockwork show circadian fluctuations.
liver (Balsalobre et al., 2000; Reddy et al., 2007). About 60% of the circadian liver transcriptome
Rhythmic corticosteroid production in the adult is driven by GCs (Reddy et al., 2007). In the fetal
adrenal cortex is driven by the SCN by influencing rat liver (E20), detection of clock genes by in situ
both the hypothalamus-pituitary-adrenal (HPA) hybridization did not reveal any circadian
neuroendocrine axis (Abe et al., 1979; Buijs and rhythms except for a low amplitude in Cry1
Kalsbeek, 2001; Buijs et al., 2003) and the sympa- (Sladek et al., 2007). This study has also shown
thetic nervous system (Buijs et al., 2003; Ishida that the molecular clockwork in the liver matures
et al., 2005; Lemos et al., 2006; Ulrich-Lai et al., even more slowly during postnatal development
2006). Melatonin inhibits the activation of cortisol than the fetal SCN oscillator and rhythmic expres-
secretion induced by ACTH (Torres-Farfan et al., sion of all clock genes has been achieved only at
2003, 2004). The adult adrenal cortex possesses a P30 (Sladek et al., 2007; Sumova et al., 2004). In
molecular clockwork (Bittman et al., 2003; contrast, the fetal liver of Per1-luc transgenic rats
Fahrenkrug et al., 2008; Ishida et al., 2005; Lemos shows a circadian rhythm in Per1 expression for
et al., 2006; Oster et al., 2006b; Torres-Farfan at least 2 days in vitro that, similarly to the fetal
et al., 2006b; Valenzuela et al., 2008) which gates SCN and the adult liver, can be entrained by
rhythmic cortisol secretion in response to ACTH maternal feeding (Ohta et al., 2008). Thus, the
(Oster et al., 2006b). Melatonin-proficient mouse fetal liver represents a damped circadian oscilla-
strains show circadian rhythms in clock gene tor that gradually matures during postnatal life.
expression in the adrenal cortex in contrast to
melatonin-deficient mouse strains (Torres-Farfan
et al., 2006b) suggesting an important role of mel- Summary
atonin for driving the adrenal cortex peripheral
oscillator. The fetal rat adrenal gland (E18) shows In fetal rodents, the SCN, adrenal gland, and liver
circadian rhythms in clock gene expression and in show a circadian rhythm of clock gene expression,
cortisol secretion which persist at least for 2 days which persists for at least 2 days in vitro,
in vitro (Torres-Farfan et al., 2011). Melatonin demonstrating the existence of circadian fetal
affects the phase of clock gene expression in oscillators. These can be entrained by maternal
cultured adrenal glands (Torres-Farfan et al., cues such as melatonin (impact on SCN and adre-
2011), suggesting that maternal melatonin might nal gland), dopamine (SCN), and maternal feed-
trigger the molecular clockwork in this develop- ing (SCN and liver). It is still a matter of debate
ing peripheral oscillator. However, in the adrenal whether the fetal SCN and the fetal adrenal gland
gland of fetal capuchin monkeys, the circadian produce rhythmic output signals, such as vaso-
rhythm in clock gene expression is not pressin, and cortisol, respectively, which might
112

Mother

Rhythmic Melatonin
signals
(e.g.,
feeding)

Placenta

Fetus

Rhythmic Fetal peripheral


signals? oscillators
(e.g., liver,
adrenal, PT)

Fig. 1. Model depicting synchronization of fetal circadian oscillators by rhythmic maternal cues. The endogenous rhythm generator
in the maternal SCN is synchronized by light and generates circadian rhythms in melatonin and other hormones, body temperature,
and metabolites. Via the placenta these rhythmic maternal signals reach the fetus in which they synchronize the SCN, and self-
sustained peripheral oscillators, such as the liver and the adrenal gland, and drive input-dependent peripheral oscillators, such as
the pars tuberalis of the pituitary (PT). It is still a matter of debate whether vasopressin or other rhythmic signals generated by
the fetal SCN are capable of driving/synchronizing fetal peripheral oscillators. Modified after Reppert and Weaver (2002) and
Ansari et al. (2009).

1200
Relative optical density
Number of cells/SCN

1.0 Adult

900 P10
P2
600 0.5 E18

300

0 0.0
00 06 12 18 24 00 06 12 18 24
Circadian time Circadian time

Fig. 2. Analysis of mPER1-immunoreaction (Ir) in SCN (left) and PT (right) during mouse ontogenetic development (mean  SEM
of five animals per time point and ontogenetic stage. Data points at CT00/24 are double-plotted). Fetal SCN and PT show low- and
high-amplitude circadian oscillation of mPER1-Ir, respectively. Modified after Ansari et al. (2009).
113

provide intrinsic entrainment signals to other the mouse liver transcriptome, as revealed by cDNA micro-
peripheral oscillators in the fetus (Fig. 1). Alter- array, is driven by the suprachiasmatic nucleus. Current
Biology, 12, 540–550.
nately, peripheral oscillators in the fetus may be Albrecht, U., Sun, Z. S., Eichele, G., & Lee, C. C. (1997).
exclusively entrained by maternal signals until A differential response of two putative mammalian circadian
birth/weaning. The core molecular clockwork regulators, mper1 and mper2, to light. Cell, 91, 1055–1064.
in SCN and liver, which represent robust self- Ambriz-Tututi, M., Rocha-Gonzalez, H. I., Cruz, S. L., &
sustained oscillators in the adult, matures gradu- Granados-Soto, V. (2009). Melatonin: A hormone that
modulates pain. Life Sciences, 84, 489–498.
ally during postnatal development. In vitro stud- Ansari, N., Agathagelidis, M., Lee, C., Korf, H. W., &
ies suggest that this maturation is independent of von Gall, C. (2009). Differential maturation of circadian
rhythmic maternal cues. In contrast, the core rhythms in clock gene proteins in the suprachiasmatic
molecular clockwork in strictly input-dependent nucleus and the pars tuberalis during mouse ontogeny. The
oscillators, such as the PT, appears to be driven European Journal of Neuroscience, 29, 477–489.
Aschoff, J. (1965). Circadian rhythms in man—A self-
by maternal cues and matures already during fetal sustained oscillator with an inherent frequency underlies
life (Fig. 2). In conclusion, maternal cues contrib- human 24-hour periodicity. Science, 148, 1427–1432.
ute to the ontogenetic development of circadian Balsalobre, A. (2002). Clock genes in mammalian peripheral
oscillators; depending on the organ, the signals tissues. Cell and Tissue Research, 309, 193–199.
from the mother can either drive or entrain the Balsalobre, A., Brown, S. A., Marcacci, L., Tronche, F.,
Kellendonk, C., Reichardt, H. M., et al. (2000). Resetting
fetal and newborn clockwork. of circadian time in peripheral tissues by glucocorticoid sig-
naling. Science, 289, 2344–2347.
Bittman, E. L., Doherty, L., Huang, L., & Paroskie, A. (2003).
Period gene expression in mouse endocrine tissues. Ameri-
Abbreviations can Journal of Physiology. Regulatory, Integrative and Com-
parative Physiology, 285, R561–R569.
ACTH adrenocorticotropin Buijs, R. M., & Kalsbeek, A. (2001). Hypothalamic integration
AVP arginine vasopressin of central and peripheral clocks. Nature Reviews. Neurosci-
ence, 2, 521–526.
CRH corticotrophin-releasing hormone Buijs, R. M., van Eden, C. G., Goncharuk, V. D., &
E embryonic day Kalsbeek, A. (2003). The biological clock tunes the organs
GC glucocorticoid of the body: Timing by hormones and the autonomic ner-
HPA hypothalamo-pituitary-adrenal vous system. The Journal of Endocrinology, 177, 17–26.
luc luciferase Bur, I. M., Zouaoui, S., Fontanaud, P., Coutry, N., Molino, F.,
Martin, A. O., et al. (2010). The comparison between circa-
MT1 melatonin receptor 1 dian oscillators in mouse liver and pituitary gland reveals
NRC negative regulator complex different integration of feeding and light schedules. PLoS
P postnatal day One, 5, e15316.
PT pars tuberalis Cajochen, C., Krauchi, K., & Wirz-Justice, A. (2003). Role of
PVN paraventricular nucleus melatonin in the regulation of human circadian rhythms
and sleep. Journal of Neuroendocrinology, 15, 432–437.
RHT retinohypothalamic tract Cheng, M. Y., Bullock, C. M., Li, C., Lee, A. G.,
SCN suprachiasmatic nucleus Bermak, J. C., Belluzzi, J., et al. (2002). Prokineticin 2 tra-
nsmits the behavioural circadian rhythm of the
suprachiasmatic nucleus. Nature, 417, 405–410.
Damiola, F., Le Minh, N., Preitner, N., Kornmann, B., Fleury-
References Olela, F., & Schibler, U. (2000). Restricted feeding
uncouples circadian oscillators in peripheral tissues from
Abe, K., Kroning, J., Greer, M. A., & Critchlow, V. (1979). Effects the central pacemaker in the suprachiasmatic nucleus. Genes
of destruction of the suprachiasmatic nuclei on the circadian & Development, 14, 2950–2961.
rhythms in plasma corticosterone, body temperature, feeding Davis, F. C., & Gorski, R. A. (1988). Development of hamster
and plasma thyrotropin. Neuroendocrinology, 29, 119–131. circadian rhythms: Role of the maternal suprachiasmatic
Akhtar, R. A., Reddy, A. B., Maywood, E. S., Clayton, J. D., nucleus. Journal of Comparative Physiology. A, 162,
King, V. M., Smith, A. G., et al. (2002). Circadian cycling of 601–610.
114

Davis, F. C., & Mannion, J. (1988). Entrainment of hamster Johnson, C. H., & Golden, S. S. (1999). Circadian programs in
pup circadian rhythms by prenatal melatonin injections to cyanobacteria: Adaptiveness and mechanism. Annual
the mother. The American Journal of Physiology, 255, Review of Microbiology, 53, 389–409.
R439–R448. Kalsbeek, A., Buijs, R. M., Vanheerikhuize, J. J., Arts, M., &
Deguchi, T. (1975). Ontogenesis of a biological clock for Vanderwoude, T. P. (1992). Vasopressin-containing neurons
serotonin: Acetyl coenzyme A N-acetyltransferase in pineal of the suprachiasmatic nuclei inhibit corticosterone release.
gland of rat. Proceedings of the National Academy of Brain Research, 580, 62–67.
Sciences of the United States of America, 72, 2814–2818. Kalsbeek, A., Fliers, E., Franke, A. N., Wortel, J., &
Deguchi, T. (1982). Sympathetic regulation of circadian Buijs, R. M. (2000). Functional connections between the
rhythm of serotonin N-acetyltransferase activity in pineal suprachiasmatic nucleus and the thyroid gland as revealed
gland of infant rat. Journal of Neurochemistry, 38, 797–802. by lesioning and viral tracing techniques in the rat. Endocri-
Dunlap, J. C. (1999). Molecular bases for circadian clocks. nology, 141, 3832–3841.
Cell, 96, 271–290. Kalsbeek, A., La Fleur, S., Van Heijningen, C., & Buijs, R. M.
Dvornyk, V., Vinogradova, O., & Nevo, E. (2003). Origin and (2004). Suprachiasmatic GABAergic inputs to the para-
evolution of circadian clock genes in prokaryotes. Pro- ventricular nucleus control plasma glucose concentrations
ceedings of the National Academy of Sciences of the United in the rat via sympathetic innervation of the liver. The Jour-
States of America, 100, 2495–2500. nal of Neuroscience, 24, 7604–7613.
Fahrenkrug, J., Hannibal, J., & Georg, B. (2008). Diurnal Kalsbeek, A., Perreau-Lenz, S., & Buijs, R. M. (2006). A net-
rhythmicity of the canonical clock genes Per1, Per2 and work of (Autonomic) clock outputs. Chronobiology Interna-
Bmal1 in the rat adrenal gland is unaltered after hypophy- tional, 23, 521–535.
sectomy. Journal of Neuroendocrinology, 20, 323–329. Kaufman, C. M., & Menaker, M. (1994). Ontogeny of light-
Gau, D., Lemberger, T., von Gall, C., Kretz, O., Le Minh, N., induced Fos-like immunoreactivity in the hamster
Gass, P., et al. (2002). Phosphorylation of CREB Ser142 suprachiasmatic nucleus. Brain Research, 633, 162–166.
regulates light-induced phase shifts of the circadian clock. Kennaway, D. J., Stamp, G. E., & Goble, F. C. (1992). Devel-
Neuron, 34, 245–253. opment of melatonin production in infants and the impact of
Hastings, M. H., Reddy, A. B., Garabette, M., King, V. M., prematurity. The Journal of Clinical Endocrinology and
Chahad-Ehlers, S., O’Brien, J., et al. (2003). Expression of Metabolism, 75, 367–369.
clock gene products in the suprachiasmatic nucleus in relation Klein, D. C., Moore, R. Y., & Reppert, S. M. (1991). The
to circadian behaviour. Novartis Foundation Symposium, 253, suprachiasmatic nucleus: The mind’s clock. New York, NY:
203–217 (Discussion 102–209, 218–222, 281–204). Oxford University press.
Hazlerigg, D. G. (2001). Review—What is the role of melato- Kovacikova, Z., Sladek, M., Bendova, Z., Illnerova, H., &
nin within the anterior pituitary? The Journal of Endocrinol- Sumova, A. (2006). Expression of clock and clock-driven
ogy, 170, 493–501. genes in the rat suprachiasmatic nucleus during late fetal
Huang, J., Lu, C., Chen, S., Hua, L., & Qian, R. (2010). Post- and early postnatal development. Journal of Biological
natal ontogenesis of clock genes in mouse suprachiasmatic Rhythms, 21, 140–148.
nucleus and heart. Lipids in Health and Disease, 9, 22. Kramer, A., Yang, F. C., Snodgrass, P., Li, X. D.,
Illnerova, H., Buresova, M., & Presl, J. (1993). Melatonin Scammell, T. E., Davis, F. C., et al. (2001). Regulation of
rhythm in human milk. The Journal of Clinical Endocrinol- daily locomotor activity and sleep by hypothalamic EGF
ogy and Metabolism, 77, 838–841. receptor signaling. Science, 294, 2511–2515.
Ishida, A., Mutoh, T., Ueyama, T., Bando, H., Masubuchi, S., Kraves, S., & Weitz, C. J. (2006). A role for cardiotrophin-like
Nakahara, D., et al. (2005). Light activates the adrenal cytokine in the circadian control of mammalian locomotor
gland: Timing of gene expression and glucocorticoid release. activity. Nature Neuroscience, 9, 212–219.
Cell Metabolism, 2, 297–307. Leard, L. E., Macdonald, E. S., Heller, H. C., & Kilduff, T. S.
Jilg, A., Moek, J., Weaver, D. R., Korf, H. W., Stehle, J. H., (1994). Ontogeny of photic-induced c-fos mRNA expression
& von Gall, C. (2005). Rhythms in clock proteins in the in rat suprachiasmatic nuclei. Neuroreport, 5, 2683–2687.
mouse pars tuberalis depend on MT1 melatonin receptor Lemos, D. R., Downs, J. L., & Urbanski, H. F. (2006).
signalling. The European Journal of Neuroscience, 22, Twenty-four-hour rhythmic gene expression in the rhesus
2845–2854. macaque adrenal gland. Molecular Endocrinology, 20,
Jin, X., Shearman, L. P., Weaver, D. R., Zylka, M. J., 1164–1176.
de Vries, G. J., & Reppert, S. M. (1999). A molecular mech- LeSauter, J., Lehman, M. N., & Silver, R. (1996). Restoration
anism regulating rhythmic output from the suprachiasmatic of circadian rhythmicity by transplants of SCN “micro-
circadian clock. Cell, 96, 57–68. punches” Journal of Biological Rhythms, 11, 163–171.
115

Li, X., & Davis, F. C. (2005). Developmental expression of pharmacology and physiological significance. Neurochemis-
clock genes in the Syrian hamster. Brain Research. Develop- try International, 24, 101–146.
mental Brain Research, 158, 31–40. Morgan, P. J., & Williams, L. M. (1996). The pars tuberalis of
Lincoln, G. A. (2002). Neuroendocrine regulation of seasonal the pituitary: A gateway for neuroendocrine output.
gonadotrophin and prolactin rhythms: Lessons from the Reviews of Reproduction, 1, 153–161.
Soay ram model. Reproduction Supplement, 59, 131–147. Munoz Llamosas, M., Huerta, J. J., Cernuda-Cernuda, R., &
Liu, A. C., Welsh, D. K., Ko, C. H., Tran, H. G., Zhang, E. E., Garcia-Fernandez, J. M. (2000). Ontogeny of a photic
Priest, A. A., et al. (2007). Intercellular coupling confers response in the retina and suprachiasmatic nucleus in the
robustness against mutations in the SCN circadian clock net- mouse. Brain Research. Developmental Brain Research,
work. Cell, 129, 605–616. 120, 1–6.
Mateju, K., Bendova, Z., El-Hennamy, R., Sladek, M., Nowak, R., Young, I. R., & McMillen, I. C. (1990). Emergence
Sosniyenko, S., & Sumova, A. (2009). Development of the of the diurnal rhythm in plasma melatonin concentrations in
light sensitivity of the clock genes Period1 and Period2, and newborn lambs delivered to intact or pinealectomized ewes.
immediate-early gene c-fos within the rat suprachiasmatic The Journal of Endocrinology, 125, 97–102.
nucleus. The European Journal of Neuroscience, 29, 490–501. Ohta, H., Honma, S., Abe, H., & Honma, K. (2003). Periodic
Maywood, E. S., Chesham, J. E., O’Brien, J. A., & absence of nursing mothers phase-shifts circadian rhythms
Hastings, M. H. (2011). A diversity of paracrine signals of clock genes in the suprachiasmatic nucleus of rat pups.
sustains molecular circadian cycling in suprachiasmatic The European Journal of Neuroscience, 17, 1628–1634.
nucleus circuits. Proceedings of the National Academy of Ohta, H., Xu, S., Moriya, T., Iigo, M., Watanabe, T.,
Sciences of the United States of America, 108, 14306–14311. Nakahata, N., et al. (2008). Maternal feeding controls fetal
McMillen, I. C., Nowak, R., Walker, D. W., & Young, I. R. biological clock. PLoS One, 3, e2601.
(1990). Maternal pinealectomy alters the daily pattern of Oishi, K., Sakamoto, K., Okada, T., Nagase, T., & Ishida, N.
fetal breathing in sheep. The American Journal of Physiol- (1998). Antiphase circadian expression between BMAL1
ogy, 258, R284–R287. and period homologue mRNA in the suprachiasmatic
McNamara, P., Seo, S. B., Rudic, R. D., Sehgal, A., nucleus and peripheral tissues of rats. Biochemical and Bio-
Chakravarti, D., & FitzGerald, G. A. (2001). Regulation of physical Research Communications, 253, 199–203.
CLOCK and MOP4 by nuclear hormone receptors in the Okamura, H., Yamaguchi, S., & Yagita, K. (2002). Molecular
vasculature: A humoral mechanism to reset a peripheral machinery of the circadian clock in mammals. Cell and Tis-
clock. Cell, 105, 877–889. sue Research, 309, 47–56.
Messager, S., Garabette, M. L., Hastings, M. H., & Oster, H., Damerow, S., Hut, R. A., & Eichele, G. (2006a).
Hazlerigg, D. G. (2001). Tissue-specific abolition of Per1 Transcriptional profiling in the adrenal gland reveals
expression in the pars tuberalis by pinealectomy in the circadian regulation of hormone biosynthesis genes and
Syrian hamster. Neuroreport, 12, 579–582. nucleosome assembly genes. Journal of Biological Rhythms,
Messager, S., Hazlerigg, D. G., Mercer, J. G., & Morgan, P. J. 21, 350–361.
(2000). Photoperiod differentially regulates the expression Oster, H., Damerow, S., Kiessling, S., Jakubcakova, V.,
of Per1 and ICER in the pars tuberalis and the Abraham, D., Tian, J., et al. (2006b). The circadian rhythm
suprachiasmatic nucleus of the Siberian hamster. The Euro- of glucocorticoids is regulated by a gating mechanism residing
pean Journal of Neuroscience, 12, 2865–2870. in the adrenal cortical clock. Cell Metabolism, 4, 163–173.
Moore, R. Y. (1991). Development of the supraciasmatic Panda, S., Antoch, M. P., Miller, B. H., Su, A. I.,
nucleus. In D. C. Klein, R. J. Moore & S. M. Reppert Schook, A. B., Straume, M., et al. (2002). Coordinated tran-
(Eds.), Suprachiasmatic nucleus: The mind’s clock scription of key pathways in the mouse by the circadian
(pp. 197–216). New York, NY: Oxford University Press. clock. Cell, 109, 307–320.
Moore, R. Y., & Bernstein, M. E. (1989). Synaptogenesis in Perreau-Lenz, S., Kalsbeek, A., Garidou, M. L., Wortel, J.,
the rat suprachiasmatic nucleus demonstrated by electron van der Vliet, J., van Heijningen, C., et al. (2003).
microscopy and synapsin I immunoreactivity. The Journal Suprachiasmatic control of melatonin synthesis in rats:
of Neuroscience, 9, 2151–2162. Inhibitory and stimulatory mechanisms. The European Jour-
Moore, R. Y., & Leak, R. K. (2001). Suprachiasmatic nucleus. nal of Neuroscience, 17, 221–228.
In J. S. Takahashi, F. W. Turek & R. Y. Moore (Eds.), Peschke, E. (2008). Melatonin, endocrine pancreas and diabe-
Handbook of behavioral neurobiology 12, circadian clocks tes. Journal of Pineal Research, 44, 26–40.
(pp. 141–171). New York, NY: Kluwer Academic. Ralph, M. R., Foster, R. G., Davis, F. C., & Menaker, M.
Morgan, P. J., Barrett, P., Howell, H. E., & Helliwell, R. (1990). Transplanted suprachiasmatic nucleus determines
(1994). Melatonin receptors: Localization, molecular circadian period. Science, 247, 975–978.
116

Ralph, M. R., & Menaker, M. (1988). A mutation of the circa- and abused drugs. Advanced Drug Delivery Reviews, 62,
dian system in golden hamsters. Science, 241, 1225–1227. 918–927.
Reddy, A. B., Karp, N. A., Maywood, E. S., Sage, E. A., Shimomura, H., Moriya, T., Sudo, M., Wakamatsu, H.,
Deery, M., O’Neill, J. S., et al. (2006). Circadian orchestra- Akiyama, M., Miyake, Y., et al. (2001). Differential daily
tion of the hepatic proteome. Current Biology, 16, expression of Per1 and Per2 mRNA in the suprachiasmatic
1107–1115. nucleus of fetal and early postnatal mice. The European
Reddy, A. B., Maywood, E. S., Karp, N. A., King, V. M., Journal of Neuroscience, 13, 687–693.
Inoue, Y., Gonzalez, F. J., et al. (2007). Glucocorticoid sig- Silver, R., LeSauter, J., Tresco, P. A., & Lehman, M. N.
naling synchronizes the liver circadian transcriptome. (1996). A diffusible coupling signal from the transplanted
Hepatology, 45, 1478–1488. suprachiasmatic nucleus controlling circadian locomotor
Reppert, S. M., Coleman, R. J., Heath, H. W., & rhythms. Nature, 382, 810–813.
Swedlow, J. R. (1984). Pineal N-acetyltransferase activity Sladek, M., Jindrakova, Z., Bendova, Z., & Sumova, A.
in 10-day-old rats: A paradigm for studying the developing (2007). Postnatal ontogenesis of the circadian clock within
circadian system. Endocrinology, 115, 918–925. the rat liver. American Journal of Physiology. Regulatory,
Reppert, S. M., & Schwartz, W. J. (1983). Maternal coordina- Integrative and Comparative Physiology, 292,
tion of the fetal biological clock in utero. Science, 220, R1224–R1229.
969–971. Stephan, F. K. (2002). The “other” circadian system: Food as a
Reppert, S. M., & Schwartz, W. J. (1984a). Functional activity Zeitgeber. Journal of Biological Rhythms, 17, 284–292.
of the suprachiasmatic nuclei in the fetal primate. Neurosci- Stoeckel, M. E., Hindelang-Gertner, C., & Porte, A. (1979).
ence Letters, 46, 145–149. Embryonic development and secretory differentiation in
Reppert, S. M., & Schwartz, W. J. (1984b). The suprachiasmatic the pars tuberalis of the mouse hypophysis. Cell and Tissue
nuclei of the fetal rat: Characterization of a functional circa- Research, 198, 465–476.
dian clock using 14C-labeled deoxyglucose. The Journal of Sumova, A., Bendova, Z., Sladek, M., Kovacikova, Z., El-
Neuroscience, 4, 1677–1682. Hennamy, R., Laurinova, K., et al. (2006). The rat circadian
Reppert, S. M., & Schwartz, W. J. (1986). Maternal clockwork and its photoperiodic entrainment during devel-
suprachiasmatic nuclei are necessary for maternal coordina- opment. Chronobiology International, 23, 237–243.
tion of the developing circadian system. The Journal of Neu- Sumova, A., Jac, M., Sladek, M., Sauman, I., & Illnerova, H.
roscience, 6, 2724–2729. (2003). Clock gene daily profiles and their phase relation-
Reppert, S. M., & Uhl, G. R. (1987). Vasopressin messenger ship in the rat suprachiasmatic nucleus are affected by pho-
ribonucleic acid in supraoptic and suprachiasmatic nuclei: toperiod. Journal of Biological Rhythms, 18, 134–144.
Appearance and circadian regulation during development. Sumova, A., Sladek, M., Kovacikova, Z., Bendova, Z.,
Endocrinology, 120, 2483–2487. Laurinova, K., & Illnerova, H. (2004). Insight into molecular
Reppert, S. M., & Weaver, D. R. (2002). Coordination of cir- core clock mechanism of embryonic and early postnatal rat
cadian timing in mammals. Nature, 418, 935–941. suprachiasmatic nucleus. Proceedings of the National Acad-
Roca, A. L., Godson, C., Weaver, D. R., & Reppert, S. M. emy of Sciences of the United States of America, 101,
(1996). Structure, characterization, and expression of the 6231–6236.
gene encoding the mouse Mel1a melatonin receptor. Endo- Sun, Z. S., Albrecht, U., Zhuchenko, O., Bailey, J.,
crinology, 137, 3469–3477. Eichele, G., & Lee, C. C. (1997). RIGUI, a putative mam-
Schibler, U., Ripperger, J., & Brown, S. A. (2003). Peripheral malian ortholog of the Drosophila period gene. Cell, 90,
circadian oscillators in mammals: Time and food. Journal of 1003–1011.
Biological Rhythms, 18, 250–260. Thomas, L., Purvis, C. C., Drew, J. E., Abramovich, D. R., &
Schwartz, W. J., & Zimmerman, P. (1990). Circadian time- Williams, L. M. (2002). Melatonin receptors in human fetal
keeping in BALB/c and C57BL/6 inbred mouse strains. brain: 2-[(125)I]iodomelatonin binding and MT1 gene
The Journal of Neuroscience, 10, 3685–3694. expression. Journal of Pineal Research, 33, 218–224.
Shearman, L. P., Zylka, M. J., Weaver, D. R., Torres-Farfan, C., Mendez, N., Abarzua-Catalan, L.,
Kolakowski, L. F., Jr., & Reppert, S. M. (1997). Two period Vilches, N., Valenzuela, G. J., & Seron-Ferre, M. (2011).
homologs: Circadian expression and photic regulation in the A circadian clock entrained by melatonin is ticking in the
suprachiasmatic nuclei. Neuron, 19, 1261–1269. rat fetal adrenal. Endocrinology, 152, 1891–1900.
Shibata, S., & Moore, R. Y. (1987). Development of neuronal Torres-Farfan, C., Richter, H. G., Germain, A. M.,
activity in the rat suprachiasmatic nucleus. Brain Research, Valenzuela, G. J., Campino, C., Rojas-Garcia, P., et al.
431, 311–315. (2004). Maternal melatonin selectively inhibits cortisol pro-
Shibata, S., Tahara, Y., & Hirao, A. (2010). The adjustment duction in the primate fetal adrenal gland. The Journal of
and manipulation of biological rhythms by light, nutrition, Physiology, 554, 841–856.
117

Torres-Farfan, C., Richter, H. G., Rojas-Garcia, P., Rhythmic gene expression in pituitary depends on heterolo-
Vergara, M., Forcelledo, M. L., Valladares, L. E., et al. gous sensitization by the neurohormone melatonin. Nature
(2003). mt1 Melatonin receptor in the primate adrenal Neuroscience, 5, 234–238.
gland: Inhibition of adrenocorticotropin-stimulated cortisol von Gall, C., Noton, E., Lee, C., & Weaver, D. R. (2003).
production by melatonin. The Journal of Clinical Endocri- Light does not degrade the constitutively expressed BMAL1
nology and Metabolism, 88, 450–458. protein in the mouse suprachiasmatic nucleus. The Euro-
Torres-Farfan, C., Rocco, V., Monso, C., Valenzuela, F. J., pean Journal of Neuroscience, 18, 125–133.
Campino, C., Germain, A., et al. (2006a). Maternal melato- von Gall, C., Schneider-Huther, I., Pfeffer, M., Dehghani, F.,
nin effects on clock gene expression in a nonhuman primate Korf, H. W., & Stehle, J. H. (2001). Clock gene protein
fetus. Endocrinology, 147, 4618–4626. mPER1 is rhythmically synthesized and under cAMP control
Torres-Farfan, C., Seron-Ferre, M., Dinet, V., & Korf, H. W. in the mouse pineal organ. Journal of Neuroendocrinology,
(2006b). Immunocytochemical demonstration of day/night 13, 313–316.
changes of clock gene protein levels in the murine adrenal von Gall, C., Stehle, J. H., & Weaver, D. R. (2002b). Mamma-
gland: Differences between melatonin-proficient (C3H) lian melatonin receptors: Molecular biology and signal
and melatonin-deficient (C57BL) mice. Journal of Pineal transduction. Cell and Tissue Research, 309, 151–162.
Research, 40, 64–70. von Gall, C., Weaver, D. R., Moek, J., Jilg, A., Stehle, J. H., &
Tousson, E., & Meissl, H. (2004). Suprachiasmatic nuclei Korf, H. W. (2005). Melatonin plays a crucial role in the reg-
grafts restore the circadian rhythm in the paraventricular ulation of rhythmic clock gene expression in the mouse pars
nucleus of the hypothalamus. The Journal of Neuroscience, tuberalis. Annals of the New York Academy of Sciences,
24, 2983–2988. 1040, 508–511.
Ulrich-Lai, Y. M., Arnhold, M. M., & Engeland, W. C. (2006). Vrang, N., Larsen, P. J., Moller, M., & Mikkelsen, J. D. (1995).
Adrenal splanchnic innervation contributes to the diurnal Topographical organization of the rat suprachiasmatic-
rhythm of plasma corticosterone in rats by modulating adre- paraventricular projection. The Journal of Comparative
nal sensitivity to ACTH. American Journal of Physiology. Neurology, 353, 585–603.
Regulatory, Integrative and Comparative Physiology, 290, Weaver, D. R., & Reppert, S. M. (1989). Periodic feeding of
R1128–R1135. SCN-lesioned pregnant rats entrains the fetal biological
Unfried, C., Ansari, N., Yasuo, S., Korf, H. W., & von Gall, C. clock. Brain Research. Developmental Brain Research, 46,
(2009). Impact of melatonin and molecular clockwork com- 291–296.
ponents on the expression of thyrotropin beta-chain (Tshb) Weaver, D. R., & Reppert, S. M. (1995). Definition of the
and the Tsh receptor in the mouse pars tuberalis. Endocri- developmental transition from dopaminergic to photic regu-
nology, 150, 4653–4662. lation of c-fos gene expression in the rat suprachiasmatic
Unfried, C., Burbach, G., Korf, H. W., & von Gall, C. (2010). nucleus. Brain Research. Molecular Brain Research, 33,
Melatonin receptor 1-dependent gene expression in the 136–148.
mouse pars tuberalis as revealed by cDNA microarray anal- Weaver, D. R., & Reppert, S. M. (1996). The Mel1a melatonin
ysis and in situ hybridization. Journal of Pineal Research, 48, receptor gene is expressed in human suprachiasmatic nuclei.
148–156. Neuroreport, 8, 109–112.
Valenzuela, F. J., Torres-Farfan, C., Richter, H. G., Weaver, D. R., Roca, A. L., & Reppert, S. M. (1995). c-fos and
Mendez, N., Campino, C., Torrealba, F., et al. (2008). Clock jun-B mRNAs are transiently expressed in fetal rodent
gene expression in adult primate suprachiasmatic nuclei and suprachiasmatic nucleus following dopaminergic stimula-
adrenal: Is the adrenal a peripheral clock responsive to mel- tion. Brain Research. Developmental Brain Research, 85,
atonin? Endocrinology, 149, 1454–1461. 293–297.
Viswanathan, N., Weaver, D. R., Reppert, S. M., & Weinert, D. (2005). Ontogenetic development of the mamma-
Davis, F. C. (1994). Entrainment of the fetal hamster circa- lian circadian system. Chronobiology International, 22,
dian pacemaker by prenatal injections of the dopamine ago- 179–205.
nist SKF 38393. The Journal of Neuroscience, 14, 5393–5398. Yamaguchi, S., Kobayashi, M., Mitsui, S., Ishida, Y., van der
von Gall, C., Duffield, G. E., Hastings, M. H., Kopp, M. D., Horst, G. T., Suzuki, M., et al. (2001). View of a mouse
Dehghani, F., Korf, H. W., et al. (1998). CREB in the clock gene ticking. Nature, 409, 684.
mouse SCN: A molecular interface coding the phase- Yamazaki, S., Numano, R., Abe, M., Hida, A., Takahashi, R.,
adjusting stimuli light, glutamate, PACAP, and melatonin Ueda, M., et al. (2000). Resetting central and peripheral
for clockwork access. The Journal of Neuroscience, 18, circadian oscillators in transgenic rats. Science, 288,
10389–10397. 682–685.
von Gall, C., Garabette, M. L., Kell, C. A., Frenzel, S., Yamazaki, S., Yoshikawa, T., Biscoe, E. W., Numano, R.,
Dehghani, F., Schumm-Draeger, P. M., et al. (2002a). Gallaspy, L. M., Soulsby, S., et al. (2009). Ontogeny of
118

circadian organization in the rat. Journal of Biological Yasuo, S., Yoshimura, T., Ebihara, S., & Korf, H. W. (2010b).
Rhythms, 24, 55–63. Photoperiodic control of TSH-beta expression in the mam-
Yasuo, S., Koch, M., Schmidt, H., Ziebell, S., Bojunga, J., malian pars tuberalis has different impacts on the induction
Geisslinger, G., et al. (2010a). An endocannabinoid system and suppression of the hypothalamo-hypopysial gonadal
is localized to the hypophysial pars tuberalis of Syrian axis. Journal of Neuroendocrinology, 22, 43–50.
hamsters and responds to photoperiodic changes. Cell and Yellon, S. M., & Longo, L. D. (1988). Effect of maternal pine-
Tissue Research, 340, 127–136. alectomy and reverse photoperiod on the circadian melato-
Yasuo, S., & Korf, H. W. (2011). The hypophysial pars nin rhythm in the sheep and fetus during the last trimester
tuberalis transduces photoperiodic signals via multiple pat- of pregnancy. Biology of Reproduction, 39, 1093–1099.
hways and messenger molecules. General and Comparative Yoo, S. H., Yamazaki, S., Lowrey, P. L., Shimomura, K.,
Endocrinology, 172, 15–22. Ko, C. H., Buhr, E. D., et al. (2004). PERIOD2::LUCIFER-
Yasuo, S., Yoshimura, T., Ebihara, S., & Korf, H. W. (2007). ASE real-time reporting of circadian dynamics reveals
Temporal dynamics of type 2 deiodinase expression after persistent circadian oscillations in mouse peripheral tissues.
melatonin injections in Syrian hamsters. Endocrinology, Proceedings of the National Academy of Sciences of the
148, 4385–4392. United States of America, 101, 5339–5346.
Yasuo, S., Yoshimura, T., Ebihara, S., & Korf, H. W. (2009). Zylka, M. J., Shearman, L. P., Weaver, D. R., & Reppert, S. M.
Melatonin transmits photoperiodic signals through the (1998). Three period homologs in mammals: Differential light
MT1 melatonin receptor. The Journal of Neuroscience, 29, responses in the suprachiasmatic circadian clock and oscillating
2885–2889. transcripts outside of brain. Neuron, 20, 1103–1110.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 7

The circadian output signals from


the suprachiasmatic nuclei

Jia-Da Li{,*, Wang-Ping Hu{ and Qun-Yong Zhou}

{
The State Key Laboratory of Medical Genetics, Central South University, Changsha, Hunan, PR China
{
Department of Pharmacology, Hubei University of Science and Technology, Xianning, Hubei, PR China
}
Department of Pharmacology, University of California Irvine, Irvine, CA, USA

Abstract: The suprachiasmatic nuclei (SCN) of the anterior hypothalamus comprise a self-sustained
biological clock generating an endogenous  24-h circadian rhythm, driving many overt daily rhythms in
the body. An important remaining question is how the SCN neurons communicate with their efferent
targets to control the daily oscillations in behavior and physiology. In this chapter, we summarize several
signaling factors that may serve as such SCN output factors. Whereas vasopressin may be involved in
the regulation of circadian hormone rhythms, SCN-derived prokineticin 2 (PK2), TGF-a, and
cardiotrophin-like cytokine (CLC) may serve as output factors for other circadian rhythms, including
locomotor activity, body temperature, and energy metabolism. The circadian rhythm in firing activity of
SCN neurons is also likely to be a critical output signaling mechanism. The likely involvement of these
output factors in the generation of the circadian rhythm in SCN neuronal firing activity is also discussed.

Keywords: circadian; suprachiasmatic nuclei; prokineticin 2; vasopressin; TGF-a; cardiotrophin-like


cytokine.

Introduction earth. In mammals, the endogenous pacemaker


that drives circadian rhythms resides in the
Almost all organisms, from bacteria to mammals, suprachiasmatic nuclei (SCN) of the anterior hypo-
exhibit circadian ( 24 h) rhythms in physiology thalamus (Lowrey and Takahashi, 2004; Reppert
and behavior to adapt to the environmental and Weaver, 2002; Schibler, 2005). The SCN coor-
changes imposed by the daily revolutions of the dinate daily cycles of behavior and physiology,
such as the daily rhythms of locomotor activity
and body temperature (Lowrey and Takahashi,
*Corresponding author. 2004; Reppert and Weaver, 2002; Schibler, 2005).
Tel.: þ0086-731-84805339; Fax: þ0086-731-84805339
E-mail: lijiada@sklmg.edu.cn
In the past decades, the molecular mechanisms

http://dx.doi.org/10.1016/B978-0-444-59427-3.00028-9 119
120

of this endogenous clock in the SCN have been et al., 1974). The vasopressin-containing neurons
elucidated as consisting of several autoregulatory constitute an important subpopulation of the het-
transcriptional/translational feedback loops erogeneous SCN neurons: between 10% and 30%
(Dibner et al., 2010; Lowrey and Takahashi, 2004; of the approximately 10,000 SCN neurons on each
Reppert and Weaver, 2002; Schibler, 2005). How- side of the rat brain may contain AVP (Kalsbeek
ever, how the information in the SCN is mechanis- et al., 2010). The concentration of AVP in the cere-
tically transmitted out of them to regulate the overt brospinal fluid varies in a circadian cycle, with
circadian rhythms is still largely unknown. morning levels approximately five times higher
Dye tracing experiments have revealed that the than those during night hours (Reppert et al.,
primary efferent target areas of the SCN are quite 1981), corresponding to the circadian variation of
limited and predominantly located in the hypothala- AVP content in the SCN (Jansen et al., 2007;
mus and the midline thalamus. Primary SCN target Sodersten et al., 1985; Tominaga et al., 1992; Uhl
areas include lateral septum, bed nucleus of the stria and Reppert, 1986). Molecular studies have rev-
terminalis, subparaventricular zone, para- ealed that the transcription of the AVP gene in
ventricular hypothalamic nucleus, dorsomedial the SCN is rhythmically regulated by the same pos-
hypothalamic nucleus, as well as paraventricular itive and negative elements that control the core
thalamic nucleus (Buijs, 1996; Kalsbeek et al., molecular loops. The CLOCK/BMAL1 complex
1993; Leak and Moore, 2001; Morin and Allen, positively activates the expression of the AVP
2006; Watts and Swanson, 1987; Watts et al., 1987). gene, which is suppressed by the period (Per) and
Lesion and transplant studies have indicated that cryptochrome (Cry) proteins (Jin et al., 1999).
the SCN drive locomotor activity cycles by rhythmi- The role of AVP as an output factor was elegantly
cally releasing factors that act on receptors in the demonstrated by Tousson and Meissl with electro-
hypothalamus near the wall of the third ventricle physiological recordings (Tousson and Meissl,
(Lehman et al., 1987; Ralph et al., 1990). The dem- 2004). They demonstrated circadian rhythms in
onstration that transplanted SCN tissue encaged in spontaneous firing rate in the SCN and adjacent
a semipermeable membrane can still restore a circa- hypothalamic regions, with the activity rhythms
dian rhythm in locomotor activity proved that dif- recorded from structures outside the SCN in phase
fusible factors emanate from the transplanted SCN with the rhythms in the SCN. Cyclic firing in the
(Silver et al., 1996). However, transplantation of hypothalamus was lost after ablation of the SCN
embryonic SCN tissue does not restore the endo- but could be restored by SCN grafts. Further, the
crine and other physiological rhythms, underscoring restoration of a circadian rhythm in the para-
the importance of axonal connections between the ventricular nucleus cocultured with an SCN graft
SCN and their target sites for the regulation of many was abolished by the AVP receptor antagonist.
other circadian processes (Meyer-Bernstein Conversely, periodic application of AVP was able
et al., 1999). In this chapter, we summarize a few to induce firing rhythmicity in the hypothalamus
candidate factors that may transmit the circadian (Tousson and Meissl, 2004).
information from the SCN to other brain areas. However, central infusion of either vasopressin or
a vasopressin V1-antagonist failed to produce signif-
icant effects on the daily patterns of wheel-running
Circadian control of hormone rhythms by behavior (Albers et al., 1984; Stoynev and Nagai,
arginine vasopressin 1996). Moreover, AVP-deficient Brattleboro rats
did not show dramatic changes in the locomotor
Arginine vasopressin (AVP) is one of the first activity rhythms (Groblewski et al., 1981); thus, it
neurotransmitters identified in the SCN (Burlet was supposed that AVP may not be critically
and Marchetti, 1975; Swaab et al., 1975; Vandesande involved in the control of circadian locomotor
121

rhythm. Nevertheless, deficiency in AVP receptor hormone (GnRH) and AVP in cocultures of the
V1a, a dominant AVP receptor that is expressed in preoptic area and the SCN (Funabashi et al.,
the SCN, led to an attenuated circadian rhythmicity 2000). The period of the GnRH circadian rhythm
of locomotor activity (Li et al., 2009). Interestingly, was the same as that of the AVP circadian rhythm.
the rhythmicity of prokineticin 2 (PK2), one of the Administration of AVP significantly increased
candidate output genes, was attenuated in the SCN GnRH release in single preoptic area cultures in
of V1a-deficient mice (Li et al., 2009), suggesting a the presence of estrogen (Funabashi et al., 2000).
role of AVP/V1a signaling in SCN gene expression Further, Clock mutant mice lack the LH surge on
and, subsequently, in the control of behavior and the afternoon of proestrus. Miller et al. showed that
physiology. By using bioluminescence recording, intracerebroventricular (ICV) injection of AVP on
Maywood et al. recently demonstrated that AVP the afternoon of proestrus is sufficient to induce
signaling within the SCN plays a role in the synchro- LH secretion in Clock mutant mice. The effect of
nization of SCN cellular pacemaking (Maywood AVP on the Clock mutant LH surge seems to be
et al., 2011). Recent studies in common voles mediated by V1a receptors, as coinfusion of AVP
also indicate a possible role of SCN vasopressin in and a V1a-specific antagonist prevents the AVP-
circadian locomotor activity. Arrhythmic voles induced LH release (Miller et al., 2006).
exhibited constantly high populations of vasopres-
sin-immunoreactive neurons in the SCN (Jansen
et al., 2007). Further, vasopressin release from The circadian role of PK2 as an output molecule
cultured SCN neurons was attenuated in arrhythmic
voles (Jansen et al., 2000). Prokineticins are a pair of cysteine-rich peptides
The role of AVP in the circadian regulation involved in a variety of biological functions, includ-
of hormone release has been extensively studied. ing gastrointestinal smooth muscle contraction,
Circulating plasma corticosterone levels display a angiogenesis, neurogenesis, nociception, feeding,
circadian rhythm, whereas SCN-lesioned animals and circadian and emotional regulation (Zhou
display constant, but elevated, levels of plasma and Cheng, 2005). Prokineticins execute their
corticosterone. Kalsbeek and colleagues demon- functions by activating two closely related
strated that AVP was able to suppress elevated G-protein-coupled receptors, PKR1 and PKR2
levels of corticosterone in SCN-lesioned animals (Lin et al., 2002; Masuda et al., 2002; Soga et al.,
to basal daytime values. In agreement with this 2002). Activation of these receptors leads to mobi-
suppressive effect of AVP signaling, infusion of lization of calcium, phosphoinositide hydrolysis,
an AVP antagonist induced a sevenfold increase and activation of the mitogen-activated protein
in basal corticosterone levels in intact animals kinase and protein kinase B pathways (Lin et al.,
(Kalsbeek et al., 1992, 1996a,b). 2002; Masuda et al., 2002; Soga et al., 2002).
AVP release from the SCN is probably also The possible link of PK2 with the circadian clock
important for the control of the daily rhythm in was first indicated by the observation that PK2
other hormonal axes, such as the hypothalamic– mRNA levels in the SCN oscillate with a high
pituitary–gonadal axis. SCN lesions lead to basal, amplitude, with high levels during the day and
nonfluctuating luteinizing hormone (LH) levels, low or undetectable levels at night (Cheng et al.,
but a 2-h administration of AVP in the medial 2002). Studies using a bacterial artificial chromo-
preoptic area is sufficient to reinstate a complete some (BAC) transgenic mouse line, in which the
LH surge comparable to the estrogen-induced enhanced green fluorescence protein (EGFP)
surges in SCN-intact animals, both in shape and reporter gene expression was driven by the PK2
amplitude (Palm et al., 1999). Funabashi et al. promoter, indicated a circadian oscillation of a
measured the release of gonadotropin-releasing number of EGFP-positive neurons in the SCN
122

(Zhang et al., 2009). The data from this strain of expression (Hu et al., 2007; Li et al., 2006). The
transgenic mice also revealed EGFP-expressing PKR2/ mice showed a similar attenuation in
neurons in the SCN projected to many known their daily rhythms (Prosser et al., 2007). Further,
SCN target areas, including the ventral lateral clock gene expression is not altered in the SCN of
septum, medial preoptic area, subparaventricular either PK2/ or PKR2/ mice (Li et al., 2006;
zone, paraventricular nucleus, dorsomedial hypo- Prosser et al., 2007), consistent with the supposed
thalamic nucleus, lateral hypothalamic area, and role of PK2 as an output molecule. Moreover,
paraventricular thalamic nucleus (Zhang et al., Morton et al. observed a correlation between
2009). In the same brain areas, the receptor for increased daytime activity and reduced expres-
PK2 (PKR2) was also found to be expressed, as sion of the PK2 molecular rhythm in the SCN of
identified with in situ hybridization (Cheng transgenic mice expressing a mutant Huntington’s
et al., 2002). In situ hybridization studies also rev- gene (Morton et al., 2005). This finding suggests
ealed that PK2 mRNA-positive neurons were that the reduced PK2 rhythm may contribute to
scattered in both the dorsomedial and ventrolat- the sleep disturbances and abnormal circadian
eral SCN (Masumoto et al., 2006). Double label- behavior seen in this strain of transgenic mice.
ing in situ hybridization indicated that more than In addition to expression in the primary SCN
50% of the PK2 mRNA-containing neurons co- targets, PKR2 is also intensively expressed in the
expressed vasoactive intestinal peptide (VIP), SCN neurons, implying a possible signaling role
gastrin-releasing peptide (GRP), or AVP in the of PK2/PKR2 signaling within the SCN. Recently,
SCN (Masumoto et al., 2006). Ren et al. investigated the effect of PK2 on sponta-
The rhythm of PK2 is directly regulated by neous firing and miniature inhibitory postsynaptic
the transcriptional/translational feedback loop of currents (mIPSCs) of SCN neurons using whole-
the central clock (Cheng et al., 2002). Both the cell patch-clamp recordings (Ren et al., 2011).
human and mouse PK2 promoters contain four PK2 dose-dependently increased spontaneous
E-boxes within the first 2 kb upstream of the tran- firing rates in most neurons from the dorsal SCN.
scription start site. In vitro transcription assays Further, PK2 seems to act postsynaptically to
indicate that PK2 transcription is tightly regulated reduce g-aminobutyric acid (GABA)-ergic func-
by clockwork gene products through activation of tion within the SCN (Ren et al., 2011). This study
the E-boxes residing in its proximal promoter. indicated that the local effect of PK2 on the cellular
These in vitro findings were validated in vivo, as activity of SCN neurons may also be linked to
PK2 mRNA expression in the SCN is completely its regulatory role in behavioral and physiological
absent or blunted in mutant mice lacking a func- circadian rhythms.
tional clockwork, including Clock mutant mice
and Cry1/Cry2/ mice (Cheng et al., 2002).
ICV delivery of PK2 into the lateral ventricle TGF-a and CLC are identified as inhibitory
during subjective night, when endogenous PK2 factors by systematic screening
was low, inhibited the nocturnal wheel-running
activity of rats (Cheng et al., 2002). The impor- In their elegant temporal chimera experiments,
tance of PK2/PKR2 signaling in the circadian Vogelbaum and Menaker demonstrated that there
regulation was also demonstrated in mice defi- are both stimulatory and inhibitory output signals
cient in PK2 or PKR2, respectively. The PK2/ from the SCN, at least for locomotor activity
mice showed attenuated circadian rhythmicity in (Vogelbaum and Menaker, 1992). Weitz and
a variety of behaviors and physiology, including coworkers utilized a systematic approach to identify
locomotor activity, sleep/wake, body tempera- peptides that inhibit locomotor activity when infused
ture, hormone as well as peripheral clock gene into the third ventricle (Kramer et al., 2005). As a
123

result of this approach, Weitz et al. reported two Weitz, 2006). However, genetic studies to verify
secreted proteins that could be involved in the gener- the possible role of CLC in circadian rhythms
ation of circadian locomotor rhythmicity. In 2001, have not been reported.
they reported TGF-a as such a candidate output
molecule (Kramer et al., 2001). When TGF-a
was infused into the third ventricle, it reversibly Neuronal firing as an SCN output
inhibited locomotor activity and disrupted circadian
sleep–wake cycles (Kramer et al., 2001). The mRNA The SCN neurons express a cell autonomous circa-
of TGF-a is expressed rhythmically in the SCN, with dian rhythm in spontaneous action potential fre-
peak and trough levels around subjective dawn and quency (Kuhlman and McMahon, 2006). In vivo
dusk, respectively (Kramer et al., 2001). However, recordings revealed circadian rhythms in the elec-
in contrast to the circadian rhythms of TGF-a trical activity in the SCN, with spike activity being
mRNA levels in the hamster SCN, there is no signif- high during the day and low at night (Inouye and
icant circadian change in the levels of TGF-a protein Kawamura, 1979). Rhythmic SCN electrical activ-
in the hamster SCN (Lindley et al., 2008). TGF- ity is likely a critical component of SCN output.
a-positive cells were found throughout the SCN, Several ion channels have been identified in the
but more abundantly in the core than the shell SCN. One class of fast delayed rectifier (FDR)
area, while AVP was predominantly found in the channels, the Kv3.1/3.2 Kþ channels, is rhythmically
shell. Fluorescent double labeling revealed a total expressed in the SCN, with its members’ expression
lack of coexpression for the two proteins in SCN peaking during the day phase. Blocking the FDR
cells. Further, TGF-a was found to be colocalized potassium currents appeared to block the daily
with glial fibrillary acidic protein (GFAP), but not rhythm in firing rate in SCN neurons (Itri et al.,
with the general neuronal marker NeuN, or 2005). Further, the iberiotoxin-sensitive BK
calbindin-D28K (Li et al., 2002; Lindley et al., channels are rhythmically expressed in the SCN,
2008). Thus, the role of TGF-a in the circadian regu- and the daily expression of BK channel in the
latory system remains controversial. SCN is controlled by the intrinsic circadian clock
In 2006, Weitz and coworkers identified car- (Meredith et al., 2006). The mRNA level of the tran-
diotrophin-like cytokine (CLC) as another candi- script encoding for BK channel KCNMA1 is high
date circadian regulator of locomotor activity during the night phase, so is the BK current. Both
(Kraves and Weitz, 2006). When CLC was pharmacological blockade of BK current and
acutely infused into the third ventricle, it pro- genetic knockout of the channel gene (Kcnma1/)
duced a transient blockade of locomotor activity blunt, but do not completely block, the circadian
without affecting the core molecular oscillation rhythm in spike frequency (Meredith et al., 2006).
(Kraves and Weitz, 2006). In the mouse, CLC is Specifically, Kcnma1/ mice have increased spon-
expressed in a subpopulation of SCN vasopressin taneous firing rates in SCN neurons at night, imply-
neurons with a circadian rhythm that peaks dur- ing that the nightly BK current might reduce SCN
ing the late day, a daily period with locomotor spike frequency. Kcnma1/ mice also displayed
quiescence. The rhythm of CLC seems directly reduced circadian amplitudes in multiple behaviors
regulated by the transcriptional/tranlational nega- timed by the SCN, including locomotor activity
tive feedback loop of the central clock (Kraves and core body temperature. Further, Kcnma1/
and Weitz, 2006).The CLC receptors flank the mice showed normal expression of clock genes such
third ventricle, and the hypothalamic infusion of as Bmal1, consistent with a role of BK channels
neutralizing antibodies to the CLC receptor pro- in SCN output, rather than in the intrinsic time-
duces extra daily locomotor activity at a time keeping mechanism of the master pacemaker
when CLC is maximally expressed (Kraves and (Meredith et al., 2006).
124

Thermal regulation

PK2
Sleep-wake

B C
AVP Hormone
E

CLC Locomotor activity

Fig. 1. A diagram depicting the relationship between three putative circadian output molecules. The transcription of PK2, AVP,
and CLC is regulated by Bmal1/Clock heterodimer by binding to the E-box located at their promoters. However, the PK2,
AVP, and CLC mRNA levels peak at approximately CT 3, CT 6, and CT 9, respectively. In addition, they influence different
behavioral and physiological rhythms.

Perspectives rodents. However, as it seems that PK2, AVP,


and TGF-a expression is similar in diurnal and
Several candidate signaling factors have been nocturnal animals (Lambert et al., 2005; Mahoney
reported as output signals connecting the SCN et al., 2009; Tournier et al., 2007), there is no evi-
and their target areas. Temporally, these factors dence indicating that a difference in the phase of
might exert overlapping functions, as knockout output gene expression accounts for diurnality.
mice for a single output gene or blockade of a Although the temporal chimera experiments
single pathway did not result in complete carried out by Vogelbaum and Menaker proposed
arrhythmicity. PK2, AVP, and CLC mRNA peak the existence of both stimulatory and inhibitory
at approximately circadian time 3 (CT 3), CT 6, signals in the SCN (Vogelbaum and Menaker,
and CT 9, respectively (Cheng et al., 2002; Jin 1992), no experiments have thus far proposed
et al., 1999; Kraves and Weitz, 2006). Thus, the SCN factors that promote arousal or locomotor
dawn and dusk time may represent unique zones activity. Further, most of the research appeared
for PK2 and CLC, respectively (Fig. 1). Defi- to focus on locomotor activity. As the SCN also
ciency in PK2 resulted in increased locomotor regulate many other physiological and neuroendo-
activity, wakefulness, body temperature, as well crine rhythms, it might be worthwhile to investi-
as corticosterone levels during the early morning gate the other putative output pathways.
period (Li et al., 2006), whereas blocking CLC
signaling by infusion of anti-gp130 caused an
increase in locomotor activity specifically during Acknowledgments
CT 9–12 (Kraves and Weitz, 2006).
The circadian rhythms of neural, metabolic, neu- The research in Jia-Da Li’s lab was partly supported
rotransmitter, and clock gene expression in the by the Projects in the Major State Basic Research
SCN neurons are similarly phased in nocturnal Development Program of China (973 Program)
and diurnal animals. It is thus challenging to iden- (grant number 2012CB517904), National Natural
tify the switch responsible for diurnality. Some Science Foundation of China (grants 30970958
research has been performed to see if there is and 81070481), and the Lotus Scholar Professorship
any difference in the output genes in the diurnal Funds from Hunan Province Government, China.
125

The research in Qun-Yong Zhou’s lab was Dibner, C., Schibler, U., & Albrecht, U. (2010). The mamma-
supported in part by grants from NIH (NIMH67753, lian circadian timing system: Organization and coordination
of central and peripheral clocks. Annual Review of Physiology,
HD15788) and the American Heart Association 72, 517–549.
(855156 F). Funabashi, T., Shinohara, K., Mitsushima, D., & Kimura, F.
(2000). Gonadotropin-releasing hormone exhibits circadian
rhythm in phase with arginine-vasopressin in co-cultures of
Abbreviations the female rat preoptic area and suprachiasmatic nucleus.
Journal of Neuroendocrinology, 12(6), 521–528.
AVP arginine vasopressin Groblewski, T. A., Nunez, A. A., & Gold, R. M. (1981). Circa-
BAC bacterial artificial chromosome dian rhythms in vasopressin deficient rats. Brain Research
Bulletin, 6(2), 125–130.
CLC cardiotrophin-like cytokine Hu, W. P., Li, J. D., Zhang, C., Boehmer, L., Siegel, J. M., &
Cry cryptochrome Zhou, Q. Y. (2007). Altered circadian and homeostatic sleep
CT circadian time regulation in prokineticin 2-deficient mice. Sleep, 30(3),
EGFP enhanced green fluorescence protein 247–256.
FDR fast delayed rectifier Inouye, S. T., & Kawamura, H. (1979). Persistence of circadian
rhythmicity in a mammalian hypothalamic “island” con-
GABA g-aminobutyric acid taining the suprachiasmatic nucleus. Proceedings of the
GFAP glial fibrillary acidic protein National Academy of Sciences of the United States of America,
GnRH gonadotropin-releasing hormone 76(11), 5962–5966.
GRP gastrin-releasing peptide Itri, J. N., Michel, S., Vansteensel, M. J., Meijer, J. H., &
ICV intracerebroventricular Colwell, C. S. (2005). Fast delayed rectifier potassium
current is required for circadian neural activity. Nature Neu-
LH luteinizing hormone roscience, 8(5), 650–656.
mIPSCs miniature inhibitory postsynaptic Jansen, K., Van der Zee, E. A., & Gerkema, M. P. (2000).
currents Being circadian or not: Vasopressin release in cultured SCN
Per period mirrors behavior in adult voles. Neuroreport, 11(16),
PK2 prokineticin 2 3555–3558.
Jansen, K., Van der Zee, E. A., & Gerkema, M. P. (2007). Vaso-
SCN suprachiasmatic nuclei pressin immunoreactivity, but not vasoactive intestinal poly-
VIP vasoactive intestinal peptide peptide, correlates with expression of circadian rhythmicity
in the suprachiasmatic nucleus of voles. Neuropeptides, 41
(4), 207–216.
Jin, X., Shearman, L. P., Weaver, D. R., Zylka, M. J., de
References Vries, G. J., & Reppert, S. M. (1999). A molecular mecha-
nism regulating rhythmic output from the suprachiasmatic
Albers, H. E., Ferris, C. F., Leeman, S. E., & Goldman, B. D. circadian clock. Cell, 96(1), 57–68.
(1984). Avian pancreatic polypeptide phase shifts hamster cir- Kalsbeek, A., Buijs, R. M., van Heerikhuize, J. J., Arts, M., &
cadian rhythms when microinjected into the suprachiasmatic van der Woude, T. P. (1992). Vasopressin-containing
region. Science, 223(4638), 833–835. neurons of the suprachiasmatic nuclei inhibit corticosterone
Buijs, R. M. (1996). The anatomical basis for the expression release. Brain Research, 580(1–2), 62–67.
of circadian rhythms: The efferent projections of the Kalsbeek, A., Fliers, E., Hofman, M. A., Swaab, D. F., &
suprachiasmatic nucleus. Progress in Brain Research, 111, Buijs, R. M. (2010). Vasopressin and the output of the hypo-
229–240. thalamic biological clock. Journal of Neuroendocrinology, 22
Burlet, A., & Marchetti, J. (1975). Immunoreactive vasopres- (5), 362–372.
sin in the supra-chiasmatic nucleus. Preliminary data in rats. Kalsbeek, A., Teclemariam-Mesbah, R., & Pevet, P. (1993).
Comptes rendus des séances de la Société de biologie et de ses Efferent projections of the suprachiasmatic nucleus in the
filiales, 169(1), 148–151. golden hamster (Mesocricetus auratus). The Journal of
Cheng, M. Y., Bullock, C. M., Li, C., Lee, A. G., Bermak, J. C., Comparative Neurology, 332(3), 293–314.
Belluzzi, J., et al. (2002). Prokineticin 2 transmits the behav- Kalsbeek, A., van der Vliet, J., & Buijs, R. M. (1996).
ioural circadian rhythm of the suprachiasmatic nucleus. Decrease of endogenous vasopressin release necessary for
Nature, 417(6887), 405–410. expression of the circadian rise in plasma corticosterone:
126

A reverse microdialysis study. Journal of Neuroendocrinology, Lindley, J., Deurveilher, S., Rusak, B., & Semba, K. (2008).
8(4), 299–307. Transforming growth factor-alpha and glial fibrillary acidic
Kalsbeek, A., van Heerikhuize, J. J., Wortel, J., & Buijs, R. M. protein in the hamster circadian system: Daily profile and
(1996). A diurnal rhythm of stimulatory input to cellular localization. Brain Research, 1197, 94–105.
the hypothalamo-pituitary-adrenal system as revealed by Lowrey, P. L., & Takahashi, J. S. (2004). Mammalian circadian
timed intrahypothalamic administration of the vasopressin biology: Elucidating genome-wide levels of temporal organi-
V1 antagonist. The Journal of Neuroscience, 16(17), zation. Annual Review of Genomics and Human Genetics, 5,
5555–5565. 407–441.
Kramer, A., Yang, F. C., Kraves, S., & Weitz, C. J. (2005). Mahoney, M. M., Ramanathan, C., Hagenauer, M. H.,
A screen for secreted factors of the suprachiasmatic nucleus. Thompson, R. C., Smale, L., & Lee, T. (2009). Daily
Methods in Enzymology, 393, 645–663. rhythms and sex differences in vasoactive intestinal polypep-
Kramer, A., Yang, F. C., Snodgrass, P., Li, X., tide, VIPR2 receptor and arginine vasopressin mRNA in the
Scammell, T. E., Davis, F. C., et al. (2001). Regulation of suprachiasmatic nucleus of a diurnal rodent, Arvicanthis
daily locomotor activity and sleep by hypothalamic EGF niloticus. The European Journal of Neuroscience, 30(8),
receptor signaling. Science, 294(5551), 2511–2515. 1537–1543.
Kraves, S., & Weitz, C. J. (2006). A role for cardiotrophin-like Masuda, Y., Takatsu, Y., Terao, Y., Kumano, S., Ishibashi, Y.,
cytokine in the circadian control of mammalian locomotor Suenaga, M., et al. (2002). Isolation and identification of
activity. Nature Neuroscience, 9(2), 212–219. EG-VEGF/prokineticins as cognate ligands for two orphan
Kuhlman, S. J., & McMahon, D. G. (2006). Encoding the ins G-protein-coupled receptors. Biochemical and Biophysical
and outs of circadian pacemaking. Journal of Biological Research Communications, 293(1), 396–402.
Rhythms, 21(6), 470–481. Masumoto, K. H., Nagano, M., Takashima, N., Hayasaka, N.,
Lambert, C. M., Machida, K. K., Smale, L., Nunez, A. A., & Hiyama, H., Matsumoto, S., et al. (2006). Distinct localiza-
Weaver, D. R. (2005). Analysis of the prokineticin 2 system tion of prokineticin 2 and prokineticin receptor 2 mRNAs
in a diurnal rodent, the unstriped Nile grass rat (Arvicanthis in the rat suprachiasmatic nucleus. The European Journal
niloticus). Journal of Biological Rhythms, 20(3), 206–218. of Neuroscience, 23(11), 2959–2970.
Leak, R. K., & Moore, R. Y. (2001). Topographic organization Maywood, E. S., Chesham, J. E., O’Brien, J. A., &
of suprachiasmatic nucleus projection neurons. The Journal Hastings, M. H. (2011). A diversity of paracrine signals
of Comparative Neurology, 433(3), 312–334. sustains molecular circadian cycling in suprachiasmatic nucleus
Lehman, M. N., Silver, R., Gladstone, W. R., Kahn, R. M., circuits. Proceedings of the National Academy of Sciences of
Gibson, M., & Bittman, E. L. (1987). Circadian rhythmicity the United States of America, 108(34), 14306–14311.
restored by neural transplant. Immunocytochemical charac- Meredith, A. L., Wiler, S. W., Miller, B. H., Takahashi, J. S.,
terization of the graft and its integration with the host brain. Fodor, A. A., Ruby, N. F., et al. (2006). BK calcium-
The Journal of Neuroscience, 7(6), 1626–1638. activated potassium channels regulate circadian behavioral
Li, J. D., Burton, K. J., Zhang, C., Hu, S. B., & Zhou, Q. Y. rhythms and pacemaker output. Nature Neuroscience, 9(8),
(2009). Vasopressin receptor V1a regulates circadian 1041–1049.
rhythms of locomotor activity and expression of clock- Meyer-Bernstein, E. L., Jetton, A. E., Matsumoto, S. I.,
controlled genes in the suprachiasmatic nuclei. American Markuns, J. F., Lehman, M. N., & Bittman, E. L. (1999).
Journal of Physiology. Regulatory, Integrative and Compara- Effects of suprachiasmatic transplants on circadian rhythms
tive Physiology, 296(3), R824–R830. of neuroendocrine function in golden hamsters. Endocrinol-
Li, J. D., Hu, W. P., Boehmer, L., Cheng, M. Y., Lee, A. G., ogy, 140(1), 207–218.
Jilek, A., et al. (2006). Attenuated circadian rhythms in mice Miller, B. H., Olson, S. L., Levine, J. E., Turek, F. W.,
lacking the prokineticin 2 gene. The Journal of Neurosci- Horton, T. H., & Takahashi, J. S. (2006). Vasopressin regu-
ence, 26(45), 11615–11623. lation of the proestrous luteinizing hormone surge in wild-
Li, X., Sankrithi, N., & Davis, F. C. (2002). Transforming type and Clock mutant mice. Biology of Reproduction, 75
growth factor-alpha is expressed in astrocytes of the (5), 778–784.
suprachiasmatic nucleus in hamster: Role of glial cells in cir- Morin, L. P., & Allen, C. N. (2006). The circadian visual sys-
cadian clocks. Neuroreport, 13(16), 2143–2147. tem, 2005. Brain Research Reviews, 51(1), 1–60.
Lin, D. C., Bullock, C. M., Ehlert, F. J., Chen, J. L., Tian, H., Morton, A. J., Wood, N. I., Hastings, M. H., Hurelbrink, C.,
& Zhou, Q. Y. (2002). Identification and molecular charac- Barker, R. A., & Maywood, E. S. (2005). Disintegration of
terization of two closely related G protein-coupled receptors the sleep-wake cycle and circadian timing in Huntington’s
activated by prokineticins/endocrine gland vascular endo- disease. The Journal of Neuroscience, 25(1), 157–163.
thelial growth factor. The Journal of Biological Chemistry, Palm, I. F., Van Der Beek, E. M., Wiegant, V. M.,
277(22), 19276–19280. Buijs, R. M., & Kalsbeek, A. (1999). Vasopressin induces
127

a luteinizing hormone surge in ovariectomized, estradiol- Swaab, D. F., Pool, C. W., & Nijveldt, F. (1975). Immuno-
treated rats with lesions of the suprachiasmatic nucleus. fluorescence of vasopressin and oxytocin in the rat
Neuroscience, 93(2), 659–666. hypothalamo-neurohypophypopseal system. Journal of
Prosser, H. M., Bradley, A., Chesham, J. E., Ebling, F. J., Neural Transmission, 36(3–4), 195–215.
Hastings, M. H., & Maywood, E. S. (2007). Prokineticin Tominaga, K., Shinohara, K., Otori, Y., Fukuhara, C., &
receptor 2 (Prokr2) is essential for the regulation of circa- Inouye, S. T. (1992). Circadian rhythms of vasopressin con-
dian behavior by the suprachiasmatic nuclei. Proceedings tent in the suprachiasmatic nucleus of the rat. Neuroreport, 3
of the National Academy of Sciences of the United States of (9), 809–812.
America, 104(2), 648–653. Tournier, B. B., Dardente, H., Vuillez, P., Pevet, P., &
Ralph, M. R., Foster, R. G., Davis, F. C., & Menaker, M. Challet, E. (2007). Expression of Tgfalpha in the
(1990). Transplanted suprachiasmatic nucleus determines suprachiasmatic nuclei of nocturnal and diurnal rodents.
circadian period. Science, 247(4945), 975–978. Neuroscience, 145(3), 1138–1143.
Ren, P., Zhang, H., Qiu, F., Liu, Y. Q., Gu, H., Tousson, E., & Meissl, H. (2004). Suprachiasmatic nuclei
O’Dowd, D. K., et al. (2011). Prokineticin 2 regulates the grafts restore the circadian rhythm in the paraventricular
electrical activity of rat suprachiasmatic nuclei neurons. nucleus of the hypothalamus. The Journal of Neuroscience,
PLoS One, 6(6), e20263. 24(12), 2983–2988.
Reppert, S. M., Artman, H. G., Swaminathan, S., & Uhl, G. R., & Reppert, S. M. (1986). Suprachiasmatic nucleus
Fisher, D. A. (1981). Vasopressin exhibits a rhythmic daily vasopressin messenger RNA: Circadian variation in normal
pattern in cerebrospinal fluid but not in blood. Science, 213 and Brattleboro rats. Science, 232(4748), 390–393.
(4513), 1256–1257. Vandesande, F., DeMey, J., & Dierickx, K. (1974). Identifica-
Reppert, S. M., & Weaver, D. R. (2002). Coordination of cir- tion of neurophysin producing cells. I. The origin of the
cadian timing in mammals. Nature, 418(6901), 935–941. neurophysin-like substance-containing nerve fibres of
Schibler, U. (2005). The daily rhythms of genes, cells and the external region of the median eminence of the rat. Cell
organs. Biological clocks and circadian timing in cells. and Tissue Research, 151(2), 187–200.
EMBO Reports, 6(Spec No), S9–S13. Vogelbaum, M. A., & Menaker, M. (1992). Temporal
Silver, R., LeSauter, J., Tresco, P. A., & Lehman, M. N. chimeras produced by hypothalamic transplants. The Jour-
(1996). A diffusible coupling signal from the transplanted nal of Neuroscience, 12(9), 3619–3627.
suprachiasmatic nucleus controlling circadian locomotor Watts, A. G., & Swanson, L. W. (1987). Efferent projections of
rhythms. Nature, 382(6594), 810–813. the suprachiasmatic nucleus: II. Studies using retrograde
Sodersten, P., De Vries, G. J., Buijs, R. M., & Melin, P. (1985). transport of fluorescent dyes and simultaneous peptide
A daily rhythm in behavioral vasopressin sensitivity and immunohistochemistry in the rat. The Journal of Compara-
brain vasopressin concentrations. Neuroscience Letters, 58 tive Neurology, 258(2), 230–252.
(1), 37–41. Watts, A. G., Swanson, L. W., & Sanchez-Watts, G. (1987).
Soga, T., Matsumoto, S., Oda, T., Saito, T., Hiyama, H., Efferent projections of the suprachiasmatic nucleus: I. Stud-
Takasaki, J., et al. (2002). Molecular cloning and characteri- ies using anterograde transport of Phaseolus vulgaris
zation of prokineticin receptors. Biochimica et Biophysica leucoagglutinin in the rat. The Journal of Comparative Neu-
Acta, 1579(2–3), 173–179. rology, 258(2), 204–229.
Stoynev, A. G., & Nagai, K. (1996). Lack of effect Zhang, C., Truong, K. K., & Zhou, Q. Y. (2009). Efferent
of suprachiasmatic infusion of a vasopressin antagonist projections of prokineticin 2 expressing neurons in the
on the circadian rhythm of wheel-running activity in rats. mouse suprachiasmatic nucleus. PLoS One, 4(9), e7151.
Acta Physiologica et Pharmacologica Bulgarica, 22(2), Zhou, Q. Y., & Cheng, M. Y. (2005). Prokineticin 2 and circa-
39–43. dian clock output. The FEBS Journal, 272(22), 5703–5709.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 8

Suprachiasmatic nucleus: Cellular


clocks and networks

Sato Honma{,{,*, Daisuke Ono{,}, Yohko Suzuki}, Natsuko Inagaki{,1,


Tomoko Yoshikawa}, Wataru Nakamura{,} and Ken-ichi Honma{

{
Department of Physiology, Hokkaido University Graduate School of Medicine, Sapporo, Japan
{
Department of Chronomedicine, Hokkaido University Graduate School of Medicine, Sapporo, Japan
}
Advanced Photonic Bioimaging Section, Research Center for Cooperative Projects, Hokkaido University
Graduate School of Medicine, Sapporo, Japan
}
Department of Chronodentistry, Osaka University Graduate School of Dentistry, Suita, Osaka, Japan

Abstract: The suprachiasmatic nucleus (SCN), the master circadian clock of mammals, is composed of
multiple circadian oscillator neurons. Most of them exhibit significant circadian rhythms in their clock
gene expression and spontaneous firing when cultured in dispersed cells, as well as in an organotypic
slice. The distribution of periods depends on the SCN tissue organization, suggesting that cell-to-cell
interaction is important for synchronization of the constituent oscillator cells. This cell-to-cell interaction
involves both synaptic interactions and humoral mediators. Cellular oscillators form at least three
separate but mutually coupled regional pacemakers, and two of them are involved in the photoperiodic
regulation of behavioral rhythms in mice. Coupling of cellular oscillators in the SCN tissue compensates
for the dysfunction due to clock gene mutations, on the one hand, and desynchronization within and
between the regional pacemakers that suppresses the coherent rhythm expression from the SCN, on the
other hand. The multioscillator pacemaker structure of the SCN is advantageous for responding to a
wide range of environmental challenges without losing coherent rhythm outputs.
Keywords: suprachiasmatic nucleus; neural network; clock genes; multi-electrode array; oscillatory
coupling; photoperiods; bioluminescence imaging.

*Corresponding author.
Tel.: þ81-11-706-4778; Fax: þ81-11-706-4737
E-mail: sathonma@med.hokudai.ac.jp
Introduction

1
Present address: Laboratory of Cell Growth and Differentia- The suprachiasmatic nucleus (SCN), the master
tion Institute of Molecular and Cellular Biosciences, The circadian clock of mammals, is composed of about
University of Tokyo, Tokyo, Japan. 20,000 neurons in rats and mice (Abrahamson

http://dx.doi.org/10.1016/B978-0-444-59427-3.00029-0 129
130

and Moore, 2001; van den Pol, 1980). Monitoring 2004; Welsh et al., 1995). In the SCN neurons of
spontaneous firing activity from SCN tissue wild-type Wistar rats, circadian periods of sponta-
cultured on a multi-electrode array dish (MED) neous firing range from 20 to 30 h when cultured
showed that most of the neurons exhibited robust as enzymatically dispersed ones, whereas they
circadian firing rhythms with periodicities slightly range from 22 to 27 h when the SCN neurons
different from each other (Herzog et al., 1997; are cultured as an organotypic coronal slice
Honma et al., 1998; Welsh et al., 1995). These (Honma et al., 2004; Fig. 1). Periods from both
results suggest that each SCN neuron contains a cultures followed a Gaussian distribution. The
cellular circadian oscillator. The current under- mean periods were not statistically different, but
standing of the molecular core of the cellular the distribution range was substantially different
oscillator involves a transcriptional–translational between the two cultures. The circadian periods
feedback loop in which heterodimeric transcrip- are well synchronized within a cultured SCN slice,
tion factors, CLOCK and BMAL1, activate tran- where cell-to-cell interactions are much stronger
scription of Period (Per) and Cryptochrome than those in the dispersed cell culture. In both
(Cry) genes and the protein products, PERs and cultures, the average periods are close to the free-
CRYs, in turn, suppress their own transcription, running period of animals’ behavior, suggesting
thus closing the auto-feedback loop, whose single that mutual synchronization among the constituent
turn takes about 24 h (Reppert and Weaver, cellular oscillators in the SCN determines the
2002). Robust circadian rhythms in the transcrip- rhythm output from the SCN for behavioral
tional activity of most clock genes and in the pro- rhythms (Fig. 1a; Honma et al., 2004). Considering
tein levels are advantageous for monitoring the the narrow distribution of free-running periods in
tick of the circadian clock by bioluminescence the behavioral rhythms, intact neural networks and
and fluorescent reporters (Kuhlman et al., 2000; close cell–cell apposition for efficient paracrine sig-
Yamazaki et al., 2000). This recent progress naling seem to effectively increase the coupling
in molecular neurobiology enabled long-term, intensity among oscillator cells in the SCN in vivo.
real-time monitoring of multiple functions from Further, input signals from extra SCN tissues may
hundreds of SCN neurons simultaneously. Com- also affect the stability of periodicity. Interestingly,
parison of circadian rhythms of different functions at a small percentage, anti-phasic firing rhythms
from a single dispersed SCN cell makes it possible were observed in a cultured SCN slice, indicating
to analyze intracellular circadian signal transduc- a large diversity of neuronal outputs from the pace-
tion. Further, long-term and simultaneous moni- maker (Nakamura et al., 2001).
toring of different circadian rhythms from single We also analyzed circadian firing rhythms of
cells of an SCN explant revealed novel roles of single SCN neurons in the slice and dispersed
the oscillatory networks. Dependence of regions cell cultures of C57BL/6 J mice (Fig. 1b). The
in the SCN and developmental stages is also of results were basically the same as those for rats
special interest. Here, we review cellular and net- except for the periods, which were slightly
work properties of the SCN circadian rhythms shorter than the respective period for rats. In
revealed by these new rhythm monitoring tools. both rats and mice, mean circadian period of
behavioral rhythm was slightly but significantly
longer than the respective SCN neurons. The
Cellular oscillators and coupling reason for the difference is not known. Although
the circadian period is systematically changed
Enzymatically isolated SCN neurons of rats and depending on the environmental light in con-
mice exhibit circadian rhythms in spontaneous stant conditions, the periodicity in the behavioral
firing and clock gene expression (Herzog et al., rhythms is not due to the environmental light,
131

(a) rats (Wistar) (b) mice (C57BL/6J)


30 15
23.1±1.9 h

Number of neurons
dispersed

Number of neurons
25 disperse 24.1±1.5 h
72.4% (223/308) 98% (121/123)
20 10
15
10 5
5
0 0
16 18 20 22 24 26 28 30 16 18 20 22 24 26 28 30
circadian period (h) circadian period (h)
70 15
24.2±0.7 h slice
Number of neurons

slice

Number of neurons
60 23.3±0.4 h
82.6% (185 / 224)
50 100% (n=53)
10
40
30
20 5
10
0 0
16 18 20 22 24 26 28 30 16 18 20 22 24 26 28 30
circadian period (h) circadian period (h)
80 60
behavior behavior
Number of animals

Number of animals

24.4±0.2 h 50 23.8±0.1 h
60 (n=86)
(n=147) 40
40 30
20
20
10
0 0
16 18 20 22 24 26 28 30 16 18 20 22 24 26 28 30
circadian period (h) circadian period (h)

Fig. 1. Distribution of circadian periods of spontaneous firing rhythms in individual SCN neurons and behavior. Data show the
distribution of circadian periods of spontaneous firing rhythms in rats (a) and mice (b) together with that of behavioral rhythms.
Upper, middle, and lower panels demonstrate period distribution of SCN neurons in dispersed cell culture, that in organotypic
slice culture, and spontaneous behavior, respectively. Behavioral rhythms were monitored by an infrared thermal sensor from
blinded rats and mice kept in constant darkness. Numbers in each panel indicate average period (SD) and percentage of
neurons with significant circadian periodicity. Numbers in parentheses are those of SCN neurons with significant rhythms out of
the total number examined.

because behavioral rhythms were monitored Pacemaker composed of oscillator and


from blinded rats and mice kept in constant non-oscillator cells
darkness. Since the SCN had to be sliced for cul-
turing, we might have lost some components of The percentage of neurons exhibiting a significant
the SCN, which would have lengthened the peri- circadian rhythm in a culture depends on the cell
odicity during the culture preparation. Alterna- density and cellular architecture (Herzog et al.,
tively, some factors in the culture media might 2004; Nakamura et al., 2002), suggesting that not
affect the periodicity in vitro. every SCN neuron is a cell-autonomous clock cell.
132

The percentage of neurons exhibiting a circadian (a)


20
firing rhythm was significantly higher in slice dispersed 22.8±2.1 h
cultures than that in dispersed cell cultures 100% (n=168)
15

Number of cells
(Honma et al., 2004). Luciferase reporters and
bioluminescence imaging of clock gene expres-
10
sion made the analysis of single cell rhythms
much easier compared to that of firing rhythms.
In addition, cellular rhythms can now be analyzed 5
in adult animals that have been exposed to
different experimental conditions. Using an SCN 0
16 18 20 22 24 26 28 30
slice culture from transgenic mice carrying a circadian period (h)
mouse Per1 promoter-driven luciferase reporter (b)
50
gene (Per1-luc) (Inagaki et al., 2007), we com- newborn slice 24.0±0.3 h
pared circadian periodicities of single SCN cells 40 100% (n=109)

Number of cells
in the slice and those in dispersed cell cultures.
30
All bioluminescent cells exhibited significant cir-
cadian Per1-luc rhythms in dispersed cell cultures 20
as well as in organotypic slice cultures (Fig. 2). In
the slice culture, circadian periods of single SCN 10
cells of adult mice were distributed in a wider
0
range than those of newborns. The reason for this 16 18 20 22 24 26 28 30
difference is not known. Culture condition might circadian period (h)
have affected the period distribution, since new- (c) 50
adult slice 23.7±0.9 h
born SCN slices were precultured 1–2 weeks
40 100% (n=108)
before the start of recording, which might have
Number of cells

promoted the network formation within a slice. 30


On the other hand, neural networks might have
been dissected by the slice preparation and never 20
recovered in adult SCN.
10
The percentage of neurons with significant
circadian periods depends on the cell density in 0
dispersed cell cultures: the higher the density, 16 18 20 22 24 26 28 30
the higher the percentage of rhythmic neurons. circadian period (h)
Using dispersed SCN cells from mice carrying
Fig. 2. Distribution of circadian periods in Per1-luc rhythms
a PERIOD2::LUCIFERASE fusion protein as depends on culture system and age. Data show the
a real-time reporter (PER2::LUC, Yoo et al., distribution of circadian periods of Per1-luc rhythms in
2005), Webb et al. (2009) demonstrated that 71% individual mouse SCN cells. Distribution in dispersed cell
of dispersed SCN cells did not show PER2::LUC culture of newborns (a), slice culture of newborns (b), and
rhythms longer than 3 days when cultured in a adults (c) are shown. For the details, see the legend of Fig. 1.
very low density of 100 cells/mm2. Among the
remaining cells, 65% were rhythmic with intrinsic rhythm-positive and -negative cells, suggesting that
circadian periods. They also reported that the neither the AVP nor the VIP cells are solely the
marker neuropeptides of the SCN, vasopressin oscillator cells (Webb et al., 2009). Surprising is a
(AVP) and vasoactive intestinal polypeptide (VIP), relatively small number of rhythm-positive cells
were detected at almost the same percentage in (19%) in the culture with almost no physical contact.
133

Mathematical models predict a stable oscillation Clock gene expression rhythms in the ventrolateral
by coupling among cells with self-sustained and SCN immediately shifted to the new LD cycle,
damped oscillators (Abraham et al., 2010; Ko whereas those in the dorsomedial SCN gradually
et al., 2010). Therefore, the oscillatory coupling in shifted with several days of transients. These
the intact SCN seems to be a specific character of data not only confirmed the uncoupling of two
the central clock rather than the oscillation within regional oscillators seen previously in vitro but
single neurons. also suggested a functional significance of these
oscillators.
Recently, we found at least three photoperiodi-
Regional oscillators in the SCN for cally regulated oscillator cell groups in the SCN
different functions by measuring Per1-expressing rhythms. Further,
two of them were phase-locked to the onset and
An organotypic slice culture of the SCN enables offset of behavioral activity separately (Inagaki
us to analyze networks and regional specificity et al., 2007). Animals living in the middle and
within the SCN. Previously, we reported desyn- high latitudes change their behavioral rhythms
chronization between circadian rhythms in AVP depending on season, which is important for their
and VIP releases (Shinohara et al., 1995). We survival in different seasons. The long-lasting
measured the concentration of the two peptides hypothesis for photoperiodic regulation of behav-
in the culture medium of an SCN slice cultured ioral rhythms of nocturnal rodents consists of
in a roller tube every 2 h, and demonstrated two separate, but mutually coupled, oscillators:
robust circadian rhythms in both peptides. They evening (E) and morning (M) oscillators, which
were synchronized in the control cultures, but in regulate activity onset and offset, respectively
the cultures treated with anti-mitotics in the begin- (Pittendrigh and Daan, 1976b). We exposed
ning of culturing, the rhythms of the two peptides Per1-luc mice, which were reared under a LD
free-ran with different periods and eventually cycle of 12:12 h, to either short day (LD 6:18) or
desynchronized with each other during the culture. long day (LD 18:6) conditions for 3 weeks. Two
This result suggests that the SCN contains at least serial coronal SCN slices of 300-mm thickness
two regional pacemakers, one composed of AVP- were made and their Per1-luc bioluminescence
releasing neurons and the other composed of was measured for more than 5 days (Fig. 3). In
VIP-releasing neurons. The two pacemakers were mice kept under LD 12:12 and LD 6:18, the circa-
uncoupled by treatment with anti-mitotics; never- dian peak of the Per1-luc rhythm in the anterior
theless, synchronization among cellular oscillators SCN always phase-lagged compared with that in
of each peptidergic neuron was kept basically the posterior counterpart. The phase difference
intact. We still do not know the mechanism for was smaller in LD 6:18 than in LD 12:12. Surpris-
the desynchronization by anti-mitotics. Anti- ingly, only in the anterior SCN of mice under LD
mitotic treatment prevents glial overgrowth; thus, 18:6, Per1-luc exhibited two peaks per day, which
treated SCN slices tended to extend, whereas tended to merge during culturing (Fig. 3). Biolumi-
untreated slice remained rather compact, which nescence imaging revealed that this bimodal Per1-
might promote better physical contact. Alterna- luc pattern was due to two differentially peaking
tively, the glia might exert some functions in cell groups with a period of about 24 h (Inagaki
rhythm coupling among SCN neurons mechani- et al., 2007). Since the morning peak in the anterior
cally or chemically. A similar desynchronization SCN is expressed only in the long photoperiods,
between the dorsal and ventral SCN was also the evening Per1-luc peak was selected for compar-
observed in the rat SCN in vivo after an abrupt shift ison with those in other photoperiods. We found
in the light–dark (LD) cycle (Nagano et al., 2003). that, irrespective of the photoperiods, the anterior
134

anterior SCN
posterior SCN Time of day (hr)
(a)
20 0 4 8 12 16 20 24
RLU (⫻102 counts/min)

Days in culture (day)


10 1

0 2

3
-10
4
-20
12 24 12 24 12 24 12
Time of day (hr)
(b) Time of day (hr)
20 0 4 8 12 16 20 24
RLU (⫻102 counts/min)

Days in culture (day)


10 1

2
0

3
-10
4
-20
12 24 12 24 12 24 12
Time of day (hr)

Fig. 3. Per1-luc rhythms in anterior and posterior SCN slices from mice exposed to long photoperiod. Data show Per1-luc rhythms
of cultured SCN slices from two representative mice (a and b) exposed to LD 18:6. Left panels indicate Per1-luc records of two
consecutive SCN slices, which were detrended by 12 h-moving average from the original data. Solid and broken lines indicate
Per1-luc records of the posterior and anterior slices from the same mouse, respectively. Horizontal dark and white bars for the
first day of culture indicate dark and light phases to which mice were exposed. Right panels indicate peak phases of Per1-luc
rhythms shown in the left panels for 5-day culture. (Reproduced from Fig. 3 of Honma et al., 2008, with permission.)

Per1-luc peak phase-locked with the onset of knock-out mice, Daan et al. (2001) formulated
the behavioral activity rhythm and the posterior a hypothesis that proposed E and M oscillators
peak with the offset of behavioral rhythms (Fig. 3). consisting of two molecular clocks; the E oscillator
Neither anterior nor posterior Per1-luc peaks tracks evening lights and is composed of Per2 and
phase-locked to light on or light off. The findings Cry2, and the M oscillator tracks morning light
suggest that the discrete regions of the SCN were and is composed of Per1 and Cry1. Our data sug-
separately involved in the regulation of the phase gest that Per1 is involved in both E and M clocks,
of behavioral rhythms: the anterior SCN being but that only the region in the SCN is different.
involved in the regulation of the activity onset (thus The role of the third oscillator exhibiting an
the site of E oscillator) and the posterior SCN, in early morning peak in the anterior SCN is not
that of activity offset (M oscillator). Previously, known. Per1-luc rhythms in horizontal SCN slices,
on the basis of the phase responses of various gene by covering entire rostro-caudal parts in a single
135

SCN slice, demonstrated phase difference of cel- in specific regions of the SCN separately entrain to
lular rhythms (Fig. 4). Time of Per1-luc expres- light on or off and regulate the offset and onset of
sion corresponded roughly to the subjective light behavioral rhythms separately. On the other hand,
phase of an LD cycle to which mice were in the SCN slices from mice exposed to short days,
exposed. Cells in the caudal SCN peaked earlier the Per1 peaks are distributed over only a few
than the rostral ones. In our experiments, we only hours. Irrespective of the photoperiods, the phases
detected Per1-luc peaks in LD 18:6 very early in of circadian Per1-luc rhythms in the anterior SCN
the morning at around light on in the lateral part were distributed over a wider range than those in
of the anterior-to-middle SCN (indicated by an the posterior SCN, indicating that not only phase
arrow in Fig. 4). Newly developed spatiotemporal position but also phase distribution is different
maps based on Fast Fourier Transform (FFT) depending on the region in the SCN.
analysis (Fricker et al., 2007) clearly demonstra-
ted the localization of these different cell groups
(Fig. 5, see the online colored version). Since Coupling of oscillator cells and that of
the late Per1 peak in the anterior SCN appeared regional pacemakers
a few hours before the onset of subjective night
on the first day of culture, in total, the peak Although it still remains to be studied as to how
phases were distributed over about 15 h in SCN cells communicate their rhythmic informa-
the subjective day. The SCN cells dynamically tion to express coherent circadian rhythms,
change their circadian phases depending on the several candidates have been found. Among
photoperiod. These results suggest that cell groups them, VIP seems to be an important coupling

LD6:18
2:00 4:00 6:00 8:00 10:00 12:00 14:00 16:00 18:00 20:00 22:00
C

R
LD12:12
C

LD15:9
C

LD18:6
C

Fig. 4. Per1-luc bioluminescence images in horizontal SCN slice. Horizontal SCN slices were cultured from mice exposed to four
different photoperiods, LD 6:18 (L 9:00–15:00), LD 12:12 (L 6:00–18:00), LD 15:9 (L 4:30–19:30), and LD 18:6 (3:00–21:00).
Bioluminescence images were captured by a CCD camera (ORCA-II, Hamamatsu Photonics), cooled at 60  C, by exposing
cultured SCN for 59 min. Exposure was started at the local time indicated at the top of each image. Shown are the time series
images from 2:00 to 22:00 on the second day of culturing. Scale bars indicate 80 mm. Arrows in images at 2:00, 6:00 and 16:00 of
LD18:6 indicate regions where Per1-luc differentially peaked.
136

Amplitude high Phase

caudal
(a) (b)
VIII

rostral

OC

low

0 6 12 18 0
Local time (h)

Fig. 5. Amplitude and phase maps of Per1-luc expression rhythms in a horizontal SCN slice. Time series of bioluminescence images
were analyzed at pixel level by Fast Fourier Transform (FFT). Resultant amplitude maps (a) and phase maps (b) are demonstrated
in psuedocolor. Rhythm amplitude of Per1-luc was high in the posterior SCN. The relative scale of amplitude is shown at the right
side of the amplitude map. The phase map shows differentially phasing cell groups in specific regions in the SCN on the second day
of culturing. Arrows in the SCN from LD 18:6 indicate localization of three cell groups. Horizontal dark and yellow bar indicates
subjective night and day. Horizontal pseudocolor bar shows peak phase of Per1-luc rhythm in each pixel on the second day of
culture. OC, optic chiasm; VIII, the third ventricle; scale bar measures 50 mm.

cue within the SCN. VIP deficiency and VIP of synaptic interaction by TTX may have reduced
receptor, VPAC2, knockout cause severe dis- the number of synapses.
turbances in the behavioral rhythms, which TTX treatment also affected the synchroniza-
are basically due to the desynchronization of tion among neurons of cultured SCN slices. Webb
SCN cellular oscillators (Maywood et al., 2006). et al. (2009) treated the SCN slice with TTX
Furthermore, Maywood et al. (2011) recently twice, with each treatment continuing for 6 days.
demonstrated, by co-culturing SCN slices from About 13% of cells remained rhythmic during
mice with clock gene mutations with a wild-type both treatments, suggesting that they are cell-
SCN slice, that paracrine signals are able to autonomous clock cells. But these cells were not
restore circadian PER2::LUC rhythms. Diffusible localized in specified regions of the SCN. In addi-
signals are important not only in synchronization tion, they also observed switching of rhythmic
but also in sustaining the cellular oscillation. neurons during the repeated TTX challenges.
Synaptic interactions are also involved in rhythm Some rhythm-positive cells became rhythm nega-
synchronization. We have studied the role of synap- tive in the second trial, and vice versa. Thus, they
tic interaction in rhythm coupling by measuring the concluded that circadian rhythm generation is not
spontaneous firing of single SCN neurons before localized to a specific cell groups, but that instead
and after treatment with tetrodotoxin (TTX), a each SCN cell has about 25% of chance to exhibit
sodium channel blocker. Dispersed SCN neurons a circadian rhythm. In their experiment, however,
showed synchronized circadian rhythms when they two SCN slices were treated with TTX for 6 days.
had functional synapses (Shirakawa et al., 2000), The duration of treatment might be long enough
but the inter-spike intervals of the synchronized to exert an irreversible change in synaptic trans-
neuron pairs were lengthened after TTX treatment mission. We suppressed synaptic transmission with
(Honma et al., 2000), suggesting a reduced efficacy TTX for 48 h in the wild-type rat SCN slices
of the synaptic transmission. Prolonged suppression cultured on the MED probe with the method
137

described elsewhere (Nakamura et al., 2001, 2002). the phase predicted by the rhythm before the treat-
Following the baseline recording of spontaneous ment as shown in Fig. 6, as was observed in those of
firing for more than 5 days, SCN slices were treated dispersed SCN cell cultures (Honma et al., 2000;
with TTX (final concentration 200 nM) for 48 h. Welsh et al., 1995). We analyzed circadian period
The medium was changed daily. After the treat- by a chi-square periodogram using a record of five
ment, they were washed with the recording consecutive days at the significance levels of
medium twice and spontaneous discharges were p  0.01. In each slice, differences between the circa-
monitored for an additional 1–4 weeks. TTX treat- dian periods before and after the TTX treatment
ment completely abolished the spontaneous firing were statistically analyzed using two-way ANOVA,
of the SCN neurons. After the washout of TTX, and difference in the period variance within a slice
most neuronal rhythms started to free-run from was analyzed by F-test. The circadian periodicity in

(a) ch12 ch14 ch15 ch18


1
25.1 25.3 25.4 25.1

10 24.0 27.5
24.8 25.1
(days)

20
24.0 25.6 23.9 24.0

0 12 0 12 0 0 12 0 12 0 0 12 0 12 0 0 12 0 12 0
(b) ch1 ch6 ch8
1 ch5
25.0 24.7 25.0 25.2

10 26.1. 24.0 25.7 24.0


(days)

20
24.0 23.9
23.9 23.7

0 12 0 12 0 0 12 0 12 0 0 12 0 12 0 0 12 0 12 0
Time of day (h) Time of day (h) Time of day (h) Time of day (h)

Fig. 6. Effects of TTX treatment on the synchronization of spontaneous firing rhythms. SCN slices of newborn rats were cultured on
an MED probe, and the numbers of spontaneous discharges with signal-to-noise ratio >2 were collected every 20 s from eight
electrodes covered by an SCN slice. Four firing rhythms simultaneously monitored from a single SCN slice are double-plotted.
Firing rhythms from two SCN slices (a and b) are plotted in histogram of mean firing rates against the time of day at 15-min
bins. The number at the upper left margin of each panel indicates channel number of the MED probe. Shadowed areas are the
time of TTX treatment (200 nM, 48 h). Numbers and arrows at the right margin of each panel are the circadian periods of firing
rhythm at the indicated time analyzed by a chi-square periodogram.
138

each channel was designated as altered by TTX channels. In five control SCN slices, the circadian
treatment when it showed a larger change than that periods were not significantly changed, but the vari-
observed in vehicle-treated controls (> the mean ation was changed in two slices: increasing in one
period difference þ 2 SD ¼ 1.06 h). Before the treat- and decreasing in the other. Since circadian periods
ment, neuronal rhythms were basically synchroni- were either desynchronized or shortened by TTX
zed within a slice. Out of 15 SCN slices examined, treatment in most SCNs examined, it is apparent
only two SCNs showed no significant difference in that the suppression of synaptic transmission for
their circadian periods after the TTX treatment only 48 h modified coupling of cellular oscillators.
(Fig. 7a). On the other hand, TTX treatment In addition, neuronal rhythms were abolished or
desynchronized the circadian neuronal rhythms in appeared in some channels, suggesting a loss or
seven SCNs and significantly shortened the periods acquisition of the interaction between neurons with
in six SCNs (Fig. 7c). In the remaining SCN slice, an autonomous oscillator. The present data suggest
rhythms either appeared or disappeared, and we that the ensemble of cellular rhythms is critical for
could not compare the difference. Interestingly, the expression of synchronized firing rhythms, which
desynchronized rhythms were partially or fully is, at least partly, accomplished through Naþ-chan-
resynchronized in 1–2 weeks. In this experiment, nel dependent mechanisms.
we recorded five to eight neuronal rhythms from a TTX-resistant circadian rhythms have also been
single SCN slice. Out of 95 channels recorded, circa- reported. Using the Ca2 þ-sensitive fluorescent pro-
dian periods were shortened in 36 and lengthened in tein Cameleon, Ikeda et al. (2003) reported that
10. Period change was smaller than 1.06 h (the mean TTX blocked multi-unit activity (MUA) but not
change þ 2 SD of controls) in 40 channels, was intracellular Ca2 þ rhythms in the SCN cells. On
abolished in five channels and appeared in four the other hand, TTX application was reported to

(a) Unchanged (b) Desynchronized (c) Shortened

28 24.0±0.1 24.1±0.2 24.7±0.3 23.8±0.3

26

24
Circadian period (h)

25.4±0.3 25.0±1.3
22

28 24.5±0.2 24.5±0.2 23.7±0.1 23.5±1.0 24.9±0.8 23.8±0.3

26

24

22
Pre Post Pre Post Pre Post

Fig. 7. TTX effects on the synchronous circadian firing rhythm of rat SCN. Effects of TTX treatment (48 h) on the circadian
periodicity were examined in each SCN slice on a MED probe. The presence of significant circadian periodicity and its period
were evaluated by a chi-square periodogram using 5-day records immediately before (pre) and after (post) the treatment.
Results of representative SCN slices in which circadian periodicity of neuronal rhythms was unchanged (a), desynchronized (b),
and shortened (c) are shown. Numbers in each panel are mean circadian periods (SD) of individual firing rhythms.
139

dramatically suppress the amplitude and to as paracrine signals (Maywood et al., 2006, 2011).
desynchronize the Per1-luc rhythms in SCN Furthermore, extreme circadian periodicities with
neurons (Yamaguchi et al., 2003), and it also a large period distribution may have affected the
suppressed AVP expression rhythms in the SCN synchronization among cellular rhythms. Clock-
(Arima et al., 2002). mutant SCN cells exhibit much longer periods with
larger variability than wild-type cells. Various clock
gene-deficient mice also exhibit either shorter or
Clock gene functions in the SCN longer periodicity compared with wild-type mice.
cellular networks Even in wild-type rats, the synchronization within
an SCN slice becomes weaker as the periodicity of
A lack or mutation of clock genes results in a individual neuronal rhythms deviates more from
genotype-specific rhythm phenotype. Clock is 24 h (Honma et al., 2004). Pittendrigh and Daan
the first clock gene cloned in mammals. Upon (1976a) demonstrated in rodents that also the circa-
exposure to DD, the D19 clock-mutant mice dian period of behavioral rhythms becomes more
exhibited circadian rhythms with extremely long variable as the species average deviates more from
periods, which eventually became aperiodic 24 h, which thus seems to be a general rule for circa-
(Vitaterna et al., 1994). SCN neurons of dian rhythms.
organotypic slice cultures from clock-mutant mice In conclusion, in the SCN, multiple oscillator
exhibit robust and synchronized circadian firing neurons with relatively diverse circadian periodic-
rhythms with a period comparable to that of the ity are mutually synchronized to form a few
behavioral rhythms; however, the percentage of regional pacemakers, which further couple to
neurons with significant circadian rhythms was express coherent rhythm outputs. The multi-
much less than that of wild-type mice (Nakamura oscillator neuronal pacemaker structure is advan-
et al., 2002). The Clock mutation similarly tageous for adapting cyclic environments of wide
affected the circadian periodicity of the SCN variability without losing stable periodicity and
neurons in the slice and dispersed cell cultures how- phase adjustment. Desynchronization within and
ever, the mutation dramatically reduced the per- between the regional pacemakers results in an
centage of rhythm-positive neurons in the arrhythmic phenotype even if each cellular clock
dispersed cell cultures. Genotype difference (clock can generate a circadian rhythm. Not all of the
mutant vs. wild type) and culture system differences SCN cells seem to possess a cell-autonomous clock;
(organotypic slices vs. dispersed cells) suggest that thus, mammals have developed a central circadian
not all of the SCN neurons possess a cell-autono- clock in the SCN via oscillatory coupling with neu-
mous circadian clock, as described above, and that ronal and paracrine signals.
the clock mutation affects rhythm synchronization,
especially that between neurons with and without
a cell-autonomous clock. The mutation may directly
References
reduce rhythm synchronization by reducing either
coupling signals or receptors. Our results suggest Abraham, U., Granada, A. E., Westermark, P. O., Heine, M.,
that SCN tissue structures compensated the effects Kramer, A., & Herzel, H. (2010). Coupling governs entrain-
of the clock mutation on single neurons (Nakamura ment range of circadian clocks. Molecular Systems Biology,
et al., 2002). Such compensation was also observed 6, 438.
Abrahamson, E. E., & Moore, R. Y. (2001). Suprachiasmatic
in various other clock gene knockouts (Liu et al.,
nucleus in the mouse: Retinal innervation, intrinsic organiza-
2007). Therefore, an SCN tissue structure with close tion and efferent projections. Brain Research, 916, 172–191.
cell-to-cell appositions may be essential to transmit Arima, H., House, S. B., Gainer, H., & Aguilera, G. (2002).
the rhythm information for synchronization, such Neuronal activity is required for the circadian rhythm of
140

vasopressin gene transcription in the suprachiasmatic robustness against mutations in the SCN circadian clock net-
nucleus in vitro. Endocrinology, 143, 4165–4171. work. Cell, 129, 605–616.
Daan, S., Albrecht, U., van der Horst, G. T., Illnerová, H., Maywood, E. S., Chesham, J. E., O’Brien, J. A., &
Roenneberg, T., Wehr, T. A., et al. (2001). Assembling a Hastings, M. H. (2011). A diversity of paracrine signals
clock for all seasons: Are there M and E oscillators in the sustains molecular circadian cycling in suprachiasmatic
genes? Journal of Biological Rhythms, 16, 105–116. nucleus circuits. Proceedings of the National Academy of
Fricker, M. D., Tlalka, M., Bebber, D., Takagi, S., Sciences of the United States of America, 108, 14306–14311.
Watkinson, S. C., & Darrah, P. R. (2007). Fourier-based Maywood, E. S., Reddy, A. B., Wong, G. K., O’Neill, J. S.,
spatial mapping of oscillatory phenomena in fungi. Fungal O’Brien, J. A., McMahon, D. G., et al. (2006). Synchroniza-
Genetics and Biology, 44, 1077–1084. tion and maintenance of timekeeping in suprachiasmatic
Herzog, E. D., Aton, S. J., Numano, R., Sakaki, Y., & Tei, H. circadian clock cells by neuropeptidergic signaling. Current
(2004). Temporal precision in the mammalian circadian sys- Biology, 16, 599–605.
tem: A reliable clock from less reliable neurons. Journal of Nagano, M., Adachi, A., Nakahama, K., Nakamura, T.,
Biological Rhythms, 19, 35–46. Tamada, M., Meyer-Bernstein, E., et al. (2003). An abrupt shift
Herzog, E. D., Geusz, M. E., Khals, S. B., Straume, M., & in the day/night cycle causes desynchrony in the mammalian
Block, G. D. (1997). Circadian rhythms in mouse circadian center. The Journal of Neuroscience, 23, 6141–6151.
suprachiasmatic nucleus explants on multimicroelectrode Nakamura, W., Honma, S., Shirakawa, T., & Honma, K.
plates. Brain Research, 757, 285–290. (2002). Clock mutation lengthens the circadian period with-
Honma, S., Nakamura, W., Shirakawa, T., & Honma, K. (2004). out damping rhythms in individual SCN neurons. Nature
Diversity in the circadian periods of single neurons of the rat Neuroscience, 5, 399–400.
suprachiasmatic nucleus depends on nuclear structure and Nakamura, W., Honma, S., Shirakawa, T., Oguchi, H., &
intrinsic period. Neuroscience Letters, 358, 173–176. Honma, K. (2001). Regional pacemakers composed of mul-
Honma, S., Shirakawa, T., Katsuno, Y., Namihira, M., & tiple oscillator neurons in the rat suprachiasmatic nucleus.
Honma, K. (1998). Circadian periods of single suprachiasmatic European Journal of Neuroscience, 14, 666–674.
neurons in rats. Neuroscience Letters, 250, 157–160. Pittendrigh, C. S., & Daan, S. (1976a). A functional analysis of
Honma, S., Shirakawa, S., Namakura, W., & Honma, K. circadian pacemakers in nocturnal rodents. I. The stability
(2000). Synaptic communication of cellular oscillations in and lability of spontaneous frequency. Journal of Compara-
the rat suprachiasmatic neurons. Neuroscience Letters, 294, tive Physiology. A, 106, 223–252.
113–116. Pittendrigh, C. S., & Daan, S. (1976b). A functional analysis
Honma, S., Inagaki, N., Ono, D., Yoshikawa, T., of circadian pacemakers in nocturnal rodents. V. Pacemaker
Hashimoto, S., & Honma, K. (2008). Clock mechanisms structure: A clock for all seasons. Journal of Comparative
for seasonal adaptation: Morning and evening oscillators in Physiology. A, 106, 333–355.
the suprachiasmatic nucleus. Sleep Biol. Rhythms, 6, 84–90. Reppert, S. M., & Weaver, D. R. (2002). Coordination of cir-
Ikeda, M., Sugiyama, T., Wallace, C. S., Gompf, H. S., cadian timing in mammals. Nature, 418, 935–941.
Yoshioka, T., Miyawaki, A., et al. (2003). Circadian dynam- Shinohara, K., Honma, S., Katsuno, Y., Abe, H., & Honma, K.
ics of cytosolic and nuclear Ca2 þ in single suprachiasmatic (1995). Two distinct oscillators in the rat suprachiasmatic
nucleus neurons. Neuron, 38, 253–263. nucleus in vitro. Proceedings of the National Academy of
Inagaki, N., Honma, S., Ono, D., Tanahashi, Y., & Honma, K. Sciences of the United States of America, 92, 7396–7400.
(2007). Separate oscillating cell groups in mouse Shirakawa, S., Honma, S., Katsuno, Y., Oguchi, H., &
suprachiasmatic nucleus couple photoperiodically to the Honma, K. (2000). Synchronization of circadian firing
onset and end of daily activity. Proceedings of the National rhythms in cultured rat suprachiasmatic neurons. European
Academy of Sciences of the United States of America, 104, Journal of Neuroscience, 12, 2833–2838.
7664–7669. van den Pol, A. (1980). The hypothalamic suprachiasmatic
Ko, H. C., Yamada, Y. R., Welsh, D. K., Buhr, E. D., nucleus of the rat: Intrinsic anatomy. The Journal of Com-
Liu, A. C., Zhang, E. E., et al. (2010). Emergence of parative Neurology, 191, 661–702.
noise-induced oscillations in the central circadian pace- Vitaterna, M. H., King, D. P., Chang, A. M.,
maker. PLoS Biology, 8, e1000513. Kornhauser, J. M., Lowrey, P. L., McDonald, J. D., et al.
Kuhlman, S. J., Quintero, J. E., & McMahon, D. G. (2000). GFP (1994). Mutagenesis and mapping of a mouse gene, Clock,
fluorescence reports Period 1 circadian gene regulation in essential for circadian behavior. Science, 264, 719–725.
the mammalian biological clock. Neuroreport, 11, 1479–1482. Webb, A. B., Angelo, N., Huettner, J. E., & Herzog, E. D.
Liu, A. C., Welsh, D. K., Ko, C. H., Tran, H. G., Zhang, E. E., (2009). Intrinsic, nondeterministic circadian rhythm genera-
Priest, A. A., et al. (2007). Intercellular coupling confers tion in identified mammalian neurons. Proceedings of the
141

National Academy of Sciences of the United States of Yamazaki, S., Numano, R., Abe, M., Hida, A., Takahashi, R.,
America, 106, 16493–16498. Ueda, M., et al. (2000). Resetting central and peripheral cir-
Welsh, D. K., Logothetis, D. E., Meister, M., & Reppert, S. M. cadian oscillators in transgenic rats. Science, 288, 682–685.
(1995). Individual neurons dissociated from rat supra- Yoo, S. H., Yamazaki, S., Lowrey, P. L., Shimomura, K.,
chiasmatic nucleus express independently phased circadian Ko, C. H., Buhr, E. D., et al. (2005). PERIOD2::LUCIFER-
firing rhythms. Neuron, 14, 697–706. ASE real-time reporting of circadian dynamics reveals per-
Yamaguchi, S., Isejima, H., Matsuo, T., Okura, R., Yagita, K., sistent circadian oscillations in mouse peripheral tissues.
Kobayashi, M., et al. (2003). Synchronization of cellular clocks Proceedings of the National Academy of Sciences of the
in the suprachiasmatic nucleus. Science, 302, 1408–1412. United States of America, 101, 5339–5346.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 9

Dynamic neuronal network organization of the


circadian clock and possible deterioration in disease

Johanna H. Meijer{,*, Christopher S. Colwell{,{, Jos H. T. Rohling{, Thijs Houben{ and


Stephan Michel{

{
Laboratory for Neurophysiology, Department of Molecular Cell Biology, Leiden University Medical Center,
Leiden, The Netherlands
{
Laboratory of Circadian and Sleep Medicine, Department of Psychiatry and Biobehavioral Sciences, David Geffen
School of Medicine, University of California, Los Angeles, CA, USA

Abstract: In mammals, the suprachiasmatic nuclei (SCNs) function as a circadian pacemaker that drives
24-h rhythms in physiology and behavior. The SCN is a multicellular clock in which the constituent
oscillators show dynamics in their functional organization and phase coherence. Evidence has emerged
that plasticity in phase synchrony among SCN neurons determines (i) the amplitude of the rhythm, (ii)
the response to continuous light, (iii) the capacity to respond to seasonal changes, and (iv) the phase-
resetting capacity. A decrease in circadian amplitude and phase-resetting capacity is characteristic
during aging and can be a result of disease processes. Whether the decrease in amplitude is caused by
a loss of synchronization or by a loss of single-cell rhythmicity remains to be determined and is
important for the development of strategies to ameliorate circadian disorders.

Keywords: circadian; SCN; network; plasticity; synchronization; entrainment; photoperiod; jet lag; aging;
circadian disorders.

Introduction environment, brought about by the rotation of the


earth around its axis. In order to anticipate these
Most animals show clear 24-h rhythms in physiol- changes, innate clocks have evolved that allow
ogy and behavior. These rhythms have developed organisms to prepare for the predictable onset of
as an adaptation to the recurring changes in the night and day. In mammals, the central clock is
located in the suprachiasmatic nucleus (SCN).
*Corresponding author. The SCN is a bilateral structure, located at the base
Tel.: þ31 71 526 9760; Fax: þ31 71 526 8270 of the brain, with the ventral aspect immediately
E-mail: j.h.meijer@lumc.nl above the optic chiasm. The SCN contains about

http://dx.doi.org/10.1016/B978-0-444-59427-3.00009-5 143
144

10,000 neurons on each side and maintains direct we raise the possibility that one mechanism by
and indirect connections with many parts of the which aging and disease can alter the function of
central nervous system (Kalsbeek et al., 2006; the circadian system is by reducing the strength
Morin and Allen, 2006). Light is the major external of the intracellular coupling within the SCN circuit.
stimulus that synchronizes the endogenous clock
to the external 24-h cycle. It reaches the neurons
of the SCN via a monosynaptic pathway, formed SCN waveform is an ensemble property
by melanopsin containing retinal ganglion cells
(Berson et al., 2002; Hattar et al., 2003), that proj- Recordings of electrical impulse frequency have
ect with glutamate- and PACAP-containing fibers been performed both in vivo in the intact animal
to the SCN (Golombek and Rosenstein, 2010; and in vitro in brain slices that contain the SCN.
Hannibal et al., 2000). A distinction has been made Pioneers of SCN in vivo recordings were Inouye
in the core and shell region of the SCN (Antle and and Kawamura (1979) who performed these
Silver, 2005; Gamble et al., 2007; Ibata et al., 1989; recordings to show that the SCN functions as an
Kiss et al., 2008; Moore and Silver, 1998). The core endogenous oscillator. SCN in vitro recordings
contains gastrin-releasing peptide (GRP)- and confirmed that the SCN functions as an endoge-
vasoactive intestinal polypeptide (VIP)-expressing nous oscillator and does not require rhythmic
neurons and is retinorecipient (Abrahamson and input to sustain rhythmicity (Green and Gillette,
Moore, 2001; Antle et al., 2005). The shell contains 1982; Groos and Hendriks, 1982; Shibata et al.,
vasopressin and receives input from the core of 1982). A consistent finding of in vitro and in vivo
the SCN (Moore et al., 2002), while the core is electrophysiological recordings is that during the
only sparsely innervated by the shell (Romijn day, electrical activity is high, and during the night
et al., 1997). it is low, both in nocturnal and in diurnal species.
Generation of circadian rhythmicity occurs at In nocturnal animals, high activity of the SCN
the single-cell level and is based on an intertwined corresponds with the resting phase and low activity
negative feedback loop between clock genes and with the active phase of the animal (Brown and
their protein products (Herzog, 2007; Reppert Piggins, 2007; Colwell, 2011; Gillette, 1996).
and Weaver, 2002). The genetic basis for rhythm In diurnal species, this is reversed and electrical
generation can explain that isolated cells of the activity is in phase with the behavioral activity pat-
SCN are capable of generating circadian rhythms tern (Challet, 2007). A major difference between
and do not require rhythmic input (Webb et al., nocturnal and diurnal animals is therefore the
2009; Welsh et al., 1995, 2010). The important phase relationship between the SCN and other
implication is that the SCN functions as a multi- parts of the central nervous system.
oscillator structure in which the different neurons The waveform of the SCN electrical activity
are mutually synchronized in phase in order to rhythm is almost sinusoidal. Simultaneous
function as a coherent pacemaker. In this chapter, recordings of electrical activity and behavioral
we review the role of synchronization within the activity have allowed us to determine at what
SCN for the functional adaptation of the SCN as level of electrical activity, behavior is initiated or
a clock and cover specific questions: How is the arrested (Houben et al., 2009). The onset of activ-
multioscillator structure used for the adaptation ity appears to occur at around the 50% level of
to shifts of the light–dark cycle? What is the the rhythm amplitude. Within a certain range
response of the SCN to constant light and to sea- around the 50% level, the chance for a transition
sonal changes? What are the consequences of from rest to activity is maximal. Also for the off-
changes in neuronal synchrony for rhythm wave- set of activity, a close correlation with the
form, amplitude, and resetting capacity? Finally, 50% level is observed. It is concluded that the
145

on- and offset regulation of behavioral activity (a)


is rather predictable from the level of SCN elec- 800

MUA (Hz)
trical activity and suggests a direct relationship 700
600
between SCN neural activity and the onset–offset
500
of behavioral patterns. These findings have
400
implications for situations in which the waveform
of the SCN changes, which will be discussed in (b)
this chapter.
The waveform of the SCN molecular and elec-
trical activity reflects the combined activity of
many neurons and is thus a composite tissue- 0 12 24 36 48 60
level property (Evans et al., 2011; Maywood Time (h)
et al., 2006; Quintero et al., 2003; Schaap et al.,
2003; Yamaguchi et al., 2003; Yamazaki et al., Fig. 1. Distribution of single neuron electrical activity pattern
explains the multiunit electrical activity pattern of the SCN.
2000). Decomposition of the ensemble pattern (a) An in vivo electrical activity recording of 60 h, with the
has revealed that small subpopulations and single gray background indicating night and the white background
SCN neurons exhibit much shorter periods of indicating day. (b) One typical recording of an electrical
enhanced electrical activity. Single units in the activity pattern of a mouse housed in a 12:12 light–dark cycle
mouse and rat SCN appear to be active for was used and plotted repetitively according to a Gaussian
distribution, to simulate the phase distribution of neuronal
durations of about 5 h (Brown and Piggins, 2009; activity patterns in the SCN. The electrical activity pattern of
Brown et al., 2006; Schaap et al., 2003; the population renders the characteristic multiunit electrical
VanderLeest et al., 2007). The time of maximal activity pattern.
activity of most neurons of the SCN is during the
day, but some neurons are active during the night single-cell oscillations. For example, recordings
(Fig. 1). The distribution of individual neuronal from individual SCN cells in vitro of rhythms
activity patterns can account for the ensemble pat- in Period1 promoter-driven GFP fluorescence
tern of the SCN, as the highest density of active rhythms have revealed that individual neurons
neurons occurs during midday, while the lowest from arrhythmic mice kept in constant light are
density of active neurons occurs at midnight. This not compromised in their rhythm-generating abil-
phase distribution renders a sinusoidal pattern in ity (Ohta et al., 2005). In fact, the individual
SCN multiunit activity at the ensemble level. neurons remained rhythmic, but the population
shows a severely distorted phase distribution.
The strong disruption of phase coherence by
Constant light effects exposure to constant light leads to a decline in cir-
cadian amplitude at the tissue level. Similarly,
Exposing organisms to continuous light causes in vivo recordings in the SCN of mice kept in
robust changes in the circadian rhythms including continuous light result in reduced overt rhythmic-
a strong reduction in the amplitude of overt ity and confirm these findings for electrical activ-
rhythms, lengthening of the endogenous cycle ity of the SCN. These recordings revealed a
length (period) (Aschoff, 1979), and can even gradual decline in amplitude and a distortion of
lead to splitting of circadian locomotor behavior electrical rhythmicity upon prolonged exposure
and arrhythmicity (Depres-Brummer et al., 1995; to continuous light (Fig. 2). The results show the
Eastman and Rechtschaffen, 1983; Pittendrigh necessity of coupling mechanisms, in addition to
and Daan, 1976). Previous data suggest that the the presence of rhythm-generating units, and indi-
impact of constant light is to desynchronize cate the vital importance of the network-level
146

(a)

700
MUA (Hz)

600

500

400

300
0 48 96 144 192 240 288 336 384 432 480 528 576 624 672 720 768 816
Time (h)
(b)
100
Rhythm amplitude (Normalized)

75

50

25

0
DD 1 2 3 4 5 6 7
Circadian cycle

Fig. 2. Effect of constant light on SCN neuronal activity and behavior. (a) Long-term in vivo recording of SCN neuronal activity in a
freely moving mouse. SCN activity was recorded with implanted microelectrodes and is depicted in 10-min bins (black trace). Gray
background indicates lights off and white background lights on. During the first days in constant light, a decrease in amplitude is
observed. During prolonged exposure to constant light, the amplitude of the rhythm was variable as the SCN appeared to lose
and then regain coherence. This may reflect temporal synchronization and desynchronization of the constituent oscillators.
(b) Effect of constant light on the mean amplitude of the SCN rhythm in seven animals. The animals were exposed to LL for at
least 7 days following exposure to constant darkness. The rhythm was normalized for each animal and expressed relative to the
amplitude in constant darkness (gray bar, SEM). White bars depict the amplitude during the first 7 days in constant light.

organization of the SCN to obtain robust and high- factors determine the phase coherence within the
amplitude rhythms. SCN and hence influence the waveform of the
Synchronization among the neurons of the SCN ensemble rhythm. Coupled oscillator models have
is critically determined by a variety of coupling found that large groups of oscillators such as
mechanisms (for review, see Welsh et al., 2010). SCN neurons can organize themselves into many
Several coupling factors have been implicated such different configurations to meet environmental
as VIP, GABA, and even gap junctions. These challenges. These conformational changes are
147

rapidly driven by alterations in the strength of peak) in short days and decompressed (broad
neuronal coupling mechanisms (Golubitsky peak) in long days (Houben et al., 2009; Mrugala
et al., 1999; Strogatz, 2003; Strogatz and Stewart, et al., 2000). The same modulation at the tissue—
1993). The flexibility in coupling strength between or population—level is observable in a number
neuronal populations enables the system to cope of clock genes (Carr et al., 2003; Johnston
with different environmental conditions such as et al., 2003, 2005; Nuesslein-Hildesheim et al.,
changes in the light–dark cycle and responses to 2000; Sumova et al., 2002, 2003; Tournier et al.,
constant light. In the next section, we will consider 2003), Fos expression (Jac et al., 2000b; Sumova
how changes in day length affect the phase syn- et al., 1995; Vuillez et al., 1996), and vasopressin
chrony among SCN neurons. (Jac et al., 2000a) and prokineticin 2 (PK2)
mRNA levels (Cheng et al., 2005). The peak width
appears to serve as an internal representation of
Seasonal changes in waveform by changes in the length of the day. Importantly, the photope-
phase distribution riodic driven changes in the SCN waveform and
activity patterns are maintained for weeks even
In response to the changes in day length that when the animal is released back into constant
occur over the course of the year, many animals conditions. We view these changes as a type of
undergo strong alterations in their anatomy, plasticity in SCN physiology.
physiology, and behavior. Seasonal breeding While photoperiodic encoding of day length is
organisms show drastic changes in behavioral explainable by the change in the width of the elec-
activity, metabolism, and reproductive physiology trical activity peak, a valid question is whether also
(Ebling and Barrett, 2008; Hazlerigg and Loudon, changes in triggering level account for the modula-
2008). The duration of the melatonin signal is a tion in activity duration. Presumably, a higher trig-
critical cue for some seasonal breeders, while gering level would result in an increase in the
other seasonal adaptations do not appear to be activity period and a lowering of the triggering
driven by melatonin. To provide an example, level in an increase in the activity duration. This
C57 mice have very low melatonin and yet display mechanism was proposed by Wever (1960) as a
robust photoperiodic driven changes in behavioral possible underlying mechanism for modulation of
activity. When these mice are exposed to long or long- and short-activity periods in finches. In order
short photoperiods, the duration of activity to explore this possibility, we have recorded the
(alpha) is systematically lengthened in short days SCN in vivo under long and short photoperiods
and shortened in long days. The changes in alpha and have simultaneously scored the onset and off-
are maintained even when the mice are placed in set of behavioral activity (Houben et al., 2009).
constant darkness. The behavioral activity pattern However, we found no evidence for changes in
that is then displayed reflects the previous photo- triggering level, and instead, we observed that
period, indicating an endogenous memory, or both in long and in short days, transitions between
encoding mechanism, which is present for days activity to rest were most likely to occur at 50% of
to weeks (Sumova et al., 2004). the electrical activity rhythm. We conclude that
The SCN plays an important role in the seasonal the change in peak width represents the major
encoding process. The sinusoidal-like waveform of mechanism by which photoperiod is internally
the SCN electrical activity profile is strongly encoded by the SCN, and that no changes in
altered under influence of long and short days, as triggering level are involved.
shown by in vivo recordings of the SCN, or by in The plasticity in the waveform of the SCN,
situ analysis of clock gene expression. The electri- observable in behavioral activity patterns, and in
cal activity profile becomes compressed (narrow the SCN in vivo is preserved when the SCN is
148

recorded in vitro, in a brain slice preparation (a)


(Mrugala et al., 2000). This finding opens up
the important possibility to investigate the mecha-
nism underlying SCN waveform changes in vitro. 8
Two studies have independently proposed that

Activity (Hz)
SCN neurons show oscillations in their molecular 6
or electrical activity pattern that are out of phase
and that these phase differences could contribute 4
to the photoperiodically induced changes in
2
waveform (Quintero et al., 2003; Schaap et al.,
2003). These predictions were confirmed in 0
recordings of subpopulations of individual SCN 0 6 12 18
neurons from mice held in different photoperiods ZT (h)
(Brown and Piggins, 2009; Hazlerigg et al., 2005; (b)
Inagaki et al., 2007; Naito et al., 2008; Sosniyenko
et al., 2009; VanderLeest et al., 2007). These stud-
ies all found evidence that exposing mice to long 8
days causes a desynchrony in the electrical activ-
Activity (Hz)
ity and gene expression rhythms within the SCN 6
(Fig. 3).
Surprisingly, the neural activity patterns 4
recorded from individual SCN neurons show little
2
change in their activity pattern in long and short
days (Brown and Piggins, 2009; Naito et al., 2008; 0
VanderLeest et al., 2007). Photoperiod changes 18 0 6 12
primarily the phase distribution but not the dura- ZT (h)
tion of electrical activity within individual cells.
While Brown and Piggins (2009) observed longer Fig. 3. The distribution of single-unit electrical activity patterns is
duration activity of cells in the dorsal SCN under responsible for the difference in multiunit electrical activity
pattern in long and short photoperiod. The multiunit electrical
long days, they also showed that these changes activity pattern in short (a) and long (b) photoperiod is shown
render very little difference in the duration of as a thick black line. This pattern is derived from an ensemble
the population activity pattern in agreement with of single-unit activity patterns that are distributed over the 24-h
simulations and predictions from other studies cycle according to a Gaussian distribution. The single-unit
(Rohling et al., 2006a,b).There is consensus on activity patterns for short and long photoperiod were determined
from a number of recorded single-unit activity patterns (25 for
the conclusion that the increment in the active short photoperiod, 22 for long photoperiod). The half-maximum
duration of the population is caused by changes values of the single-unit patterns were determined and the time-
in phase distribution, rather than by changes in point that expressed the middle between both half-max
single-cell activity pattern. values was used to align all single-unit patterns. The mean of
all aligned single-unit patterns was then determined, and
As a consequence of this decreased synchrony in
these averaged single-unit patterns were distributed over the
long photoperiods, neural activity rhythms are 24-h cycle according to a narrow Gaussian distribution for
broader and exhibit a lower peak–trough ampli- short photoperiod (a) or broad Gaussian distribution for long
tude. In contrast, exposure to short days leads to photoperiod (b). The multiunit activity pattern was determined
enhanced synchrony among SCN neuronal by summating these distributed single-unit activity patterns.
populations with a resulting narrowing of the peak Above the figure, the light–dark schedule for each photoperiod
is shown.
and increased amplitude of the SCN rhythm.
149

Phase-shifting capacity as a function of neuronal response to NMDA was not different in the short-
synchrony and long-day preparation, the phase-shifting
effect was significantly larger in slices from short
The photoperiod to which animals have been days. The data indicate that reduction in the mag-
exposed determines the phase-shifting capacity of nitude of phase shifts observed in long days is
the circadian system. The impact of photoperiod intrinsic to the SCN network organization.
on the amplitude of the phase response curve In long days, the neuronal populations of the
(PRC) was first demonstrated by Pittendrigh and SCN are desynchronized, yielding an ensemble
coworkers (1984). Hamsters kept on short days pattern with a low-amplitude rhythm. In short
exhibited a larger amplitude PRC than animals days, the neurons are highly synchronized in
kept on long photoperiod. The findings (also see phase, leading to a high-amplitude rhythm of the
Evans et al., 2004; Refinetti, 2002) raise questions population. Our finding that large shifts are
about the underlying mechanisms. One possibility obtained under conditions in which the neural
is that long light exposure desensitizes the retinal activity rhythm is exhibiting a high-amplitude
ganglion cells to light, and that the smaller shifts rhythms is in some ways surprising in that it does
in long days reflect a difference in retinal light not fit predictions from the limit cycle oscillator
response. In a recent study, we exposed mice to theory that the magnitude of the phase shift is
short and long photoperiods but doubled the inversely related to the amplitude of the rhythm
hourly number of photons in short photoperiods (Pittendrigh et al., 1991; Pulivarthy et al., 2007).
(VanderLeest et al., 2009). Accordingly, the This assumption is based on the notion that a per-
animals received the same amount of photons in turbation of similar strength changes the phase
long and short photoperiod but distributed over a angle of a low-amplitude rhythm more than the
different time span. No decline in the phase- phase angle of a high-amplitude rhythm because
shifting response was observed in the short days, the perturbation is a larger fraction of the radius
and the response remained significantly larger than of the cycle. This prediction has been confirmed
in long days. Although other interpretations are in many biological (and nonbiological) systems
possible, these findings provided a first indication and is valid for single-cell rhythms (Gonyaulax:
that retinal information processing is not responsi- Johnson and Kondo, 1992; Neurospora: Johnson,
ble for the photoperiod-driven changes in the mag- 1999; Lakin-Thomas et al., 1990). We believe that
nitude of light-induced phase shifts. the results of vanderLeest et al. (2009) show that
The other possibility is that the difference in the limit cycle theory does not hold for the net-
phase-shifting capacity is not determined at the work of the SCN in explaining differences in
level of the retina but is determined at the level phase-shifting capacity in long and short days.
of the SCN neuronal network. In this case, the Instead, we favor the view that under long-day
difference in phase-shifting capacity observed in conditions, the SCN network is desynchronized
long and short days should be preserved in vitro. and the individual neurons receive the NMDA
To test this, we prepared brain slices from mice pulse at different phases of their circadian cycle,
held on long and short photoperiods and exam- leading to diverse phase-shifting responses. When
ined the phase-shifting effects of the application the neurons are more synchronized, they would
of NMDA. A variety of evidence suggests that respond more consistently, leading to a larger
NMDA receptors are critical for light-induced net-shift of the ensemble. Using computer
phase shifts and we, and others, have found that simulations, we have found that the phase-shifting
bath application of NMDA can produce phase response of the population can be simulated by
shifts in the neural activity rhythms (Ding et al., distributing neurons over the cycle, according to
1994; VanderLeest et al., 2009). While the acute the observed phase distribution in long and short
150

days (VanderLeest et al., 2009). We supposed short days which needs to be confirmed experi-
that the individual neurons showed limit cycle mentally. Notwithstanding, it has become clear
oscillator behavior, and that they had the same that in the neuronal network of SCN oscillators,
PRCs. Distribution of neurons, in the simulations, at least one prediction of the limit cycle oscillator
was therefore similar to distribution of PRCs. theory does not hold, that is, the larger amplitude
Because the PRCs of individual SCN neurons rhythms showed the larger instead of the smaller
are not known, we gave the neurons either type phase-shifting responses. The experiments demon-
1 or type 0 PRCs, to test the feasibility of our strate that properties emerge at the population
explanation (but only one type per simulation). levels that are not present at the underlying
When we distributed neurons according to the single-cell level.
distribution observed in long photoperiods, the
delaying and advancing parts of the PRC do not
fully overlap. As a result, a particular pulse will Phase resetting is driven by a small population of
lead to delays in some neurons, but to advances highly synchronized neurons
in others. The net result of these divergent shifts
is a minor shift of the population as a whole. Shift work and jet lag cause a temporal disruption
When the PRCs are fully synchronized, a particu- of circadian rhythms that are evidenced by sleep
lar pulse will lead to a unidirectional shift in all disturbances, intestinal problems, and fatigue
neurons, and the net result will be a much larger (Arble et al., 2010; Eastman and Burgess, 2009;
phase response of the population (Fig. 4). Waterhouse et al., 2007). Animals have been
We are aware that the present simulations pres- exposed to an abrupt shift of the light–dark cycle
ent only one possible explanation for the observed to investigate the mechanism underlying phase
difference in phase response magnitude in long and adjustment of the SCN. Following such a shift, a
dissociation of the electrical profiles and the
molecular expression patterns of the ventral and
(a) (b) dorsal SCN becomes visible (Albus et al., 2005;
Pulse Pulse
Davidson et al., 2009; Nagano et al., 2003;
Nakamura et al., 2005; Reddy et al., 2002; Yan
Single cells

and Silver, 2004). Desynchrony between the


dorsal and ventral SCN areas can also be induced
by exposure of rats to short light–dark cycles
(de la Iglesia et al., 2004), but in the case of a jet
6 12 18 24 6 12 18 24 lag, the phase desynchrony is transient and cou-
Population

pling is eventually restored. Following a shifted


light–dark cycle, the neurons in the ventral SCN
appear to shift rapidly followed by a much slower
0 6 12 18 24 0 6 12 18 24
shift of the dorsal SCN (Albus et al., 2005;
CT (h) CT (h)
Nagano et al., 2003; Nakamura et al., 2005). The
Fig. 4. Distributed single-cell phase response curves (PRCs) difference in phase-resetting speed may relate to
determine the ensemble PRC. If the single-cell PRCs are all the differential innervation of the ventral versus
in the same phase of the cycle, the resulting ensemble PRC the dorsal SCN area by the retina (Antle and
will have a large amplitude as a single light pulse (arrow)
Silver, 2005, 2009; Card et al., 1981; Gamble
will trigger a similar phase-shifting response in all cells (a). If
the single-cell PRCs are out of phase, the resulting ensemble et al., 2007; van Esseveldt et al., 2000). The ventral
PRC will have a lower amplitude as the cells will have SCN receives most of the retinal input (Ibata et al.,
divergent responses (b). 1989; Kiss et al., 2008; Morin and Allen, 2006)
151

and shows more robust light-induced changes in work in other hypothalamic neurons demonstrates
electrical activity (Meijer et al., 1998; Shibata that the activity of NKCC1 can be actively
et al., 1984) and gene expression (Dardente et al., regulated such that GABA can switch from being
2002; Guido et al., 1999; Karatsoreos et al., 2004; inhibitory to excitatory as a function of physiologi-
Kuhlman et al., 2003; Schwartz et al., 2000; Yan cal demands (Kim et al., 2011). It is quite possible
et al., 1999). The initial shift of the ventral SCN that a similar dynamic regulation of GABA signal-
may thus be explainable by the stronger influence ing may be occurring in the SCN. Functionally, the
of environmental light input on this part of the exogenous application of GABA can synchronize
nucleus. the electrical activity of SCN neurons (Liu
The rapid shift of the ventral SCN can be easily and Reppert, 2000; Shirakawa et al., 2000), and
observed in an SCN slice, prepared after expo- GABA may synchronize ventral and dorsal SCN
sure to a 6-h delay of the light–dark cycle (Albus following a shifted light cycle (Albus et al., 2005);
et al., 2005). The differential shifts of ventral however, GABAergic signaling does not appear
and dorsal SCN become visible by two peaks in to be required for cultured SCN neurons to remain
the multiunit activity rhythm. The one peak is synchronized (Aton et al., 2006).
delayed by nearly 6 h, while the other peak shows It is evident that the shift of the ventral SCN
no significant shift. A surgical cut, made between ultimately causes a phase shift of the whole SCN
the ventral and dorsal part, made it clear that the network. When an SCN slice is prepared follow-
shifted component originates from the ventral ing the shift of the light–dark cycle, and after
part, and the unshifted component from the three days in constant darkness, the dissociation
dorsal part. The findings also show that, in the within the SCN is still observable (Albus et al.,
intact slice, communication and transfer of electri- 2005; Nagano et al., 2003; Nakamura et al.,
cal activity occur within the SCN circuitry, as in 2005). After day 6 in constant darkness, the
intact slices, both components are visible in the dorsal and ventral SCNs have resynchronized,
dorsal and ventral SCN. The transmittance of and the final peak time is near the time of the
information may rely on GABAergic activity as shifted ventral SCN. These findings indicate a
bicuculline blocked the information transfer, and strong effect of the ventral on the dorsal SCN,
single peaks (i.e., a shifted peak in the ventral while, in contrast, the dorsal SCN has a small
SCN and an unshifted peak in the dorsal SCN) effect on the ventral SCN. The idea that light
were observable in intact slices in the presence information flows from ventral to dorsal is consis-
of bicuculline. tent with anatomical studies (Moore et al., 2002;
GABA plays a role in the transfer of information Romijn et al., 1997).
between ventral and dorsal SCN. SCN neurons An intriguing observation in our experiments
receive a tonic GABAA receptor-mediated synap- was that the shifted component of the multiunit
tic input that, at least partly, originates within the activity pattern was consistently narrower than
SCN itself and peaks during the night (Itri et al., the unshifted component (Fig. 5). A curve fitting
2004; Jiang et al., 1997; Kim and Dudek, 1992; analysis revealed that this observation was highly
Strecker et al., 1997). Although GABA is normally consistent (Rohling et al., 2011). We performed
an inhibitory transmitter within the SCN, for a recordings of neuronal subpopulations following
certain percentage of cells within the SCN circuit, exposure to the shifted light–dark cycle and
GABA can play an excitatory role in communica- observed that only 20% of the neurons peaked
tion (Choi et al., 2008; Irwin and Allen, 2009). This in the middle of the new light phase. This
GABA-mediated excitation is critically dependent indicates that only 20% of the neurons seemed
on the activity of the chloride pump NKCC1 to have shifted on the first day after the shift in the
(Belenky et al., 2010; Choi et al., 2008). Recent light cycle (Rohling et al., 2011). The narrowness
152

(a) simulations (Schaap et al., 2003). Instead, it


indicates that the shifted neurons are highly syn-
300 chronized in phase. Recordings of the phase dis-
MUA (Hz)

tribution of the shifted neurons confirmed this


200
prediction (Rohling et al., 2011). The percentage
of neurons that shifts immediately corresponds
100
with the fraction that shows direct light
0 responsiveness (Meijer et al., 1986) and with the
18 0 6 12 18 subtype of VIP cells that receive direct retinal
ZT (h) input (Kawamoto et al., 2003). These neurons
(b) (c) showed an absence of endogenous rhythmicity
100 10 (Kawamoto et al., 2003), which could make them
more susceptible to phase-shifting influence of
80 8 the light–dark cycle.
From the coupled oscillator theory, we can
Variance (h)
Area (%)

60 6
explain that more tightly synchronized neurons
40 4 have a bigger influence on less synchronized
neurons (Mirollo and Strogatz, 1990; Strogatz,
20 2 2003; Winfree, 1967). Thus, when two groups of
neurons are out of phase, the more synchronized
0 0
Unshifted shifted Before After
group will exert a strong phase-shifting effect on
the less synchronized group. This can be under-
Fig. 5. After a 6-h delay of the LD cycle, one small population of stood as, in a firmly synchronized population,
neurons shifts its activity immediately to the new LD regime, each neuron is exerting its maximal phase-shifting
while another, larger population of cells remains active in the effect at the same time of the cycle, rendering a
old phase. (a) Example of a typical bimodal SCN rhythm,
following a phase delay (black line). The timescale indicates the
strong signal of the group at a particular phase
new Zeitgeber time, that is, ZT 6 is the middle of the shifted of the cycle. Previous studies reported that the
day. When Gaussian functions were fitted through the MUA ventral neurons are dominant over the dorsal
pattern, two different Gaussian functions were found in this neurons and determine to a large extent the final
example. One Gaussian function represents an unshifted phase (Albus et al., 2005; Nagano et al., 2003;
population and one represents a shifted population. In this
example, the percentage of action potentials responsible for the
Nakamura et al., 2005). The synchronization
unshifted population is 83%, while the shifted component among the shifted neurons may contribute to
contains 17% of the action potentials. (b) For all recordings, their dominance over the unshifted population.
subpopulation patterns were recorded and the variance Adjustment to phase advances is more complex
between the different phases of the subpopulations was than adjustments to phase delays and requires
assessed. The subpopulations before the trough, in the
unshifted component, had a much larger variability than the
more days. This cannot be explained by the endog-
subpopulations after the trough, in the shifted component (7.9 enous period of the rhythm, as in nocturnal
vs. 0.7 h, respectively). (c) Our data indicate that the neurons in animals, the period is generally shorter than 24 h,
the shifted component are much more synchronized than the which would facilitate a faster resetting. The slow
neurons in the unshifted component. adaptation to a phase advance results in part from
feedback of the central nervous system to the
SCN (Vansteensel et al., 2003). In the isolated
of the peak of the shifted component is not a triv- SCN, when the SCN is deafferented from this feed-
ial finding when a low number of subpopulations back input, the phase-advancing capacity is there-
is recorded, as can be shown by Monte Carlo fore strongly enhanced (Vansteensel et al., 2003).
153

Cav2.1 calcium channels play a role in the influence Collectively, the phase-resetting studies lead to
of the central nervous system on the SCN (van the following model for resynchronization:
Oosterhout et al., 2008). The CACNA1A gene In response to a shifted light cycle, only the
encodes the ion-conducting, pore-forming a1A sub- oscillators in the ventral SCN are initially shifted.
unit of voltage-gated Cav2.1 (P/Q type) calcium Other areas of the SCN become synchronized to
channels (Ophoff et al., 1996). These channels are the shifted light cycle through interoscillator-
predominantly localized at presynaptic nerve coupling mechanisms. GABA may be involved in
terminals in several brain regions (Westenbroek the transmission between ventral and dorsal
et al., 1995), including the SCN (Chen and van oscillators. The strong effect of the small group of
den Pol, 1998; Cloues and Sather, 2003; Nahm ventral neurons on the dorsal neurons may rely
et al., 2005), and play a key role in mediating on the excitatory effect of GABA on the dorsal
neurotransmitter release (Mintz et al., 1995; SCN as well as on the high degree of synchrony of
Wu et al., 1999). Animals with a gain of function the shifted versus the unshifted cells.
in these channels (used as a migraine model), show The differences in phase synchronization among
enhanced phase-advancing capacity, while their neuronal populations as observed under light–dark
phase delays are unaltered (van Oosterhout et al., cycles, under constant light, under shifted light
2008). However, the phase-resetting capacity of cycles, and under changes in photoperiod have
the SCN in isolation is unaltered in these animals. made clear that the multioscillator structure of the
These findings suggest a role for Cav2.1 channels SCN is strongly involved in the adaptive function of
in mediating influences of the central nervous sys- the SCN. The different environmental conditions
tem on the SCN, and indeed, the synaptic input to can lead to reconfiguration of the phases among
dorsal SCN neurons is enhanced in these animals the SCN population and determine the waveform
(van Oosterhout et al., 2008). The complexity of of the SCN rhythm. These findings are in agree-
phase advances is thus a consequence of an addi- ment with the theory of coupled oscillators, which
tional level of organization that attenuates the proposes that the same system of coupled neurons
phase-shifting capacity of the circadian system. can account for many different states just by chang-
Analysis of the SCN phase synchrony following ing its configuration (Golubitsky et al., 1999;
an advance has shown that also in response to Mirollo and Strogatz, 1990; Strogatz, 2003;
phase advances, the ventral SCN shifts first and Strogatz and Stewart, 1993). These findings also
is followed by a shift of the dorsal SCN, which lead to the question of how the coupling among
is similar to the response to delays (Albus SCN oscillators is affected by aging and diseases.
et al., 2005; Nagano et al., 2003; Nakamura
et al., 2005; Reddy et al., 2002; Yan and Silver,
2002, 2004). However, advances induce a larger Effects of aging and disease
degree of desynchrony within the SCN as com-
pared to phase delays (Rohling et al., 2011). The Disruptions in the circadian system, including
desynchrony may contribute to the inertia in decreased amplitude of rhythmic behaviors and
response to phase advances, and possibly, it is fragmentation of the activity–rest episodes, are
related to the role of the central nervous system commonly associated with aging in humans and
in phase advances, as the activation of input pat- other mammals (Carrier and Bliwise, 2003; Van
hways from other areas may well contribute to Someren, 2000). While undoubtedly many factors
phase desynchrony within the SCN. However, contribute to these changes, a variety of data
the role of extra SCN areas on the SCN is to date (Aujard et al., 2001; Biello, 2009; Satinoff
largely unexplored and deserves attention in et al., 1993; Turek et al., 1995; Watanabe et al.,
future studies. 1995) are emerging that is consistent with the
154

hypothesis that an age-related decline in the output patients and on their caregivers. While the underly-
of the central circadian clock in the SCN may be ing pathology has not yet been identified, several
key. For example, in recent work, in vivo multiunit studies have been carried out using mouse models
recordings were carried out from the SCN and of these neurodegenerative diseases (Kudo et al.,
a brain region that receives robust innervation 2011a,b; Morton et al., 2005; Oakeshott et al.,
from the SCN (subparaventricular zone) in freely 2011; Sterniczuk et al., 2010). Most of these mouse
moving animals. The amplitude of the day–night models exhibit circadian disruptions and there is at
difference in neural activity was substantially least some evidence that treatments designed to
reduced in both brain regions of middle-aged mice stabilize these rhythms can improve other, non-
(Nakamura et al., 2011). Another striking feature motor symptoms of these mice (Maywood et al.,
was the increase in variation in the levels of the 2010; Pallier et al., 2007). In one of the mouse
spontaneous activity. In contrast, the molecular models of HD (the BACHD line) in which the
clockwork in the SCN as measured by PERIOD2 mutated human Htt gene is expressed, the output
levels was not disrupted in middle-aged mice. of the circadian system as measured by locomotor
These results suggest that the age-related disrup- activity, heart rate, and body temperature was pro-
tion in the circadian output occurs before any foundly disrupted early in the life span (Kudo et al.,
disruption of the molecular clockwork. The 2011b). The neural activity rhythms in the SCN,
mechanisms underlying the age-related decline in but not rhythmic PER2 expression, were also
SCN neural activity are unknown but are an impor- reduced in the BACHD mice. A very similar story
tant area for future research. As both sleep states is emerging from a study of a line of alpha-
(Deboer et al., 2003) and locomotor activity synuclein-overexpressing (ASO) mice (Kudo
(Meijer et al., 1997; Schaap and Meijer, 2001; et al., 2011a). Selective deficits were found in the
Yamazaki et al., 1998) can “feedback” to regulate expression of circadian rhythms of locomotor
SCN neural activity, future studies will need to activity, including lower nighttime activity and
address the issue of to what extent the age-related greater fragmentation in the wheel-running activ-
decline in sleep and activity contributes to the ity in this model of PD and other synucleinopathies.
reduced SCN neural activity seen in vivo. Several The temporal distribution of sleep was also altered
studies have shown that the electrophysiological in the ASO mice compared to littermate controls.
activity of aged SCN neurons in vitro is altered in In the ASO mice, the peak–trough expression of
situations where these feedback mechanisms the clock gene PERIOD2 was normal in the SCN;
would not be in operation (Aujard et al., 2001; however, the daytime firing rate of SCN neurons
Biello, 2009; Nygard et al., 2005; Satinoff et al., was reduced in the mutant mice. Together, these
1993; Watanabe et al., 1995). The firing rate data in mouse models of AD, HD, and PD in com-
changes found in the SCN itself could be medi- bination with the clinical symptoms raise the possi-
ated by age-related alterations in synaptic trans- bility that a weakening of circadian output is a core
mission within the circuit as well as changes in feature of neurodegenerative diseases.
ion mechanisms and cellular metabolism intrinsic
to single SCN neurons.
Patients suffering from neurodegenerative Could disease processes alter the synchrony of
disorders including Alzheimer’s disease (AD), the SCN circuit?
Parkinson’s disease (PD), and Huntington’s dis-
ease (HD) commonly exhibit sleep disorders. One mechanism by which aging and disease can
These patients have difficulty sleeping at night alter the function of the circadian system is by
and staying awake during the day. These symptoms reducing the strength of the intercellular coupling
have a major impact on the quality of life of the within the SCN circuit. Recent work in a variety
155

of mouse models of neurodevelopmental and psy- et al., 2011; Vosko et al., 2007; Welsh et al., 2010).
chiatric disorders suggests that alterations in the A number of studies have found evidence that the
balance between synaptic excitation and inhibi- levels of VIP in the SCN are reduced with aging
tion are at the heart of the pathophysiology in and neurodegenerative diseases in humans and
these conditions (Dani et al., 2005; Gogolla rodents (e.g., Duncan et al., 2010; Fahrenkrug
et al., 2009; Milnerwood and Raymond, 2010; et al., 2007; Mazurek et al., 1997; Pereira et al.,
Nelson and Turrigiano, 2008; Shepherd and Katz, 2005; Zhou et al., 1995). These data add a key
2011). Within the SCN circuit, glutamate and support to our suggestion that an important con-
the peptide PACAP are the neurotransmitters sequence of disease pathology will be to alter
released from the RHT that drive the effects of intra-SCN coupling within the circadian circuit
light on the retinorecipient SCN neurons and decrease the synchrony of the SCN popula-
(Colwell, 2011; Morin and Allen, 2006). For tion. Based on the work described in this chapter,
neurons within the circuit, the main transmitter the loss of coupling and decreased synchroniza-
within the SCN is GABA (Moore and Speh, tion would have the consequence of reducing
1993; Okamura et al., 1989) with most neurons the phase-shifting effects of light and other
receiving a constant flux of GABA signaling (Itri phase-shifting agents. The reduced synchrony
et al., 2004; Jiang et al., 1997; Kim and Dudek, would also reduce SCN neural activity and, as a
1992; Strecker et al., 1997). Interestingly, GABA consequence, reduce the amplitude of any SCN-
can mediate both inhibitory and excitatory driven output. A reduction in output would be
postsynaptic effects within the SCN (Choi et al., expected to change the phase relationships
2008; Gribkoff et al., 1999; Irwin and Allen, between the SCN and the peripheral oscillators.
2009; Wagner et al., 1997) with the activity of Thus, many of the key circadian symptoms
the chloride pump NKCC1 being the critical exhibited with aging and diseases of the nervous
determinant (Belenky et al., 2010; Choi et al., system could be explained by a reduction in the
2008; Irwin and Allen, 2009). So, one likely conse- synchrony of the SCN cell population.
quence of a number of diseases of the central ner- These types of disruptions of circadian system
vous system is a change in the balance between caused by altered coupling within the SCN circuit
excitation and inhibition synaptic transmission. are likely to have profound consequences on
While we have no specific evidence for a disease patient health (Hastings et al., 2003; Reddy and
that alters synaptic transmission or coupling within O’Neill, 2010; Takahashi et al., 2008). We believe
the SCN circuit, this remains a very plausible that robust circadian rhythms are essential
mechanism to explain the disruption in circadian to good health. In recent years, a wide range of
function. Recent work shows that the blockade studies have demonstrated that disruption of the
of synaptic transmission within the SCN disrupts circadian system leads to a cluster of symptoms,
both circadian gene expression and neural activity including metabolic deficits (Gale et al.,
rhythms (Deery et al., 2009; Kim et al., 2009). 2011; Marcheva et al., 2010; Turek et al., 2005),
In addition, the SCN circuit highly expresses sev- cardiovascular problems (Bray et al., 2008; Scheer
eral peptides including VIP, GRP, vasopressin, et al., 2009), difficulty sleeping (Reid and Zee,
and PK2. We do not know whether these peptides 2009; Wulff et al., 2009), and cognitive deficits
function as neurotransmitters or more as cofactors (Gerstner et al., 2009; Loh et al., 2010; Wang
within this circuit. Among these peptides, perhaps et al., 2009). Many of these same symptoms are
the best studied is VIP and there is overwhelming seen in aging and neurodegenerative diseases.
evidence for this peptide playing a critical role in This suggests that we should put a greater empha-
coupling or synchronizing cellular oscillators within sis on the development of pharmacological tools
this circuit (Freeman and Herzog, 2011; Maywood and behavioral interventions that can boost
156

neural activity rhythms and the synchrony of the distribution of chloride transporters in the rat
SCN cell population in situations in which the suprachiasmatic nucleus. Neuroscience, 165, 1519–1537.
Berson, D. M., Dunn, F. A., & Takao, M. (2002). Photo-
molecular clock may still be working. transduction by retinal ganglion cells that set the circadian
clock. Science, 295, 1070–1073.
Biello, S. M. (2009). Circadian clock resetting in the mouse
Acknowledgment changes with age. Age (Dordrecht, Netherlands), 31,
293–303.
Bray, M. S., Shaw, C. A., Moore, M. W., Garcia, R. A.,
This work was supported by NWO travel grant Zanquetta, M. M., Durgan, D. J., et al. (2008). Disruption
(040.11.242) for collaboration between C. S. C. of the circadian clock within the cardiomyocyte influences
and S. M. myocardial contractile function, metabolism, and gene
expression. American Journal of Physiology. Heart and
Circulatory Physiology, 294, H1036–H1047.
Brown, T. M., Banks, J. R., & Piggins, H. D. (2006). A novel
References suction electrode recording technique for monitoring circa-
dian rhythms in single and multiunit discharge from brain
Abrahamson, E. E., & Moore, R. Y. (2001). Suprachiasmatic slices. The Journal of Neuroscience, 156, 173–181.
nucleus in the mouse: Retinal innervation, intrinsic organiza- Brown, T. M., & Piggins, H. D. (2007). Electrophysiology of
tion and efferent projections. Brain Research, 916, 172–191. the suprachiasmatic circadian clock. Progress in Neurobiol-
Albus, H., Vansteensel, M. J., Michel, S., Block, G. D., & ogy, 82, 229–255.
Meijer, J. H. (2005). A GABAergic mechanism is necessary Brown, T. M., & Piggins, H. D. (2009). Spatiotemporal hetero-
for coupling dissociable ventral and dorsal regional geneity in the electrical activity of suprachiasmatic nuclei
oscillators within the circadian clock. Current Biology, 15, neurons and their response to photoperiod. Journal of
886–893. Biological Rhythms, 24, 44–54.
Antle, M. C., LeSauter, J., & Silver, R. (2005). Neurogenesis Card, J. P., Brecha, N., Karten, H. J., & Moore, R. Y. (1981).
and ontogeny of specific cell phenotypes within the hamster Immunocytochemical localization of vasoactive intestinal
suprachiasmatic nucleus. Brain Research. Developmental polypeptide-containing cells and processes in the supra-
Brain Research, 157, 8–18. chiasmatic nucleus of the rat: Light and electron microscopic
Antle, M. C., & Silver, R. (2005). Orchestrating time: analysis. The Journal of Neuroscience, 1, 1289–1303.
Arrangements of the brain circadian clock. Trends in Carr, A. J., Johnston, J. D., Semikhodskii, A. G., Nolan, T.,
Neurosciences, 28, 145–151. Cagampang, F. R., Stirland, J. A., et al. (2003). Photoperiod
Antle, M. C., & Silver, R. (2009). Neural basis of timing and differentially regulates circadian oscillators in central and
anticipatory behaviors. The European Journal of Neurosci- peripheral tissues of the Syrian hamster. Current Biology,
ence, 30, 1643–1649. 13, 1543–1548.
Arble, D. M., Ramsey, K. M., Bass, J., & Turek, F. W. (2010). Carrier, J., & Bliwise, D. L. (2003). Sleep and circadian rhythms
Circadian disruption and metabolic disease: Findings from in normal aging. In M. Billiard & A. Kent (Eds.), Sleep: Phys-
animal models. Best Practice & Research. Clinical Endocri- iology investigations, and medicine (pp. 297–333). New York:
nology & Metabolism, 24, 785–800. Kluwer Academic/Plenum.
Aschoff, J. (1979). Circadian rhythms: Influences of internal Challet, E. (2007). Minireview: Entrainment of the supra-
and external factors on the period measured in constant chiasmatic clockwork in diurnal and nocturnal mammals.
conditions. Zeitschrift für Tierpsychologie, 49, 225–249. Endocrinology, 148, 5648–5655.
Aton, S. J., Huettner, J. E., Straume, M., & Herzog, E. D. Chen, G., & van den Pol, A. N. (1998). Presynaptic GABA(B)
(2006). GABA and Gi/o differentially control circadian autoreceptor modulation of P/Q-type calcium channels and
rhythms and synchrony in clock neurons. Proceedings of GABA release in rat suprachiasmatic nucleus neurons.
the National Academy of Sciences of the United States of The Journal of Neuroscience, 18, 1913–1922.
America, 103, 19188–19193. Cheng, M. Y., Bittman, E. L., Hattar, S., & Zhou, Q. Y.
Aujard, F., Herzog, E. D., & Block, G. D. (2001). Circadian (2005). Regulation of prokineticin 2 expression by light
rhythms in firing rate of individual suprachiasmatic nucleus and the circadian clock. BMC Neuroscience, 6, 17.
neurons from adult and middle-aged mice. Neuroscience, Choi, H. J., Lee, C. J., Schroeder, A., Kim, Y. S., Jung, S. H.,
106, 255–261. Kim, J. S., et al. (2008). Excitatory actions of GABA in the
Belenky, M. A., Sollars, P. J., Mount, D. B., Alper, S. L., suprachiasmatic nucleus. The Journal of Neuroscience, 28,
Yarom, Y., & Pickard, G. E. (2010). Cell-type specific 5450–5459.
157

Cloues, R. K., & Sather, W. A. (2003). Afterhyperpolarization Evans, J. A., Elliott, J. A., & Gorman, M. R. (2004). Photope-
regulates firing rate in neurons of the suprachiasmatic riod differentially modulates photic and nonphotic phase
nucleus. The Journal of Neuroscience, 23, 1593–1604. response curves of hamsters. American Journal of Physiol-
Colwell, C. S. (2011). Linking neural activity and molecular ogy. Regulatory, Integrative and Comparative Physiology,
oscillations in the SCN. Nature Reviews. Neuroscience, 12, 286, R539–R546.
553–569. Evans, J. A., Leise, T. L., Castanon-Cervantes, O., &
Dani, V. S., Chang, Q., Maffei, A., Turrigiano, G. G., Davidson, A. J. (2011). Intrinsic regulation of spatiotempo-
Jaenisch, R., & Nelson, S. B. (2005). Reduced cortical ral organization within the suprachiasmatic nucleus. PLoS
activity due to a shift in the balance between excitation One, 6, e15869.
and inhibition in a mouse model of Rett syndrome. Pro- Fahrenkrug, J., Popovic, N., Georg, B., Brundin, P., &
ceedings of the National Academy of Sciences of the United Hannibal, J. (2007). Decreased VIP and VPAC2 receptor
States of America, 102, 12560–12565. expression in the biological clock of the R6/2 Huntington’s
Dardente, H., Poirel, V. J., Klosen, P., Pevet, P., & Masson- disease mouse. Journal of Molecular Neuroscience, 31,
Pevet, M. (2002). Per and neuropeptide expression in the rat 139–148.
suprachiasmatic nuclei: Compartmentalization and differential Freeman, G. M., Jr., & Herzog, E. D. (2011). Neuropeptides
cellular induction by light. Brain Research, 958, 261–271. go the distance for circadian synchrony. Proceedings of the
Davidson, A. J., Castanon-Cervantes, O., Leise, T. L., National Academy of Sciences of the United States of
Molyneux, P. C., & Harrington, M. E. (2009). Visualizing America, 108, 13883–13884.
jet lag in the mouse suprachiasmatic nucleus and peripheral Gale, J. E., Cox, H. I., Qian, J., Block, G. D., Colwell, C. S., &
circadian timing system. The European Journal of Neurosci- Matveyenko, A. V. (2011). Disruption of circadian rhythms
ence, 29, 171–180. accelerates development of diabetes through pancreatic
de la Iglesia, H. O., Meyer, J., & Schwartz, W. J. (2004). Using beta-cell loss and dysfunction. Journal of Biological
Per gene expression to search for photoperiodic oscillators Rhythms, 26, 423–433.
in the hamster suprachiasmatic nucleus. Brain Research. Gamble, K. L., Allen, G. C., Zhou, T., & McMahon, D. G.
Molecular Brain Research, 127, 121–127. (2007). Gastrin-releasing peptide mediates light-like
Deboer, T., Vansteensel, M. J., Detari, L., & Meijer, J. H. resetting of the suprachiasmatic nucleus circadian pace-
(2003). Sleep states alter activity of suprachiasmatic nucleus maker through cAMP response element-binding protein
neurons. Nature Neuroscience, 6, 1086–1090. and Per1 activation. The Journal of Neuroscience, 27,
Deery, M. J., Maywood, E. S., Chesham, J. E., Sladek, M., 12078–12087.
Karp, N. A., Green, E. W., et al. (2009). Proteomic analysis Gerstner, J. R., Lyons, L. C., Wright, K. P., Jr., Loh, D. H.,
reveals the role of synaptic vesicle cycling in sustaining the Rawashdeh, O., Eckel-Mahan, K. L., et al. (2009). Cycling
suprachiasmatic circadian clock. Current Biology, 19, behavior and memory formation. The Journal of Neurosci-
2031–2036. ence, 29, 12824–12830.
Depres-Brummer, P., Levi, F., Metzger, G., & Touitou, Y. Gillette, M. U. (1996). Regulation of entrainment pathways by
(1995). Light-induced suppression of the rat circadian system. the suprachiasmatic circadian clock: Sensitivities to second
The American Journal of Physiology, 268, R1111–R1116. messengers. Progress in Brain Research, 111, 121–132.
Ding, J. M., Chen, D., Weber, E. T., Faiman, L. E., Gogolla, N., Leblanc, J. J., Quast, K. B., Sudhof, T. C.,
Rea, M. A., & Gillette, M. U. (1994). Resetting the Fagiolini, M., & Hensch, T. K. (2009). Common circuit
biological clock: Mediation of nocturnal circadian shifts by defect of excitatory–inhibitory balance in mouse models of
glutamate and NO. Science, 266, 1713–1717. autism. Journal of Neurodevelopmental Disorders, 1,
Duncan, M. J., Hester, J. M., Hopper, J. A., & Franklin, K. M. 172–181.
(2010). The effects of aging and chronic fluoxetine treat- Golombek, D. A., & Rosenstein, R. E. (2010). Physiology of cir-
ment on circadian rhythms and suprachiasmatic nucleus cadian entrainment. Physiological Reviews, 90, 1063–1102.
expression of neuropeptide genes and 5-HT1B receptors. Golubitsky, M., Stewart, I., Buono, P. L., & Collins, J. J.
The European Journal of Neuroscience, 31, 1646–1654. (1999). Symmetry in locomotor central pattern generators
Eastman, C. I., & Burgess, H. J. (2009). How to travel the and animal gaits. Nature, 401, 693–695.
world without jet lag. Sleep Medicine Clinics, 4, 241–255. Green, D. J., & Gillette, R. (1982). Circadian rhythm of firing
Eastman, C., & Rechtschaffen, A. (1983). Circadian tempera- rate recorded from single cells in the rat suprachiasmatic
ture and wake rhythms of rats exposed to prolonged contin- brain slice. Brain Research, 245, 198–200.
uous illumination. Physiology and Behavior, 31, 417–427. Gribkoff, V. K., Pieschl, R. L., Wisialowski, T. A., Park, W. K.,
Ebling, F. J., & Barrett, P. (2008). The regulation of seasonal Strecker, G. J., de Jeu, M. T., et al. (1999). A reexamination
changes in food intake and body weight. Journal of Neuro- of the role of GABA in the mammalian suprachiasmatic
endocrinology, 20, 827–833. nucleus. Journal of Biological Rhythms, 14, 126–130.
158

Groos, G., & Hendriks, J. (1982). Circadian rhythms in electri- mouse suprachiasmatic nucleus. Journal of Neurophysiol-
cal discharge of rat suprachiasmatic neurones recorded ogy, 92, 311–319.
in vitro. Neuroscience Letters, 34, 283–288. Jac, M., Kiss, A., Sumova, A., Illnerova, H., & Jezova, D.
Guido, M. E., de Guido, L. B., Goguen, D., Robertson, H. A., (2000a). Daily profiles of arginine vasopressin mRNA in
& Rusak, B. (1999). Daily rhythm of spontaneous immedi- the suprachiasmatic, supraoptic and paraventricular nuclei
ate-early gene expression in the rat suprachiasmatic nucleus. of the rat hypothalamus under various photoperiods. Brain
Journal of Biological Rhythms, 14, 275–280. Research, 887, 472–476.
Hannibal, J., Moller, M., Ottersen, O. P., & Fahrenkrug, J. Jac, M., Sumova, A., & Illnerova, H. (2000b). c-Fos rhythm in
(2000). PACAP and glutamate are co-stored in the ret- subdivisions of the rat suprachiasmatic nucleus under artifi-
inohypothalamic tract. The Journal of Comparative Neurol- cial and natural photoperiods. American Journal of Physiol-
ogy, 418, 147–155. ogy. Regulatory, Integrative and Comparative Physiology,
Hastings, M. H., Reddy, A. B., Garabette, M., King, V. M., 279, R2270–R2276.
Chahad-Ehlers, S., O’Brien, J., et al. (2003). Expression of Jiang, Z. G., Yang, Y., Liu, Z. P., & Allen, C. N. (1997). Mem-
clock gene products in the suprachiasmatic nucleus in relation brane properties and synaptic inputs of suprachiasmatic
to circadian behaviour. Novartis Foundation Symposium, 253, nucleus neurons in rat brain slices. The Journal of Physiol-
203–217. ogy, 499(Pt 1), 141–159.
Hattar, S., Lucas, R. J., Mrosovsky, N., Thompson, S., Johnson, C. H. (1999). Forty years of PRCs—What have we
Douglas, R. H., Hankins, M. W., et al. (2003). Melanopsin learned? Chronobiology International, 16, 711–743.
and rod–cone photoreceptive systems account for all major Johnson, C. H., & Kondo, T. (1992). Light pulses induce “sin-
accessory visual functions in mice. Nature, 424, 76–81. gular” behavior and shorten the period of the circadian pho-
Hazlerigg, D. G., Ebling, F. J., & Johnston, J. D. (2005). totaxis rhythm in the CW15 strain of Chlamydomonas.
Photoperiod differentially regulates gene expression rhythms Journal of Biological Rhythms, 7, 313–327.
in the rostral and caudal SCN. Current Biology, 15, Johnston, J. D., Cagampang, F. R., Stirland, J. A., Carr, A. J.,
R449–R450. White, M. R., Davis, J. R., et al. (2003). Evidence for an
Hazlerigg, D., & Loudon, A. (2008). New insights into ancient endogenous per1- and ICER-independent seasonal timer
seasonal life timers. Current Biology, 18, R795–R804. in the hamster pituitary gland. The FASEB Journal, 17,
Herzog, E. D. (2007). Neurons and networks in daily rhythms. 810–815.
Nature Reviews. Neuroscience, 8, 790–802. Johnston, J. D., Ebling, F. J., & Hazlerigg, D. G. (2005). Pho-
Houben, T., Deboer, T., van Oosterhout, F., & Meijer, J. H. toperiod regulates multiple gene expression in the
(2009). Correlation with behavioral activity and rest implies suprachiasmatic nuclei and pars tuberalis of the Siberian
circadian regulation by SCN neuronal activity levels. hamster (Phodopus sungorus). The European Journal of
Journal of Biological Rhythms, 24, 477–487. Neuroscience, 21, 2967–2974.
Ibata, Y., Takahashi, Y., Okamura, H., Kawakami, F., Kalsbeek, A., Palm, I. F., La Fleur, S. E., Scheer, F. A.,
Terubayashi, H., Kubo, T., et al. (1989). Vasoactive intesti- Perreau-Lenz, S., Ruiter, M., et al. (2006). SCN outputs
nal peptide (VIP)-like immunoreactive neurons located in and the hypothalamic balance of life. Journal of Biological
the rat suprachiasmatic nucleus receive a direct retinal Rhythms, 21, 458–469.
projection. Neuroscience Letters, 97, 1–5. Karatsoreos, I. N., Yan, L., LeSauter, J., & Silver, R. (2004).
Inagaki, N., Honma, S., Ono, D., Tanahashi, Y., & Honma, K. Phenotype matters: Identification of light-responsive cells
(2007). Separate oscillating cell groups in mouse in the mouse suprachiasmatic nucleus. The Journal of Neu-
suprachiasmatic nucleus couple photoperiodically to the onset roscience, 24, 68–75.
and end of daily activity. Proceedings of the National Academy Kawamoto, K., Nagano, M., Kanda, F., Chihara, K.,
of Sciences of the United States of America, 104, 7664–7669. Shigeyoshi, Y., & Okamura, H. (2003). Two types of VIP
Inouye, S. T., & Kawamura, H. (1979). Persistence of circadian neuronal components in rat suprachiasmatic nucleus. Jour-
rhythmicity in a mammalian hypothalamic “island” con- nal of Neuroscience Research, 74, 852–857.
taining the suprachiasmatic nucleus. Proceedings of the Kim, Y. I., & Dudek, F. E. (1992). Intracellular electrophysio-
National Academy of Sciences of the United States of logical study of suprachiasmatic nucleus neurons in rodents:
America, 76, 5962–5966. Inhibitory synaptic mechanisms. The Journal of Physiology,
Irwin, R. P., & Allen, C. N. (2009). GABAergic signaling 458, 247–260.
induces divergent neuronal Ca2 þ responses in the Kim, J. S., Kim, W. B., Kim, Y. B., Lee, Y., Kim, Y. S.,
suprachiasmatic nucleus network. The European Journal of Shen, F. Y., et al. (2011). Chronic hyperosmotic stress con-
Neuroscience, 30, 1462–1475. verts GABAergic inhibition into excitation in vasopressin
Itri, J., Michel, S., Waschek, J. A., & Colwell, C. S. (2004). Cir- and oxytocin neurons in the rat. The Journal of Neurosci-
cadian rhythm in inhibitory synaptic transmission in the ence, 31, 13312–13322.
159

Kim, S. M., Power, A., Brown, T. M., Constance, C. M., breast cancer cell lines with high and low glycolytic rates.
Coon, S. L., Nishimura, T., et al. (2009). Deletion of the The Journal of Biological Chemistry, 272, 4941–4952.
secretory vesicle proteins IA-2 and IA-2beta disrupts circa- Meijer, J. H., Groos, G. A., & Rusak, B. (1986). Luminance cod-
dian rhythms of cardiovascular and physical activity. The ing in a circadian pacemaker: The suprachiasmatic nucleus of
FASEB Journal, 23, 3226–3232. the rat and the hamster. Brain Research, 382, 109–118.
Kiss, J., Csaki, A., Csaba, Z., & Halasz, B. (2008). Synaptic Meijer, J. H., Schaap, J., Watanabe, K., & Albus, H. (1997).
contacts of vesicular glutamate transporter 2 fibres on chem- Multiunit activity recordings in the suprachiasmatic nuclei:
ically identified neurons of the hypothalamic suprachiasmatic In vivo versus in vitro models. Brain Research, 753, 322–327.
nucleus of the rat. The European Journal of Neuroscience, Meijer, J. H., Watanabe, K., Schaap, J., Albus, H., & Detari, L.
28, 1760–1774. (1998). Light responsiveness of the suprachiasmatic nucleus:
Kudo, T., Loh, D. H., Truong, D., Wu, Y., & Colwell, C. S. Long-term multiunit and single-unit recordings in freely
(2011a). Circadian dysfunction in a mouse model of moving rats. The Journal of Neuroscience, 18, 9078–9087.
Parkinson’s disease. Experimental Neurology, 232(1), 66–75. Milnerwood, A. J., & Raymond, L. A. (2010). Early synaptic
Kudo, T., Schroeder, A., Loh, D. H., Kuljis, D., Jordan, M. C., pathophysiology in neurodegeneration: Insights from
Roos, K. P., et al. (2011b). Dysfunctions in circadian behav- Huntington’s disease. Trends in Neurosciences, 33, 513–523.
ior and physiology in mouse models of Huntington’s disease. Mintz, I. M., Sabatini, B. L., & Regehr, W. G. (1995). Calcium
Experimental Neurology, 228, 80–90. control of transmitter release at a cerebellar synapse.
Kuhlman, S. J., Silver, R., Le Sauter, J., Bult-Ito, A., & Neuron, 15, 675–688.
McMahon, D. G. (2003). Phase resetting light pulses induce Mirollo, R. E., & Strogatz, S. H. (1990). Synchronization of
Per1 and persistent spike activity in a subpopulation of pulse-coupled biological oscillators. SIAM Journal on
biological clock neurons. The Journal of Neuroscience, 23, Applied Mathematics, 50, 1645–1662.
1441–1450. Moore, R. Y., & Silver, R. (1998). Suprachiasmatic nucleus
Lakin-Thomas, P. L., Cote, G. G., & Brody, S. (1990). Circadian organization. Chronobiology International, 15, 475–487.
rhythms in Neurospora crassa: Biochemistry and genetics. Moore, R. Y., & Speh, J. C. (1993). GABA is the principal
Critical Reviews in Microbiology, 17, 365–416. neurotransmitter of the circadian system. Neuroscience
Liu, C., & Reppert, S. M. (2000). GABA synchronizes clock Letters, 150, 112–116.
cells within the suprachiasmatic circadian clock. Neuron, Moore, R. Y., Speh, J. C., & Leak, R. K. (2002).
25, 123–128. Suprachiasmatic nucleus organization. Cell and Tissue
Loh, D. H., Navarro, J., Hagopian, A., Wang, L. M., Research, 309, 89–98.
Deboer, T., & Colwell, C. S. (2010). Rapid changes in the Morin, L. P., & Allen, C. N. (2006). The circadian visual sys-
light/dark cycle disrupt memory of conditioned fear in mice. tem, 2005. Brain Research Reviews, 51, 1–60.
PLoS One, 5, e12546. Morton, A. J., Wood, N. I., Hastings, M. H., Hurelbrink, C.,
Marcheva, B., Ramsey, K. M., Buhr, E. D., Kobayashi, Y., Barker, R. A., & Maywood, E. S. (2005). Disintegration of
Su, H., Ko, C. H., et al. (2010). Disruption of the clock com- the sleep-wake cycle and circadian timing in Huntington’s
ponents CLOCK and BMAL1 leads to hypoinsulinaemia disease. The Journal of Neuroscience, 25, 157–163.
and diabetes. Nature, 466, 627–631. Mrugala, M., Zlomanczuk, P., Jagota, A., & Schwartz, W. J.
Maywood, E. S., Chesham, J. E., O’Brien, J. A., & (2000). Rhythmic multiunit neural activity in slices of ham-
Hastings, M. H. (2011). A diversity of paracrine signals ster suprachiasmatic nucleus reflect prior photoperiod.
sustains molecular circadian cycling in suprachiasmatic American Journal of Physiology. Regulatory, Integrative
nucleus circuits. Proceedings of the National Academy of and Comparative Physiology, 278, R987–R994.
Sciences of the United States of America, 108, 14306–14311. Nagano, M., Adachi, A., Nakahama, K., Nakamura, T.,
Maywood, E. S., Fraenkel, E., McAllister, C. J., Wood, N., Tamada, M., Meyer-Bernstein, E., et al. (2003). An abrupt
Reddy, A. B., Hastings, M. H., et al. (2010). Disruption of shift in the day/night cycle causes desynchrony in the mam-
peripheral circadian timekeeping in a mouse model of malian circadian center. The Journal of Neuroscience, 23,
Huntington’s disease and its restoration by temporally sched- 6141–6151.
uled feeding. The Journal of Neuroscience, 30, 10199–10204. Nahm, S. S., Farnell, Y. Z., Griffith, W., & Earnest, D. J.
Maywood, E. S., Reddy, A. B., Wong, G. K., O’Neill, J. S., (2005). Circadian regulation and function of voltage-depen-
O’Brien, J. A., McMahon, D. G., et al. (2006). Synchroniza- dent calcium channels in the suprachiasmatic nucleus. The
tion and maintenance of timekeeping in suprachiasmatic cir- Journal of Neuroscience, 25, 9304–9308.
cadian clock cells by neuropeptidergic signaling. Current Naito, E., Watanabe, T., Tei, H., Yoshimura, T., & Ebihara, S.
Biology, 16, 599–605. (2008). Reorganization of the suprachiasmatic nucleus cod-
Mazurek, S., Michel, A., & Eigenbrodt, E. (1997). Effect of ing for day length. Journal of Biological Rhythms, 23,
extracellular AMP on cell proliferation and metabolism of 140–149.
160

Nakamura, T. J., Nakamura, W., Yamazaki, S., Kudo, T., & J. M. Collins (Eds.), Photoperiodic regulation of insect
Cutler, T., Colwell, C. S., et al. (2011). Age-related decline and molluscan hormones (pp. 26–47). Novartis Foundation
in circadian output. The Journal of Neuroscience, 31, Symposia, Ciba Foundation Symposium 104, London:
10201–10205. Pitman.
Nakamura, W., Yamazaki, S., Takasu, N. N., Mishima, K., & Pittendrigh, C. S., Kyner, W. T., & Takamura, T. (1991). The
Block, G. D. (2005). Differential response of Period 1 amplitude of circadian oscillations: Temperature depen-
expression within the suprachiasmatic nucleus. The Journal dence, latitudinal clines, and the photoperiodic time mea-
of Neuroscience, 25, 5481–5487. surement. Journal of Biological Rhythms, 6, 299–313.
Nelson, S. B., & Turrigiano, G. G. (2008). Strength through Pulivarthy, S. R., Tanaka, N., Welsh, D. K., De Haro, L.,
diversity. Neuron, 60, 477–482. Verma, I. M., & Panda, S. (2007). Reciprocity between
Nuesslein-Hildesheim, B., O’Brien, J. A., Ebling, F. J., phase shifts and amplitude changes in the mammalian circa-
Maywood, E. S., & Hastings, M. H. (2000). The circadian cycle dian clock. Proceedings of the National Academy of Sciences
of mPER clock gene products in the suprachiasmatic nucleus of the United States of America, 104, 20356–20361.
of the Siberian hamster encodes both daily and seasonal time. Quintero, J. E., Kuhlman, S. J., & McMahon, D. G. (2003).
The European Journal of Neuroscience, 12, 2856–2864. The biological clock nucleus: A multiphasic oscillator net-
Nygard, M., Hill, R. H., Wikstrom, M. A., & Kristensson, K. work regulated by light. The Journal of Neuroscience, 23,
(2005). Age-related changes in electrophysiological pro- 8070–8076.
perties of the mouse suprachiasmatic nucleus in vitro. Brain Reddy, A. B., Field, M. D., Maywood, E. S., & Hastings, M. H.
Research Bulletin, 65, 149–154. (2002). Differential resynchronisation of circadian clock
Oakeshott, S., Balci, F., Filippov, I., Murphy, C., Port, R., gene expression within the suprachiasmatic nuclei of mice
Connor, D., et al. (2011). Circadian abnormalities in motor subjected to experimental jet lag. The Journal of Neurosci-
activity in a BAC transgenic mouse model of Huntington’s ence, 22, 7326–7330.
disease. PLoS Currents, 3, RRN1225. Reddy, A. B., & O’Neill, J. S. (2010). Healthy clocks, healthy
Ohta, H., Yamazaki, S., & McMahon, D. G. (2005). Constant body, healthy mind. Trends in Cell Biology, 20, 36–44.
light desynchronizes mammalian clock neurons. Nature Refinetti, R. (2002). Compression and expansion of circadian
Neuroscience, 8, 267–269. rhythm in mice under long and short photoperiods. Integra-
Okamura, H., Berod, A., Julien, J. F., Geffard, M., tive Physiological and Behavioral Science, 37, 114–127.
Kitahama, K., Mallet, J., et al. (1989). Demonstration of Reid, K. J., & Zee, P. C. (2009). Circadian rhythm disorders.
GABAergic cell bodies in the suprachiasmatic nucleus: In Seminars in Neurology, 29, 393–405.
situ hybridization of glutamic acid decarboxylase (GAD) Reppert, S. M., & Weaver, D. R. (2002). Coordination of
mRNA and immunocytochemistry of GAD and GABA. circadian timing in mammals. Nature, 418, 935–941.
Neuroscience Letters, 102, 131–136. Rohling, J., Meijer, J. H., VanderLeest, H. T., & Admiraal, J.
Ophoff, R. A., Terwindt, G. M., Vergouwe, M. N., vanEijk, R., (2006a). Phase differences between SCN neurons and their
Oefner, P. J., Hoffman, S. M. G., et al. (1996). Familial role in photoperiodic encoding; a simulation of ensemble
hemiplegic migraine and episodic ataxia type-2 are caused patterns using recorded single unit electrical activity patterns.
by mutations in the Ca2þ channel gene CACNL1A4. Cell, Journal of Physiology, Paris, 100, 261–270.
87, 543–552. Rohling, J. H., Vanderleest, H. T., Michel, S.,
Pallier, P. N., Maywood, E. S., Zheng, Z., Chesham, J. E., Vansteensel, M. J., & Meijer, J. H. (2011). Phase resetting
Inyushkin, A. N., Dyball, R., et al. (2007). Pharmacological of the mammalian circadian clock relies on a rapid shift of
imposition of sleep slows cognitive decline and reverses a small population of pacemaker neurons. PLoS One, 6,
dysregulation of circadian gene expression in a transgenic e25437.
mouse model of Huntington’s disease. The Journal of Rohling, J., Wolters, L., & Meijer, J. H. (2006b). Simulation of
Neuroscience, 27, 7869–7878. day-length encoding in the SCN: From single-cell to tissue-
Pereira, P. A., Cardoso, A., & Paula-Barbosa, M. M. (2005). level organization. Journal of Biological Rhythms, 21,
Nerve growth factor restores the expression of vasopressin 301–313.
and vasoactive intestinal polypeptide in the suprachiasmatic Romijn, H. J., Sluiter, A. A., Pool, C. W., Wortel, J., &
nucleus of aged rats. Brain Research, 1048, 123–130. Buijs, R. M. (1997). Evidence from confocal fluorescence
Pittendrigh, C. S., & Daan, S. (1976). A functional analysis of microscopy for a dense, reciprocal innervation between
circadian pacemakers in nocturnal rodents. I. The stability AVP-, somatostatin-, VIP/PHI-, GRP-, and VIP/PHI/GRP-
and lability of spontaneous frequency. Journal of Compara- immunoreactive neurons in the rat suprachiasmatic nucleus.
tive Physiology A, 106, 223–252. The European Journal of Neuroscience, 9, 2613–2623.
Pittendrigh, C. S., Elliott, J., & Takamura, T. (1984). The cir- Satinoff, E., Li, H., Tcheng, T. K., Liu, C., McArthur, A. J.,
cadian component in photoperiodic induction. In R. Porter Medanic, M., et al. (1993). Do the suprachiasmatic nuclei
161

oscillate in old rats as they do in young ones? The American Strogatz, S. H., & Stewart, I. (1993). Coupled oscillators
Journal of Physiology, 265, R1216–R1222. and biological synchronization. Scientific American, 269,
Schaap, J., Albus, H., VanderLeest, H. T., Eilers, P. H., 102–109.
Detari, L., & Meijer, J. H. (2003). Heterogeneity of rhythmic Sumova, A., Bendova, Z., Sladek, M., Kovacikova, Z., &
suprachiasmatic nucleus neurons: Implications for circadian Illnerova, H. (2004). Seasonal molecular timekeeping within
waveform and photoperiodic encoding. Proceedings of the the rat circadian clock. Physiological Research, 53(Suppl. 1),
National Academy of Sciences of the United States of America, S167–S176.
100, 15994–15999. Sumova, A., Jac, M., Sladek, M., Sauman, I., & Illnerova, H.
Schaap, J., & Meijer, J. H. (2001). Opposing effects of behavioural (2003). Clock gene daily profiles and their phase relation-
activity and light on neurons of the suprachiasmatic nucleus. ship in the rat suprachiasmatic nucleus are affected by pho-
The European Journal of Neuroscience, 13, 1955–1962. toperiod. Journal of Biological Rhythms, 18, 134–144.
Scheer, F. A., Hilton, M. F., Mantzoros, C. S., & Shea, S. A. Sumova, A., Sladek, M., Jac, M., & Illnerova, H. (2002). The
(2009). Adverse metabolic and cardiovascular consequences circadian rhythm of Per1 gene product in the rat
of circadian misalignment. Proceedings of the National suprachiasmatic nucleus and its modulation by seasonal
Academy of Sciences of the United States of America, 106, changes in daylength. Brain Research, 947, 260–270.
4453–4458. Sumova, A., Travnickova, Z., & Illnerova, H. (1995). Memory
Schwartz, W. J., Carpino, A., Jr., de la Iglesia, H. O., Baler, R., on long but not on short days is stored in the rat
Klein, D. C., Nakabeppu, Y., et al. (2000). Differential reg- suprachiasmatic nucleus. Neuroscience Letters, 200, 191–194.
ulation of fos family genes in the ventrolateral and Takahashi, J. S., Hong, H. K., Ko, C. H., & McDearmon, E. L.
dorsomedial subdivisions of the rat suprachiasmatic nucleus. (2008). The genetics of mammalian circadian order and dis-
Neuroscience, 98, 535–547. order: Implications for physiology and disease. Nature
Shepherd, G. M., & Katz, D. M. (2011). Synaptic microcircuit Reviews. Genetics, 9, 764–775.
dysfunction in genetic models of neurodevelopmental dis- Tournier, B. B., Menet, J. S., Dardente, H., Poirel, V. J.,
orders: Focus on Mecp2 and Met. Current Opinion in Neu- Malan, A., Masson-Pevet, M., et al. (2003). Photoperiod dif-
robiology, 21, 827–833. ferentially regulates clock genes’ expression in the
Shibata, S., Liou, S., Ueki, S., & Oomura, Y. (1984). Influence suprachiasmatic nucleus of Syrian hamster. Neuroscience,
of environmental light–dark cycle and enucleation on activity 118, 317–322.
of suprachiasmatic neurons in slice preparations. Brain Turek, F. W., Joshu, C., Kohsaka, A., Lin, E., Ivanova, G.,
Research, 302, 75–81. McDearmon, E., et al. (2005). Obesity and metabolic syn-
Shibata, S., Oomura, Y., Kita, H., & Hattori, K. (1982). drome in circadian Clock mutant mice. Science, 308,
Circadian rhythmic changes of neuronal activity in the 1043–1045.
suprachiasmatic nucleus of the rat hypothalamic slice. Brain Turek, F. W., Penev, P., Zhang, Y., Van Reeth, O.,
Research, 247, 154–158. Takahashi, J. S., & Zee, P. (1995). Alterations in the circa-
Shirakawa, T., Honma, S., Katsuno, Y., Oguchi, H., & dian system in advanced age. CIBA Foundation Symposium,
Honma, K. I. (2000). Synchronization of circadian firing 183, 212–226.
rhythms in cultured rat suprachiasmatic neurons. The Euro- van Esseveldt, K. E., Lehman, M. N., & Boer, G. J. (2000).
pean Journal of Neuroscience, 12, 2833–2838. The suprachiasmatic nucleus and the circadian time-keeping
Sosniyenko, S., Hut, R. A., Daan, S., & Sumova, A. (2009). system revisited. Brain Research. Brain Research Reviews,
Influence of photoperiod duration and light–dark transitions 33, 34–77.
on entrainment of Per1 and Per2 gene and protein expres- van Oosterhout, F., Michel, S., Deboer, T., Houben, T., van de
sion in subdivisions of the mouse suprachiasmatic nucleus. Ven, R. C. G., Albus, H., et al. (2008). Enhanced circadian
The European Journal of Neuroscience, 30, 1802–1814. phase resetting in R192Q Ca(v)2.1 calcium channel
Sterniczuk, R., Colijn, M. A., Nunez, M., & Antle, M. C. migraine mice. Annals of Neurology, 64, 315–324.
(2010). Investigating the role of substance P in photic Van Someren, E. J. (2000). Circadian and sleep disturbances
responses of the circadian system: Individual and combined in the elderly. Experimental Gerontology, 35, 1229–1237.
actions with gastrin-releasing peptide. Neuropharmacology, VanderLeest, H. T., Houben, T., Michel, S., Deboer, T.,
58, 277–285. Albus, H., Vansteensel, M. J., et al. (2007). Seasonal
Strecker, G. J., Wuarin, J. P., & Dudek, F. E. (1997). encoding by the circadian pacemaker of the SCN. Current
GABAA-mediated local synaptic pathways connect neurons Biology, 17, 468–473.
in the rat suprachiasmatic nucleus. Journal of Neurophysiol- VanderLeest, H. T., Rohling, J. H., Michel, S., & Meijer, J. H.
ogy, 78, 2217–2220. (2009). Phase shifting capacity of the circadian pacemaker
Strogatz, S. H. (2003). Sync: How order emerges from chaos in the determined by the SCN neuronal network organization.
universe, nature, and daily life. New York: Hyperion Books. PLoS One, 4, e4976.
162

Vansteensel, M. J., Yamazaki, S., Albus, H., Deboer, T., Wever, R. (1960). Possibilities of phase-control, demonstrated
Block, G. D., & Meijer, J. H. (2003). Dissociation between cir- by an electronic model. Cold Spring Harbor Symposia on
cadian Per1 and neuronal and behavioral rhythms following a Quantitative Biology, 25, 197–206.
shifted environmental cycle. Current Biology, 13, 1538–1542. Winfree, A. T. (1967). Biological rhythms and the behaviour
Vosko, A. M., Schroeder, A., Loh, D. H., & Colwell, C. S. of populations of coupled oscillators. Journal of Theoretical
(2007). Vasoactive intestinal peptide and the mammalian Biology, 16, 15–42.
circadian system. General and Comparative Endocrinology, Wu, L. G., Westenbroek, R. E., Borst, J. G. G.,
152, 165–175. Catterall, W. A., & Sakmann, B. (1999). Calcium channel
Vuillez, P., Jacob, N., Teclemariam-Mesbah, R., & Pevet, P. types with distinct presynaptic localization couple differen-
(1996). In Syrian and European hamsters, the duration of tially to transmitter release in single calyx-type synapses.
sensitive phase to light of the suprachiasmatic nuclei The Journal of Neuroscience, 19, 726–736.
depends on the photoperiod. Neuroscience Letters, 208, Wulff, K., Porcheret, K., Cussans, E., & Foster, R. G. (2009).
37–40. Sleep and circadian rhythm disturbances: Multiple genes
Wagner, S., Castel, M., Gainer, H., & Yarom, Y. (1997). and multiple phenotypes. Current Opinion in Genetics and
GABA in the mammalian suprachiasmatic nucleus and its Development, 19, 237–246.
role in diurnal rhythmicity. Nature, 387, 598–603. Yamaguchi, S., Isejima, H., Matsuo, T., Okura, R., Yagita, K.,
Wang, Y., Zhang, X. S., & Chen, L. (2009). A network biology Kobayashi, M., et al. (2003). Synchronization of cellular
study on circadian rhythm by integrating various omics data. clocks in the suprachiasmatic nucleus. Science, 302,
OMICS, 13, 313–324. 1408–1412.
Watanabe, A., Shibata, S., & Watanabe, S. (1995). Circadian Yamazaki, S., Kerbeshian, M. C., Hocker, C. G., Block, G. D.,
rhythm of spontaneous neuronal activity in the supra- & Menaker, M. (1998). Rhythmic properties of the hamster
chiasmatic nucleus of old hamster in vitro. Brain Research, suprachiasmatic nucleus in vivo. The Journal of Neurosci-
695, 237–239. ence, 18, 10709–10723.
Waterhouse, J., Reilly, T., Atkinson, G., & Edwards, B. Yamazaki, S., Numano, R., Abe, M., Hida, A., Takahashi, R.,
(2007). Jet lag: Trends and coping strategies. The Lancet, Ueda, M., et al. (2000). Resetting central and peripheral cir-
369, 1117–1129. cadian oscillators in transgenic rats. Science, 288, 682–685.
Webb, A. B., Angelo, N., Huettner, J. E., & Herzog, E. D. Yan, L., & Silver, R. (2002). Differential induction and locali-
(2009). Intrinsic, nondeterministic circadian rhythm genera- zation of mPer1 and mPer2 during advancing and delaying
tion in identified mammalian neurons. Proceedings of the phase shifts. The European Journal of Neuroscience, 16,
National Academy of Sciences of the United States of 1531–1540.
America, 106, 16493–16498. Yan, L., & Silver, R. (2004). Resetting the brain clock: Time
Welsh, D. K., Logothetis, D. E., Meister, M., & Reppert, S. M. course and localization of mPER1 and mPER2 protein
(1995). Individual neurons dissociated from rat expression in suprachiasmatic nuclei during phase shifts.
suprachiasmatic nucleus express independently phased cir- The European Journal of Neuroscience, 19, 1105–1109.
cadian firing rhythms. Neuron, 14, 697–706. Yan, L., Takekida, S., Shigeyoshi, Y., & Okamura, H. (1999).
Welsh, D. K., Takahashi, J. S., & Kay, S. A. (2010). Per1 and Per2 gene expression in the rat suprachiasmatic
Suprachiasmatic nucleus: Cell autonomy and network pro- nucleus: Circadian profile and the compartment-specific
perties. Annual Review of Physiology, 72, 551–577. response to light. Neuroscience, 94, 141–150.
Westenbroek, R. E., Sakurai, T., Elliott, E. M., Hell, J. W., Zhou, J. N., Hofman, M. A., & Swaab, D. F. (1995). VIP
Starr, T. V. B., Snutch, T. P., et al. (1995). Immunochemical neurons in the human SCN in relation to sex, age, and
identification and subcellular-distribution of the alpha(1a) Alzheimer’s disease. Neurobiology of Aging, 16, 571–576.
subunits of brain calcium channels. The Journal of Neurosci-
ence, 15, 6403–6418.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 10

Interaction of central and peripheral clocks


in physiological regulation

Johanna L. Barclay{, Anthony H. Tsang{ and Henrik Oster{,{,*

{
Circadian Rhythms Group, Max Planck Institute for Biophysical Chemistry, Göttingen, Germany
{
Medical Department I, University of Lübeck, Lübeck, Germany

Abstract: In mammals, circadian rhythms of physiology and behavior are regulated by a complex
network of cellular molecular oscillators distributed throughout the brain and peripheral tissues.
A master clock in the hypothalamic suprachiasmatic nuclei (SCN) synchronizes internal time with the
external light–dark cycle, thus entraining the overall rhythmicity of the organism. Recent findings have
challenged the dominant role of the SCN in physiological regulation and it becomes increasingly
evident that close interaction between different central and peripheral clocks is necessary to maintain
robust circadian rhythms of physiology and metabolism. In this review, we summarize recent findings
regarding circadian organization in the SCN and in other central and peripheral tissues. We outline
the communication pathways between different tissue clocks and, exemplified by the regulation of
glucocorticoid release from the adrenal gland and glucose homeostasis in the blood, characterize the
interaction between different clocks in the regulation of physiological processes.

Keywords: SCN; adrenal; circadian clock; glucose; metabolism; glucocorticoids; liver; clock genes;
mammals; pancreas.

Introduction its axis. The biological timing system that organizes


such 24-h oscillations is known as the circadian
Living organisms—from unicellular prokaryotes clock. The two major functions of the circadian clock
to multicellular metazoans—have evolved diverse are to optimize the temporal manifestations of
strategies to internalize the diurnal environmental different biological activities along the course
changes brought about by Earth’s rotation around of a day through anticipating recurring environ-
mental fluctuation and to separate incompatible
biological processes such as feeding and sleeping.
*Corresponding author.
Tel.: þ49-551-201-2738; Fax: þ49-551-201-2705
Unicellular organisms can internalize external
E-mail: henrik.oster@mpibpc.mpg.de diurnal rhythms by employing a single set of

http://dx.doi.org/10.1016/B978-0-444-59427-3.00030-7 163
164

molecular clockwork; multicellular organisms, 2 (NPAS2; McNamara et al., 2001; Reick et al.,
such as mammals, on the other hand, have devel- 2001). Period (PER1–3) and cryptochrome
oped a highly complex circadian timing system (CRY1–2) constitute the negative limb of the core
that is intricately intertwined with different physi- clock. CLOCK–BMAL1 complexes activate the
ological systems. Mammalian circadian rhythms expression of Per and Cry genes during the subjec-
are a result of a close interaction between differ- tive day. PER and CRY interact to form complexes
ent central (i.e., inside the central nervous system and translocate into the nucleus. When PER/CRY
or CNS) and peripheral oscillators. An ever- complexes accumulate to a critical concentration,
growing body of evidence has demonstrated they interact with CLOCK and BMAL1 and
that a misalignment of central and peripheral cir- thereby inhibit their transactivator function, thus
cadian clocks (internal desynchrony) predisposes shutting down Per and Cry transcription (Lee
individuals to physiological complications and dis- et al., 2001). The progressive degradation of PER/
eases. This chapter describes the current state of CRY complexes throughout the subjective night
knowledge concerning the properties of clocks in releases the inhibition on CLOCK–BMAL1 tran-
the CNS and the periphery. We then discuss gen- scriptional activity and, thereby, completes the neg-
eral mechanisms of central-to-peripheral clock ative feedback loop of the circadian clock.
interaction in the regulation of physiology using In addition to the core clock TTL described
two extensively studied processes, glucose metab- above, additional auxiliary TTLs have also been
olism and glucocorticoid (GC) secretion. described. Though principally dispensable, they sta-
bilize the oscillation of the core clock TTL and help
to translate time-of-day information into physiolog-
Molecular clockworks ical signals via transcriptional control of clock target
genes (Zhang and Kay, 2010). Such loops include
In the past decades, our knowledge of the molec- the nuclear receptors REV-ERBa (NR1D1) and
ular make-up of the cellular circadian clock has RORa (NR1F1) that regulate Bmal1 expression
been significantly advanced. The current model via retinoid orphan receptor-responsive elements
suggests that the central mechanism of the mam- (ROREs) (Preitner et al., 2002; Ueda et al., 2002)
malian molecular clock is composed of a set of as well as the PAR basic leucine zipper (bZIP)
clock genes intertwined with a delayed interlocked proteins, D-box albumin-binding protein (DBP)
transcriptional–translational feedback loop (TTL) and E4 promoter-binding protein (E4BP; NFIL3;
with several auxiliary mechanisms reinforcing Cowell, 2002; Ripperger and Schibler, 2006), that
robustness and stability (Zhang and Kay, 2010). feedback on the expression of Per genes via D-box
The positive limb of this TTL comprises the basic promoter elements (Ripperger et al., 2000).
helix–loop–helix transcription factors—circadian Recently, there are accumulating evidences
locomotor output cycles kaput (CLOCK) and brain showing that the circadian clock is tightly inter-
and muscle aryl hydrocarbon receptor nuclear twined with multiple metabolic pathways. While
translocator like (BMAL1 or ARNTL). They form the temporal manifestation of several metabolic
heterodimers via their PER–ARNT–SIM (PAS) functions is one of the most important functional
domains and activate E-box element containing outputs of the circadian clock (see below), the cel-
genes (Hardin, 2004; Zhang and Kay, 2010) by rec- lular circadian clock, on the other hand, constantly
ruiting several transcriptional co-activators, chro- receives the feedbacks from the metabolic sig-
matin modifying proteins, and finally, RNA nalings of the cells. For example, the redox status
polymerase II. In certain tissues such as forebrain of the cells has been demonstrated to regulate
or the vasculature, CLOCK is functionally replaced the molecular clock, in both direct and indirect
by its homologue neuronal PAS domain protein manners. The ratio of oxidized to reduced
165

nicotinamide adenine dinucleotide (phosphate) In both scenarios, AMPK activation promotes


(NAD(P)þ/NAD(P)H) reflects the cellular redox their degradation and thereby influences the core
and metabolic status. This ratio oscillates in a circa- clock oscillation cycle.
dian manner (Nakahata et al., 2009; Ramsey et al., These and other similar examples illustrate that
2009). The binding of the CLOCK–BMAL1 and the molecular clockwork and several metabolic sig-
NPAS2–BMAL1 heterodimeric complexes to the naling pathways are so tightly intertwined that
E-box elements is inhibited by the oxidized NADþ the clear distinction between them is becoming
and NADP but stimulated by the reduced NADH somewhat blurrier. Here, we only reviewed a few
and NADPH (Rutter et al., 2002). NAD is an examples of such metabolic feedback pathways
important cofactor that involves in many cellular on the circadian clock. Several elegant reviews dis-
enzymatic reactions, for example, sirtuins-mediated cussing the orchestration of the metabolic homeo-
protein deacetylation. SIRT1 catalyzes the removal stasis and the circadian clock and the deleterious
of acetyl group from acetylated lysine residues of consequences if this orchestration is disrupted have
proteins thereby modulating their activities in the recently been published (Asher and Schibler, 2011;
expense of NAD(P)H. It has recently been shown Huang et al., 2011; Reddy and O’Neill, 2010). Most
that the activity and level of SIRT1 oscillate in a cir- of these metabolic feedback mechanisms have
cadian manner (Asher et al., 2008; Nakahata et al., been implicated in the food entrainment of the
2008). Importantly, SIRT1 physically interacts with peripheral clock but have only a little if not no
the CLOCK–BMAL1 complex and mediates effect on the suprachiasmatic nuclei (SCN) clock.
BAML1 and histone H3 deacetylation, which is However, it would be interesting to know how
important for the transcriptional activating activity these pathways influence the extra-SCN clocks in
of the CLOCK–BMAL1 complex (Nakahata et al., the brain, particularly for those regions involved
2008). In addition to acting on the positive limb in organizing the feeding schedule.
components, SIRT1 also deacetylates PER2, which
promotes its degradation (Asher et al., 2008). More
recently, another NAD-dependent poly (ADP- The master circadian pacemaker of the SCN
ribose) polymerase-1 (PARP-1)-mediated protein
poly (ADP-ribosyl)ation also showed a circadian In vertebrates, almost all cells express clock genes.
oscillation pattern (Asher et al., 2010). PARP-1, Without synchronization on a systemic level, these
on the other hand, can also poly (ADP-ribosyl)ate autonomous clocks could not produce physiologi-
CLOCK and thereby inhibit the CLOCK–BMAL1 cally meaningful signals. In mammals, circadian
DNA-binding capacity (Asher et al., 2010). The regulation is organized in a hierarchical fashion
nutrient-sensing AMP-dependent protein kinase with the hypothalamic SCN housing the master
(AMPK) represents another elegant example circadian pacemaker. The SCN is a bilaterally
integrating the metabolism to the circadian clock. paired compact brain nucleus comprising about
AMP/ATP ratio is another indicator of the meta- 20,000 neurons in mice, located directly adjacent
bolic status of the cells. AMPK is the major sensor to the third ventricle and atop the optic chiasm
of such ratio. The elevated AMP level stimulates (Welsh et al., 2010). Electrical ablation of the
the AMPK activity via liver kinase B1 (LKB1; SCN renders animals behaviorally arrhythmic
Mihaylova and Shaw, 2011). The activation of (Moore and Eichler, 1972; Stephan and Zucker,
AMPK acts on the negative limb of the core clock 1972). Transplanting SCN tissue to SCN-lesioned
TTL in both direct and indirect manners. AMPK animals restores circadian rhythmicity (Ralph
directly phosphorylates CRY1 (Lamia et al., 2009) et al., 1990). Importantly, the restored rhythm of
and indirectly leads to the phosphorylation of recipients is determined by the donor SCN period,
PER2 via casein kinase 1e (CK1e) (Um et al., 2007). indicating that the SCN indeed generates the
166

timing information that synchronizes oscillators neurons (Aton et al., 2005; Buhr et al., 2010; Liu
throughout the body (Ralph et al., 1990). SCN et al., 2007; Welsh et al., 1995).
explants as well as dispersed neurons display The SCN innervates numerous brain nuclei,
robust circadian oscillations in firing rate in vitro, thereby passing time information to other CNS
suggesting that the rhythmicity of the SCN is clocks. The paraventricular hypothalamic nucleus
autonomous in nature (Green and Gillette, 1982; (PVN) is one of the main regions transducing SCN
Groos and Hendriks, 1982; Shibata et al., 1982; circadian function to the periphery (Saeb-Parsy
Welsh et al., 1995). et al., 2000). The PVN is a relay hub for energy
The endogenous circadian clock in mammals homeostasis and projects predominantly to the pitu-
possesses a rhythm with an approximate 24-h itary where it regulates the release of hormones
free-running period. However, the daily fluctua- such as adrenocorticotrophin (ACTH; see below)
tion of the external environment is not constant, and thyroid-stimulating hormone. The PVN also
and variables such as photoperiod, temperature, innervates the sympathetic limb of the autonomous
and food availability are subject to seasonal nervous system which allows the SCN to indirectly
changes. In order to synchronize the internal circa- control melatonin release from the pineal gland
dian rhythm to the external diurnal fluctuation (Buijs et al., 2003b). Further projections of the
patterns, circadian clocks are constantly reset by SCN have been described to the dorsomedial hypo-
external environmental cues, so-called zeitgebers, thalamic nucleus (DMH; Luiten et al., 1987), the
every day in a process known as entrainment. The nucleus accumbens (NAc; Phillipson and Griffiths,
major zeitgeber in mammals is light, with the SCN 1985), and the paraventricular thalamic nucleus
acting as a relay between the external light–dark (Watts and Swanson, 1987; Watts et al., 1987)
cycle and the endogenous timing system (Hankins enabling the SCN to affect a plethora of physiologi-
et al., 2008). Other, nonphotic zeitgebers exist, cal processes such as the reward system, fee-
some of which act through the SCN, for example, ding–fasting cycles, cognitive function, locomotor
arousal (Welsh et al., 2010), while others, such as activity, and body temperature (Dibner et al.,
food intake, are more directly affecting the periph- 2010). In addition, the SCN secretes diffusible
eral circadian machinery (Huang et al., 2011). factors which can function as timing cues. This
The SCN makes use of the same TTL molecu- notion has been substantiated by an elegant experi-
lar timekeeping machinery as the peripheral ment showing that membrane-encapsulated fetal
oscillators (see below). However, the particular SCN tissue grafts, which only allowed for small mol-
robustness and resilience of SCN circadian rhyth- ecule passage, could restore the periodicity of loco-
micity are achieved through the formation of a motor activity in SCN-lesioned hamsters in the
tight interneuronal network (Welsh et al., 2010). absence of axonal outgrowth (Silver et al., 1996).
SCN slices cultured in vitro exhibit robust and Transforming growth factor (TGF) alpha (Kramer
persistent circadian oscillations in electrophysio- et al., 2001; Li et al., 2002), prokineticin-2 (PK-2)
logical activity and clock gene expression for sev- (Cheng et al., 2002), and cardiotrophin-like cyto-
eral weeks, while rhythms in slice explants from kine (CLC) (Kraves and Weitz, 2006) have been
most other brain regions and peripheral tissues implicated as SCN-secreted peptides capable of
dampen after a couple of days (Guilding and regulating behavioral rhythmicity.
Piggins, 2007; Guilding et al., 2009). SCN explant
rhythms are also more resistant to temperature
fluctuations or clock gene mutations (Abraham Extra-SCN clocks in the brain
et al., 2010; Buhr et al., 2010; Liu et al., 2007).
This rigidity and robustness has been attributed The classical view of the role of SCN as the exclu-
to the intercellular coupling of individual SCN sive circadian pacemaker that controls all circadian
167

aspects of behavior (Fig. 1a) has been changing existence of other circadian clocks in the brain.
(Guilding and Piggins, 2007). The emerging the- Here, we discuss some examples of these extra-
ory is that the SCN synchronizes and coordinates SCN neural oscillators.
numerous semiautonomous circadian clocks The retina was the first neuronal tissue outside
residing in different brain regions and peripheral the SCN shown to possess a circadian oscillator.
tissues (Fig. 1b). Thanks to recent technical Cultured hamster retinae show a circadian pat-
advancements, long-lasting and self-sustained tern of melatonin synthesis in vitro (Tosini and
circadian oscillations of clock genes have been Menaker, 1996). Importantly, this rhythm persists
revealed in a number of brain nuclei in vitro in constant darkness (self-sustainment) but can be
(Abe et al., 2002; Guilding and Piggins, 2007). reset by a light–dark cycle (entrainment). More-
These data provide compelling evidence for the over, the period of the retina clock is resistant to

A B

SCN SCN

Peripheral clocks Systemic rhythms Peripheral clocks Behavioral rhythms


Leptin ng/mL

ZT

Peripheral transcriptional Peripheral transcriptional


rhythms rhythms

Physiological rhythms Physiological rhythms

Fig. 1. Communication routes of the mammalian circadian timing system. (a) The SCN pacemaker entrains peripheral clocks.
Via transcriptional regulation of clock controlled genes in target tissues, peripheral physiological functions are reset and
synchronized to the light–dark cycle. (b) Pathways of interaction between central and peripheral clocks. The SCN resets
physiological rhythms via the entrainment of peripheral tissue clocks, but at the same time regulates behavioral and systemic
(e.g., endocrine) functions in a more direct manner. Peripheral rhythms may, in turn, directly or indirectly feedback on the SCN.
This interlocked balance system creates plasticity in the entrainment of the circadian timing system and promotes adaptation to
complex changes in environmental parameters.
168

temperature changes (temperature compensation), (Guilding et al., 2009). A subpopulation of dopami-


thereby fulfilling all the formal requirements of nergic (DA) ARC neurons displays robust circadian
a true autonomous circadian oscillator (Tosini oscillations of Per1 and Per2 transcripts (Sellix et al.,
and Menaker, 1996). More recently, circadian 2006). These neurons receive projections from
oscillations of clock gene expression have also the SCN (Gerhold et al., 2001) and themselves proj-
been detected in the retina, further supporting ect to the anterior pituitary, where they regulate
the clock properties of this tissue (Kamphuis prolactin secretion. The DA turnover in these
et al., 2005). Genetic disruption of retinal clock neurons exhibits a circadian rhythm which is light
function results in the loss of circadian rhythm entrainable and has an approximate 24-h free-
of electrical responses to light (Storch et al., running period in constant conditions (Sellix and
2007). The olfactory bulbs (OBs), similar to the Freeman, 2003). It has been reported that several
retina, have also been shown to comprise an ARC neuropeptides display circadian/diurnal
autonomous circadian clock (Granados-Fuentes expression patterns including Agouti-related protein
et al., 2004a,b). OB Per1 expression and electro- (AgRP; Lu et al., 2002), cocaine and amphetamine-
physiological activity display rhythmic circadian regulated transcript (CART; Vicentic et al., 2005),
patterns. Importantly, these rhythms persist in neuropeptide Y (NPY), and pro-opiomelanocortin
SCN-lesioned animals and under constant light (POMC; Kalra et al., 1999). Also, orexin, secreted
conditions—when behavioral rhythms are dis- from the lateral hypothalamic area (LHA), shows
rupted—indicating that the OB’s circadian oscilla- rhythmic expression (Willie et al., 2001). Impor-
tion is autonomous and independent from the SCN tantly, many of these rhythms are altered in
(Granados-Fuentes et al., 2004a). Circadian rhythms diet-induced obese animals (Kohsaka et al., 2007),
of clock gene expression and of odor responses suggesting a link between the hypothalamic clock
have been described in the OB (Granados-Fuentes and metabolic regulation.
et al., 2006), but molecular evidence for an autono- The reward system controls and regulates ani-
mous OB clock is still missing. mal behavior by inducing pleasurable sensations
The mediobasal hypothalamus (MBH), situated upon perceiving certain stimuli. Such stimuli can
posterior to the optic chiasm and overlying the be primary, that is, in response to food, sex, and
pituitary, is a collective anatomical structure com- water, which are important for the survival of
prising the DMH, ventromedial (VMH), PVN, individuals and propagation of species, or second-
supraoptic, and arcuate (ARC) nuclei. The MBH ary, for example, in the context of money/profits
is deemed as an integrating center regulating a and power/reputation (Chen et al., 2010). The
diverse array of physiological processes such as major anatomical structure of reward is found in
growth, feeding, maturation, and reproduction the mesolimbic DA system. DA neurons from
(Luiten et al., 1987). Clock gene rhythms have the ventral tegmental area in the midbrain signal
been demonstrated in the DMH (see below), the to the NAc, the dorsal striatum, and the frontal
PVN, and the ARC, as well as in the adjacent cortex (Chen et al., 2010). The links between
median eminence and the pituitary (Abe et al., circadian rhythms and the reward system are
2002; Guilding et al., 2009). However, the physio- multifaceted (Dibner et al., 2010; Webb et al.,
logical relevance of potential MBH clocks remains 2009). Patients that suffer from defective reward
largely unknown. A recent study revealed that, function, for example, in bipolar disorder or major
despite the fact that ARC slices in culture possess depression, often show altered behavioral rhythms
a sustained (more than 1 week) molecular rhythm, and sleep patterns (Westrich and Sprouse, 2010).
electrical firing rate rhythms dampen within 48 h, Conversely, patients with genetic sleep disorders
suggesting that in the MBH molecular clocks may are often predisposed to addiction (Shibley
not—or only weakly—couple to electrical activity et al., 2008). The most striking example, seasonal
169

affective disorder (SAD), provides a mechanistic shown that the lost rhythmic locomotor activity
link between altered mood status and altered circa- of SCN-lesioned arrhythmic rats can be partly
dian behavior. SAD patients suffer from depres- restored by temporally restricted feeding (RF)
sive episodes upon seasonal change, mostly in schedules (Stephan et al., 1979), a phenomenon
winter, suggesting that alterations in circadian light known as food anticipatory activity (FAA; Boulos
entrainment might trigger disease states (Levitan, and Terman, 1980; Mistlberger, 1994). RF can
2007). Along the same line, psychostimulants also restore the rhythmicity of pineal melatonin
such as cocaine and methamphetamine that can release (Feillet et al., 2008), thermogenesis,
activate the mesolimbic reward system and induce plasma rhythms of nutrient-related blood-borne
pleasurable effects are known to affect circadian hormones, and drinking patterns (Boulos and
clock. Early studies revealed that chronic exposure Terman, 1980; Mistlberger, 1994; Stephan, 2002).
to methamphetamine can disrupt the circadian Such feeding-related rhythms, once established,
rhythms of rats (Honma et al., 1986). On the other can persist (or free-run) for several days with an
hand, many aspects of addictive behavior show approximate 24-h period under fasting conditions.
a time-of-day-dependent pattern with a period Lesion studies have tried to determine the
of approximately 24 h (Abarca et al., 2002). anatomical locus of the FEO, but to date, no
Drug intake can induce clock gene expression in study has unequivocally identified a structure
several brain areas including the NAc and the stri- essential for the generation of feeding-related
atum (Uz et al., 2005; Yuferov et al., 2003). In rhythms (Davidson, 2009). Moreover, it remains
rodents, clock genes have been shown to modify controversial whether the known clock genes
psychostimulant responses (Abarca et al., 2002; are involved in the regulation of FAA (Challet
Rosenwasser et al., 2005; Spanagel et al., 2005). et al., 2009; Feillet et al., 2006; Storch and Weitz,
Importantly, ClockD19 mutant mice show increased 2009). The current prevailing opinion is that the
excitability of DA neurons and a higher rate of FEO is a diffuse system emerging from the inter-
DA synthesis, indicating a general excited, mania- play of different circuits within the CNS, and
like state of the DA circuitry (McClung et al., likely even within peripheral organs. Within the
2005; Roybal et al., 2007). Recently, Per2 has been CNS it may comprise multiple brain regions as
identified as a positive regulator of monoamine well as various signaling pathways.
oxidase A (MAOA) expression and activity, a The concept of the methamphetamine-sensitive
degrading enzyme of DA, and hence a negative oscillator (MASCO) originated in the 1980s;
regulator or DA release (Hampp et al., 2008). Honma and colleagues showed that chronic treat-
Taken together, these data illustrate an active ment of SCN-lesioned rats with methamphetamine
role of the circadian clock in modulating the could reinitiate circadian locomotor activity, core
mesolimbic reward pathway. It remains to be body temperature, and plasma corticosterone
shown, however, whether the DA neurons them- rhythms (Honma et al., 1987, 1988), which can per-
selves contain a functional circadian clock and sist for up to 2 weeks after withdrawal from the
how such a system affects reward responses to dif- drug (Ruis et al., 1990). The emergence of meth-
ferent stimuli. amphetamine-induced rhythms is independent of
Two phenomena tightly linked to the reward functional circadian clock machinery (Honma
system suggest that circadian rhythms may also et al., 2008; Masubuchi et al., 2001). Interestingly,
emerge from structures that are potentially very methamphetamine treatment can reset the rhythm
differently organized than the cellular TTL clocks of clock gene expression in several brain areas such
described above. The food-entrainable oscillator as the caudate putamen, the striatum, and the pari-
or FEO is a putative timing system that has drawn etal cortex, but not the SCN (Masubuchi et al.,
much attention in the past 30 years. It has been 2000). The anatomical structure, the endogenous
170

zeitgebers, and the output pathways of the Kalsbeek et al., 2004; Oster et al., 2006a), while
MASCO, however, remain largely unknown. others are affected by behavior-associated changes
in temperature (Brown et al., 2002; Dibner et al.,
2010). Circadian transcriptome profiling studies
Peripheral clocks suggest that local peripheral clocks, while being
reset by the SCN, independently control tissue
Outside the brain, “canonical” circadian clocks physiology via the regulation of output genes com-
have been identified in several peripheral organs prising 5–10% of the active transcriptome (Akhtar
and tissues, capable of generating oscillations with et al., 2002; Hughes et al., 2009; Kornmann
a periodicity of approximately 24 h. The intrinsic et al., 2007a; McCarthy et al., 2007). In line with
properties of such peripheral clocks have been this, peripheral clocks have been implicated in a
characterized using primary cell culture models plethora of physiological functions such as cardiac
and tissue explants cultures (Yoo et al., 2004). contraction (Bray and Young, 2009), renal excre-
Immortalized rat fibroblast (Rat-1) cells display tion (Firsov et al., 2011), adipogenesis and lipid
robust oscillations of clock gene expression after metabolism (Gimble et al., 2011), digestive pro-
brief stimulation with high concentrations of serum cesses (Gimble and Floyd, 2011), and xenobiotic
(serum shock; Balsalobre et al., 1998). Using single metabolism (Claudel et al., 2007), to name but a
cell imaging techniques, Nagoshi et al. showed that few. However, recent data suggest that the organi-
individual fibroblasts possess sustained endoge- zation of circadian molecular and physiological
nous circadian expression rhythms of clock genes, functions is more complex than originally thought
although populations of cells quickly become and involves tight interaction between different
desynchronized from each other because of indi- central and peripheral clocks (Kornmann et al.,
vidual differences in period (Nagoshi et al., 2004). 2007b; Fig. 1b). In the following section, we will
Serum shock (or stimulation with forskolin, discuss glucose metabolism and the regulation of
GCs, or phorbol esters) synchronizes the individ- GC secretion from the adrenal cortex to exemplify
ual cells, yielding a transiently phase coherent pop- the intricate interplay between the central and
ulation (Nagoshi et al., 2004). These results point peripheral circadian clocks which is essential for
to the notion that the peripheral cellular clock is the maintenance of physiological homeostasis.
actually self-sustained and autonomous in nature
but fails to maintain the coherence with neighbor-
ing cells, in contrast to the coupled nature of SCN Glucocorticoid (GC) secretion
(see above). In other words, at the cellular level,
fibroblasts do not differ from SCN neurons in GCs are steroids produced in the adrenal gland, cor-
terms of the molecular circadian machinery (Liu tisol in humans and corticosterone in rodents, which
et al., 2007). In line with this observation, tissue are essentially involved in energy metabolism,
explants from a wide array of peripheral organs immune function, and stress responses. Disruption
including heart, lung, kidney, liver, spleen, pan- of GC secretion is associated with severe patho-
creas, stomach, cornea, thyroid gland, and adrenal physiology. Patients affected with Cushing’s Syn-
gland all show robust clock gene expression drome, characterized by excess GC levels, often
rhythms (Yamazaki et al., 2000; Yoo et al., 2004). present with diabetes mellitus, osteoporosis,
It is still not fully understood how the SCN tra- hypertension, dyslipidemia, and sleep disorders
nsmits its timing signal to peripheral clocks. Endo- (Carroll and Findling, 2010). Conversely, GC
crine signals such as GCs play a role (Balsalobre insufficiency results in Addison’s disease, which is
et al., 2000; Kiessling et al., 2010). Some clocks characterized by stress sensitivity, hypoglycemia,
respond to neuronal cues (Ishida et al., 2005; hypotension, mood disturbances, and weight loss
171

(Anglin et al., 2006). In addition, chronic fatigue receptors (Chung et al., 2011a). Rhythmic GC
syndrome is associated with perturbations in GC secretion is regulated by an intrinsic circadian clock
regulation (Chung et al., 2011b). Clock-related life- located in the cortex of the adrenal gland (Oster
style factors can affect GC levels, such as shift et al., 2006a,b; Son et al., 2008). Using a model of
work, jet lag, and nighttime eating, and are adrenal transplantation between wild-type and
associated with diabetes and metabolic syndrome mutant mice lacking a functional clock, the role of
and increased risk of heart attack and cancer the adrenal clock was elucidated (Oster et al.,
(Boivin et al., 2007). GC secretion shows a very 2006b). In the absence of an SCN clock, GC
prominent diurnal/circadian rhythm peaking rhythms remain entrainable by light, but in the
around wake-up time (morning in humans, even- absence of light, the rhythm is rapidly lost. Con-
ing in rodents). SCN and adrenal clocks are both versely, in the absence of a functional adrenal clock,
required for the circadian production of GCs. Fur- GC rhythms are dampened, suggesting that the
ther, GC actions are circadian gated through the function of the adrenal clock is to gate the
rhythmic expression of its receptors. Finally, GC responsiveness of the adrenal to ACTH through
feedback directly affects the phase of clock gene the rhythmic expression of steriodogenic enzymes
transcription in other peripheral oscillators, com- or modulators. A Bmal1 knockdown study supports
pleting an elegant cycle of integration. these findings, suggesting that the adrenal clock
The nycthemeral production of GCs occurs plays a dominant role in the regulation of local GC
through the hypothalamus/pituitary/adrenal (HPA) production in the adrenal, though not of circulating
axis. The role of the SCN was determined in GC levels in the blood (Son et al., 2008).
the 1970s and 1980s, with studies demonstrating In addition to HPA axis control, GC release is
that the ACTH release from the pituitary is no lon- regulated by neuronal signals. Virus tracing stud-
ger rhythmic in SCN-lesioned animals, disrupting ies reveal multisynaptic autonomic connections
adrenal GC secretion rhythms (Cascio et al., 1987; between the SCN and the adrenal gland (Buijs
Moore and Eichler, 1972; Szafarczyk et al., 1983). et al., 1999). Jasper et al. showed that splanchnic
The neuropeptide arginine vasopressin (AVP) is denervation results in dampening of diurnal GC
released rhythmically from SCN neurons (Earnest rhythms and increased sensitivity to ACTH stimu-
and Sladek, 1986; Gillette and Reppert, 1987) lation (Jasper and Engeland, 1997). A direct effect
projecting into the rostral PVN. There, it inhibits of light on the adrenal gland was characterized by
the production of corticotrophin-releasing hormone Ishida et al., who showed that light exposure
(CRH; Buijs et al., 1993; Gomez et al., 1997; induces Per gene expression in the adrenal gland
Kalsbeek et al., 1992), which is ultimately responsible via the SCN sympathetic nervous system, resulting
for ACTH release from hypophyseal adreno- in an upregulation of GC release (Ishida et al.,
corticotrophs in the pituitary. In this manner, circa- 2005), thus offering a mechanism for the observed
dian rhythms in the SCN result in the circadian light entrainment of adrenal clock transplants in
release of ACTH, peaking at the beginning of the otherwise arrhythmic animals mentioned above
active phase. ACTH then regulates the production (Oster et al., 2006b).
and release of GC from the adrenal gland. GC receptors are expressed throughout the
Anatomically, the adrenal gland is divided into periphery and the brain, with the notable excep-
the cortex and the medulla, which are structurally tion of the SCN (Rosenfeld et al., 1988). Activated
and functionally discrete. The medulla is responsi- GC receptors act as transcription factors via activa-
ble for the secretion of epinephrine and norepi- tion or repression of GC target genes (Surjit et al.,
nephrine, while the cortex produces various 2011). Disruption of GC signaling, for example, by
steroid hormones. GCs are produced in the zona adrenalectomy, affects gene transcription in the
fasciculata cells of the cortex, which express ACTH periphery and the brain. In the central nucleus of
172

the amygdala, it causes a loss of PER2 rhythmic hepatic glucose output, while parasympathetic
expression, while in the liver, it alters the regulation input stimulates insulin-dependent glucose uptake
of numerous genes involved in metabolism (Lamont and storage in the form of glycogen (Puschel,
et al., 2005; Oishi et al., 2005). Timed GC or GC ana- 2004). The influence of the master clock on glu-
log treatment in mice has a powerful resetting effect cose regulation was shown by Buijs and colleagues
on liver (clock) gene rhythms (Balsalobre et al., using SCN-lesioned rats, which have no diurnal
2000; Reddy et al., 2007; Segall et al., 2006; Son glucose rhythm, compared to fasted and arrhyth-
et al., 2008). Additionally, GC receptors are rhyth- mic-fed (six daily feeds) animals, which retain their
mically transcribed in various tissues and subjected glucose rhythms (La Fleur et al., 1999). The pre-
to acetylation—and subsequent inactivation—by cise nature of this regulation has been dissected
CLOCK (Nader et al., 2009; Yao et al., 2006). In in a series of studies using euglycemic hyper-
conclusion, GC production, secretion, and signaling insulinemic clamps in combination with selective
are an example of the complex system of integration autonomic denervation of the liver. Initial retro-
afforded by the presence of multiple circadian grade tracing studies from the liver revealed
clocks organized in a hierarchical manner. projections via both sympathetic and parasympa-
thetic systems to third-order neurons in the hypo-
thalamus, specifically the SCN (La Fleur et al.,
Glucose metabolism 2000). Studies designed to distinguish between
sympathetic and parasympathetic output from the
The maintenance of glucose homeostasis is essential SCN showed that exclusive populations of neurons
for mammalian physiology. Plasma glucose levels within the SCN are responsible for each of these
display diurnal rhythms in mammals, peaking signals, projecting to preautonomic neurons in the
before the onset of activity while remaining constant PVN (Buijs et al., 2003a). In 2004, it was shown
throughout the remainder of the day. This peak that the SCN-derived sympathetic inputs to the
does not coincide with food intake, clearly PVN were inhibitory GABAergic inputs, and their
illustrating the extent of endogenous regulation inhibition resulted in increased hepatic glucose
dedicated to glucose circulation (La Fleur et al., production (Kalsbeek et al., 2004). It was addition-
1999). Circulating glucose is altered by absorption ally demonstrated that these GABAergic inputs
from the gut following feeding, glucose uptake into provide the circadian timing information for the
tissues, and glucose production. These latter pro- liver as well as for the insulin response of the pan-
cesses are tightly regulated in a temporal manner creas (Kalsbeek et al., 2008). Interestingly, it was
to assure sufficient glucose availability, for example, shown that complete denervation of the liver in
for the brain, while avoiding extended postprandial conjunction with constant feeding does not abolish
hyperglycemia. The liver plays a pivotal role in this diurnal glucose rhythms; however, this can be
process as a site of glucose uptake from the circula- achieved by inactivation of either the sympathetic
tion, as well as being the major source of de novo- or the parasympathetic inputs (Cailotto et al.,
synthesized glucose in times of need (Kalsbeek 2008). Collectively, these studies indicate that the
et al., 2010). The regulation of the diurnal glucose autonomic modulation of glucose rhythms requires
rhythm has been shown to be maintained by a balance in both branches of the autonomic ner-
both neuroendocrine and neuronal pathways and vous system by the SCN.
involve a large number of different central and Further studies revealed that orexin, a hypo-
peripheral circadian clocks. thalamic neuropeptide involved in wakefulness
Early studies on the autonomic innervation of and feeding behavior, is an important regulator
the liver in regard with glucose homeostasis show of glucose homeostasis and the main effector in
that sympathetic input predominantly increases the preactive phase glucose peak (Yi et al., 2009).
173

Intracerebroventricular infusion of orexin results Microarray analyses of the ClockD19 mutant liver
in increased glucose production. Inhibition of transcriptome reveal that metabolic gene rhyth-
GABAergic inputs—originating from the SCN— micity is dampened (Oishi et al., 2003). ClockD19
to the perifornical orexin area (PF-Oa) has a mutant and Bmal1/ mice show impaired
similar effect, correlating with activation of gluconeogenic potential, correlating with dec-
orexin-positive neurons. Given that GABAergic reased phosphoenolpyruvate carboxykinase 1
inhibition of hyperglycemia, GABAergic inputs (PEPCK) expression in the liver (Rudic et al.,
to orexin neurons in the PF-Oa, and orexin 2004). More recently, studies have shown that
release all show clear diurnal rhythmicity (Alam CRY1 negatively regulates gluconeogenesis
et al., 2005; Kalsbeek et al., 2008; Zhang et al., through the inhibition of G-protein-coupled recep-
2004), it is feasible to propose that orexin- tor-mediated cAMP accumulation (Zhang et al.,
containing PF-Oa neurons translate SCN-derived 2010). Perhaps most significantly, liver-specific
GABAergic rhythms into glucose rhythms via Bmal1/ (L-Baml1/) mice display per-
the sympathetic nervous system. turbations in rhythmic expression of glucose regu-
Interestingly, hepatic sympathetic denervation latory genes and glucose metabolism, including
of the liver results in the loss of diurnal glucose circulating blood glucose (Lamia et al., 2008).
rhythms, without affecting gene expression rhythms These mice display hypoglycemia in the middle
of liver clock genes, indicating that the liver clock is and end of the inactive phase and increased glucose
not essential in this process (Cailotto et al., 2005). uptake during this time, arguing for a significant
However, glucose production is generally thought role of hepatic clocks in the maintenance of glucose
to be further regulated via clock target genes homeostasis.
involved in glucose metabolism in the liver. The Glucose uptake is dependent on insulin and can
liver clock has been well described, and rhythmic be influenced by insulin release from the pancreas
liver genes are highly enriched for metabolic func- or by insulin sensitivity in peripheral tissues. Insu-
tion (Lamia et al., 2008; Oishi et al., 2003; Panda lin levels display diurnal rhythms, and similarly to
et al., 2002; Storch et al., 2002). SCN ablation studies glucose, time-RF results in changes to this rhythm
show that liver clock gene and clock output gene in rodents (Diaz-Munoz et al., 2000). In the case
rhythmicity are abolished or severely dampened in of insulin, however, the diurnal rhythm is lost fol-
the absence of synchronization by the central clock lowing fasting and appears to be primarily depen-
(Akhtar et al., 2002). The liver clock is highly dent on feeding rhythms rather than being clock
responsive to RF regimes, which change the phase driven (La Fleur et al., 1999). On the other hand,
of core clock gene expression as well as the diurnal feeding-stimulated insulin responses are rhythmic
rhythm of circulating glucose (Damiola et al., 2000; (Kalsbeek and Strubbe, 1998). Indeed, changes in
Escobar et al., 1998). Interestingly, these RF- glucose uptake over the course of the day correlate
induced changes to the liver clock are inhibited to alterations in insulin sensitivity rather than insu-
by GCs (Le Minh et al., 2001). Conversely, regular lin release, and this rhythm is lost in SCN-lesioned
short-period feeding paradigms prevent dis- animals, indicating central clock involvement (La
turbances in clock gene expression in the liver and Fleur et al., 2001). Conflictingly, SCN-lesioned
leave circulating glucose rhythms intact (La Fleur animals display increased glucose uptake as well
et al., 1999). as decreased meal-induced insulin secretion. This
Many clock-deficient animal models show per- is hypothesized to be due to insulin-independent
turbations in glucose metabolism. Clock mutant glucose uptake, perhaps via autonomic inputs
mice display impaired glycogen storage, which cor- to muscle tissues (La Fleur, 2003). Functional
relates with dampened glycogen synthase 2 (Gys2) clocks have been identified in the pancreatic b-cells
expression rhythms in the liver (Doi et al., 2010). (Marcheva et al., 2010; Sadacca et al., 2011).
174

In ClockD19 mutant mice, a diabetic phenotype com- models lacking a functional clock (Turek et al.,
prising increased blood glucose levels, loss of active 2005). Leptin signals predominantly in the ARC,
phase insulin peak, reduced glucose tolerance, and which contains two distinct populations of neurons:
retarded glucose-stimulated insulin release is seen the first expressing NPY and AgRP and possessing
(Marcheva et al., 2010). Generation of a b-cell- orexigenic function and the second expressing
specific Bmal1/ mouse confirmed that these POMC and CART and possessing anorexigenic
phenotypes are due to the loss of a functioning clock function. In addition to receiving leptin signals,
in islets of Langerhans and seem to be the result of these cells may also be directly responsive to glu-
impaired glucose-stimulated insulin release rather cose and function as sensors of peripheral glucose
than insulin production (Sadacca et al., 2011). fluctuations. Some studies have demonstrated that
As discussed above, euglycemia is maintained in NPY neurons are inhibited by glucose, whereas
the fasting state predominantly through hepatic pro- POMC neurons are excited, although it should be
duction and release of glucose. Of note, GCs can noted that other studies have failed to see these
increase gluconeogenesis directly through the acti- effects (Claret et al., 2007; Fioramonti et al., 2007;
vation of a number of key enzymes involved in this Ibrahim et al., 2003; Muroya et al., 1999). Despite
pathway (Jin et al., 2004; Sasaki et al., 1984; Vander discrepancies, it seems plausible that ARC neurons
Kooi et al., 2005) as well as through the production are capable of responding directly to changes in
of suitable substrates from increased lipolysis circulating glucose levels. ARC neurons signal to
(Campbell et al., 2011), thus suggesting indirect various regions of the brain such as the LHA, the
effects of adrenal—and possibly adipose—clocks VMH, the mediodorsal nucleus of the thalamus,
in glucose regulation. GCs further affect gluconeo- the dentate gyrus, the piriform cortex, the ventral
genesis by decreasing the sensitivity of the liver to basolateral amygdala, and the bed nucleus of stria
insulin, the major inhibitor of gluconeogenesis. This terminalis (DeFalco et al., 2001; Muroya et al.,
is achieved in two ways, first, by directly decreasing 2004), and some of these neurons may also possess
the release of insulin from b-cells in the pancreas glucose-sensing properties. For example, orexin/
and, second, by decreasing the insulin-mediated glu- hypocretin-containing neurons of the LHA are
cose uptake in adipocytes and muscle (Delaunay activated by hypoglycemia (Cai et al., 2001) and
et al., 1997; Sakoda et al., 2000; Weinstein et al., inhibited by glucose in mice (Guyon et al., 2009).
1998), potentially involving further circadian clocks Melanin-concentrating hormone (MCH)-containing
active in the respective tissues. Under conditions of neurons—also in the LHA—respond to glucose
HPA axis dysregulation where GC rhythms are stimulation (Burdakov et al., 2005). Given the
affected, abnormal glucose homeostasis is observed, apparent opposing functions of these subsets of
promoting the development of diabetes. neurons, with orexin/hypocretin neurons involved
The hypothalamus contains distinct populations in wakefulness and MCH-containing neurons
of neurons, identifiable according to their signa- involved in sleep and decreased activity, circulating
ture expression profiles of receptors and neuro- glucose levels can feedback to the hypothalamus
transmitters. These nuclei form a complex to direct appropriate behaviors. How clock disrup-
network of excitatory and inhibitory signals that tion in these neurons affects metabolic homeostasis
regulate peripheral nutritional status and both and glucose levels, however, remains to be shown.
homeostatic and hedonic feeding. These feeding
nuclei receive inputs from the SCN, as discussed
above, and many have been suggested to contain Conclusion
functional molecular clocks (Guilding et al.,
2009). Arguably the most important input in the While the circadian clock is traditionally seen as a
regulation of feeding is leptin, which shows a diur- top-down-controlled system in which the SCN pace-
nal circulation rhythm, and is disrupted in mouse maker synchronizes peripheral clocks throughout
175

the body, which, in turn, regulate local physiolog- phenomenon. The Journal of Neuropsychiatry and Clinical
ical rhythms via transcriptional programs (Fig. 1a), Neurosciences, 18(4), 450–459.
Asher, G., Gatfield, D., et al. (2008). SIRT1 regulates circa-
recent data clearly suggest that a coordinated dian clock gene expression through PER2 deacetylation.
interplay between different central and peripheral Cell, 134(2), 317–328.
clocks is necessary to maintain robust rhythms of Asher, G., Reinke, H., et al. (2010). Poly(ADP-ribose) poly-
certain humoral factors (e.g., GCs) or restrict merase 1 participates in the phase entrainment of circadian
fluctuations in others within physiologically toler- clocks to feeding. Cell, 142(6), 943–953.
Asher, G., & Schibler, U. (2011). Crosstalk between compo-
able ranges (e.g., blood glucose; Fig. 1b). We are nents of circadian and metabolic cycles in mammals. Cell
only now developing the genetic tools to dissect Metabolism, 13(2), 125–137.
the role of organ-specific circadian clocks in these Aton, S. J., Colwell, C. S., et al. (2005). Vasoactive intestinal poly-
processes in living animals, and new technologies peptide mediates circadian rhythmicity and synchrony in mam-
will be needed to specifically manipulate several malian clock neurons. Nature Neuroscience, 8(4), 476–483.
Balsalobre, A., Brown, S. A., et al. (2000). Resetting of circa-
tissue clocks in a coordinated manner which would dian time in peripheral tissues by glucocorticoid signaling.
allow the interactivity of the circadian oscillatory Science, 289(5488), 2344–2347.
network at the systemic level to be analyzed. Balsalobre, A., Damiola, F., et al. (1998). A serum shock
induces circadian gene expression in mammalian tissue cul-
ture cells. Cell, 93(6), 929–937.
Acknowledgments Boivin, D. B., Tremblay, G. M., et al. (2007). Working on
atypical schedules. Sleep Medicine, 8(6), 578–589.
Boulos, Z., & Terman, M. (1980). Food availability and daily
H. O. is an Emmy Noether Fellow of the German biological rhythms. Neuroscience and Biobehavioral
Research Foundation (DFG) and a Lichtenberg Reviews, 4(2), 119–131.
Fellow of the Volkswagen Foundation. A. H. T. Bray, M. S., & Young, M. E. (2009). The role of cell-specific
is supported by a GGNB Fellowship of the circadian clocks in metabolism and disease. Obesity Reviews,
10(Suppl. 2), 6–13.
University of Göttingen. Brown, S. A., Zumbrunn, G., et al. (2002). Rhythms of mam-
malian body temperature can sustain peripheral circadian
clocks. Current Biology, 12(18), 1574–1583.
References Buhr, E. D., Yoo, S. H., et al. (2010). Temperature as a univer-
sal resetting cue for mammalian circadian oscillators. Sci-
Abarca, C., Albrecht, U., et al. (2002). Cocaine sensitization ence, 330(6002), 379–385.
and reward are under the influence of circadian genes and Buijs, R. M., Kalsbeek, A., et al. (1993). Suprachiasmatic nucleus
rhythm. Proceedings of the National Academy of Sciences lesion increases corticosterone secretion. American Journal of
of the United States of America, 99(13), 9026–9030. Physiology. Regulatory, Integrative and Comparative Physiol-
Abe, M., Herzog, E. D., et al. (2002). Circadian rhythms in ogy, 264(6), R1186–R1192.
isolated brain regions. Journal of Neuroscience, 22(1), Buijs, R. M., La Fleur, S. E., et al. (2003a). The
350–356. suprachiasmatic nucleus balances sympathetic and para-
Abraham, U., Granada, A. E., et al. (2010). Coupling governs sympathetic output to peripheral organs through separate
entrainment range of circadian clocks. Molecular Systems preautonomic neurons. The Journal of Comparative
Biology, 6, 438. Neurology, 464(1), 36–48.
Akhtar, R. A., Reddy, A. B., et al. (2002). Circadian cycling of Buijs, R. M., van Eden, C. G., et al. (2003b). The biological
the mouse liver transcriptome, as revealed by cDNA micro- clock tunes the organs of the body: Timing by hormones
array, is driven by the suprachiasmatic nucleus. Current and the autonomic nervous system. Journal of Endocrinol-
Biology, 12(7), 540–550. ogy, 177(1), 17–26.
Alam, M. N., Kumar, S., et al. (2005). GABA-mediated control Buijs, R. M., Wortel, J., et al. (1999). Anatomical and func-
of hypocretin—But not melanin-concentrating hormone- tional demonstration of a multisynaptic suprachiasmatic
immunoreactive neurones during sleep in rats. The Journal nucleus adrenal (cortex) pathway. European Journal of
of Physiology, 563(Pt 2), 569–582. Neuroscience, 11(5), 1535–1544.
Anglin, R. E., Rosebush, P. I., et al. (2006). The neuropsychi- Burdakov, D., Gerasimenko, O., et al. (2005). Physiological
atric profile of Addison’s disease: Revisiting a forgotten changes in glucose differentially modulate the excitability
176

of hypothalamic melanin-concentrating hormone and orexin the central pacemaker in the suprachiasmatic nucleus. Genes
neurons in situ. Journal of Neuroscience, 25(9), 2429–2433. and Development, 14(23), 2950–2961.
Cai, X. J., Evans, M. L., et al. (2001). Hypoglycemia activates Davidson, A. J. (2009). Lesion studies targeting food-anticipatory
orexin neurons and selectively increases hypothalamic activity. European Journal of Neuroscience, 30(9), 1658–1664.
orexin-B levels: Responses inhibited by feeding and possibly DeFalco, J., Tomishima, M., et al. (2001). Virus-assisted
mediated by the nucleus of the solitary tract. Diabetes, 50(1), mapping of neural inputs to a feeding center in the hypo-
105–112. thalamus. Science, 291(5513), 2608–2613.
Cailotto, C., La Fleur, S. E., et al. (2005). The suprachiasmatic Delaunay, F., Khan, A., et al. (1997). Pancreatic beta cells are
nucleus controls the daily variation of plasma glucose via the important targets for the diabetogenic effects of
autonomic output to the liver: Are the clock genes involved? glucocorticoids. The Journal of Clinical Investigation, 100(8),
European Journal of Neuroscience, 22(10), 2531–2540. 2094–2098.
Cailotto, C., van Heijningen, C., et al. (2008). Daily rhythms in Diaz-Munoz, M., Vazquez-Martinez, O., et al. (2000). Antici-
metabolic liver enzymes and plasma glucose require a bal- patory changes in liver metabolism and entrainment of insu-
ance in the autonomic output to the liver. Endocrinology, lin, glucagon, and corticosterone in food-restricted rats.
149(4), 1914–1925. American Journal of Physiology. Regulatory, Integrative
Campbell, J. E., Peckett, A. J., et al. (2011). Adipogenic and and Comparative Physiology, 279(6), R2048–R2056.
lipolytic effects of chronic glucocorticoid exposure. Ameri- Dibner, C., Schibler, U., et al. (2010). The mammalian circa-
can Journal of Physiology. Cell Physiology, 300(1), dian timing system: Organization and coordination of cen-
C198–C209. tral and peripheral clocks. Annual Review of Physiology,
Carroll, T., & Findling, J. (2010). The diagnosis of Cushing’s 72, 517–549.
syndrome. Reviews in Endocrine and Metabolic Disorders, Doi, R., Oishi, K., et al. (2010). CLOCK regulates circadian
11(2), 147–153. rhythms of hepatic glycogen synthesis through transcrip-
Cascio, C. S., Shinsako, J., et al. (1987). The suprachiasmatic tional activation of Gys2. Journal of Biological Chemistry,
nuclei stimulate evening ACTH secretion in the rat. Brain 285(29), 22114–22121.
Research, 423(1–2), 173–178. Earnest, D. J., & Sladek, C. D. (1986). Circadian rhythms of
Challet, E., Mendoza, J., et al. (2009). Neurogenetics of food vasopressin release from individual rat suprachiasmatic
anticipation. European Journal of Neuroscience, 30(9), explants in vitro. Brain Research, 382(1), 129–133.
1676–1687. Escobar, C., Diaz-Munoz, M., et al. (1998). Persistence of met-
Chen, B. T., Hopf, F. W., et al. (2010). Synaptic plasticity in abolic rhythmicity during fasting and its entrainment by
the mesolimbic system: Therapeutic implications for sub- restricted feeding schedules in rats. American Journal of
stance abuse. Annals of the New York Academy of Sciences, Physiology, 274(5 Pt 2), R1309–R1316.
1187, 129–139. Feillet, C. A., Mendoza, J., et al. (2008). Restricted feeding
Cheng, M. Y., Bullock, C. M., et al. (2002). Prokineticin 2 tra- restores rhythmicity in the pineal gland of arrhythmic
nsmits the behavioural circadian rhythm of the suprachiasmatic-lesioned rats. European Journal of Neuro-
suprachiasmatic nucleus. Nature, 417(6887), 405–410. science, 28(12), 2451–2458.
Chung, S., Son, G. H., et al. (2011a). Adrenal peripheral oscilla- Feillet, C. A., Ripperger, J. A., et al. (2006). Lack of food
tor in generating the circadian glucocorticoid rhythm. Annals anticipation in Per2 mutant mice. Current Biology, 16(20),
of the New York Academy of Sciences, 1220(1), 71–81. 2016–2022.
Chung, S., Son, G. H., et al. (2011b). Circadian rhythm of adre- Fioramonti, X., Contie, S., et al. (2007). Characterization of
nal glucocorticoid: Its regulation and clinical implications. glucosensing neuron subpopulations in the arcuate nucleus:
Biochimica et Biophysica Acta (BBA)—Molecular Basis of Integration in neuropeptide Y and pro-opio melanocortin
Disease, 1812(5), 581–591. networks? Diabetes, 56(5), 1219–1227.
Claret, M., Smith, M. A., et al. (2007). AMPK is essential for Firsov, D., Tokonami, N., et al. (2011). Role of the renal circa-
energy homeostasis regulation and glucose sensing by dian timing system in maintaining water and electrolytes
POMC and AgRP neurons. The Journal of Clinical Investi- homeostasis. Molecular and Cellular Endocrinology, 349
gation, 117(8), 2325–2336. (1), 51–55.
Claudel, T., Cretenet, G., et al. (2007). Crosstalk between Gerhold, L. M., Horvath, T. L., et al. (2001). Vasoactive intes-
xenobiotics metabolism and circadian clock. FEBS Letters, tinal peptide fibers innervate neuroendocrine dopaminergic
581(19), 3626–3633. neurons. Brain Research, 919(1), 48–56.
Cowell, I. G. (2002). E4BP4/NFIL3, a PAR-related bZIP fac- Gillette, M. U., & Reppert, S. M. (1987). The hypothalamic
tor with many roles. BioEssays, 24(11), 1023–1029. suprachiasmatic nuclei: Circadian patterns of vasopressin
Damiola, F., Le Minh, N., et al. (2000). Restricted feeding secretion and neuronal activity in vitro. Brain Research
uncouples circadian oscillators in peripheral tissues from Bulletin, 19(1), 135–139.
177

Gimble, J. M., & Floyd, Z. E. (2011). Metabolism: What Honma, S., Honma, K., et al. (1988). Rhythms in behaviors,
causes the gut’s circadian instincts? Current Biology, 21 body temperature and plasma corticosterone in SCN
(16), R624–R626. lesioned rats given methamphetamine. Physiology and
Gimble, J. M., Sutton, G. M., et al. (2011). Prospective Behavior, 44(2), 247–255.
influences of circadian clocks in adipose tissue and metabo- Honma, S., Yasuda, T., et al. (2008). Circadian behavioral
lism. Nature Reviews. Endocrinology, 7(2), 98–107. rhythms in Cry1/Cry2 double-deficient mice induced by
Gomez, F., Chapleur, M., et al. (1997). Arginine vasopressin methamphetamine. Journal of Biological Rhythms, 23(1),
(AVP) depletion in neurons of the suprachiasmatic nuclei 91–94.
affects the AVP content of the paraventricular neurons Huang, W., Ramsey, K. M., et al. (2011). Circadian rhythms,
and stimulates adrenocorticotrophic hormone release. Jour- sleep, and metabolism. The Journal of Clinical Investigation,
nal of Neuroscience Research, 50(4), 565–574. 121(6), 2133–2141.
Granados-Fuentes, D., Prolo, L. M., et al. (2004a). The Hughes, M. E., DiTacchio, L., et al. (2009). Harmonics of cir-
suprachiasmatic nucleus entrains, but does not sustain, circa- cadian gene transcription in mammals. PLoS Genetics, 5(4),
dian rhythmicity in the olfactory bulb. Journal of Neurosci- e1000442.
ence, 24(3), 615–619. Ibrahim, N., Bosch, M. A., et al. (2003). Hypothalamic pro-
Granados-Fuentes, D., Saxena, M. T., et al. (2004b). Olfactory opiomelanocortin neurons are glucose responsive and
bulb neurons express functional, entrainable circadian express K(ATP) channels. Endocrinology, 144(4), 1331–1340.
rhythms. European Journal of Neuroscience, 19(4), 898–906. Ishida, A., Mutoh, T., et al. (2005). Light activates the adrenal
Granados-Fuentes, D., Tseng, A., et al. (2006). A circadian gland: Timing of gene expression and glucocorticoid release.
clock in the olfactory bulb controls olfactory responsivity. Cell Metabolism, 2(5), 297–307.
Journal of Neuroscience, 26(47), 12219–12225. Jasper, M. S., & Engeland, W. C. (1997). Splanchnicotomy
Green, D. J., & Gillette, R. (1982). Circadian rhythm of firing increases adrenal sensitivity to ACTH in nonstressed rats.
rate recorded from single cells in the rat suprachiasmatic American Journal of Physiology. Endocrinology and Metab-
brain slice. Brain Research, 245(1), 198–200. olism, 273(2), E363–E368.
Groos, G., & Hendriks, J. (1982). Circadian rhythms in electri- Jin, J. Y., DuBois, D. C., et al. (2004). Receptor/gene-
cal discharge of rat suprachiasmatic neurones recorded mediated pharmacodynamic effects of methylprednisolone
in vitro. Neuroscience Letters, 34(3), 283–288. on phosphoenolpyruvate carboxykinase regulation in rat
Guilding, C., Hughes, A., et al. (2009). A riot of rhythms: Neu- liver. Journal of Pharmacology and Experimental Therapeu-
ronal and glial circadian oscillators in the mediobasal hypo- tics, 309(1), 328–339.
thalamus. Molecular Brain, 2(1), 28. Kalra, S. P., Dube, M. G., et al. (1999). Interacting appetite-
Guilding, C., & Piggins, H. D. (2007). Challenging the omnip- regulating pathways in the hypothalamic regulation of body
otence of the suprachiasmatic timekeeper: Are circadian weight. Endocrine Reviews, 20(1), 68–100.
oscillators present throughout the mammalian brain? Euro- Kalsbeek, A., Buijs, R. M., et al. (1992). Vasopressin-
pean Journal of Neuroscience, 25(11), 3195–3216. containing neurons of the suprachiasmatic nuclei inhibit cor-
Guyon, A., Tardy, M. P., et al. (2009). Glucose inhibition per- ticosterone release. Brain Research, 580(1–2), 62–67.
sists in hypothalamic neurons lacking tandem-pore Kþ Kalsbeek, A., Foppen, E., et al. (2008). Circadian control of
channels. Journal of Neuroscience, 29(8), 2528–2533. the daily plasma glucose rhythm: An interplay of GABA
Hampp, G., Ripperger, J. A., et al. (2008). Regulation of mono- and glutamate. PLoS One, 3(9), e3194.
amine oxidase A by circadian-clock components implies Kalsbeek, A., La Fleur, S., et al. (2004). Suprachiasmatic
clock influence on mood. Current Biology, 18(9), 678–683. GABAergic inputs to the paraventricular nucleus control
Hankins, M. W., Peirson, S. N., et al. (2008). Melanopsin: An plasma glucose concentrations in the rat via sympathetic
exciting photopigment. Trends in Neurosciences, 31(1), innervation of the liver. Journal of Neuroscience, 24(35),
27–36. 7604–7613.
Hardin, P. E. (2004). Transcription regulation within the circa- Kalsbeek, A., & Strubbe, J. H. (1998). Circadian control of
dian clock: The E-box and beyond. Journal of Biological insulin secretion is independent of the temporal distribution
Rhythms, 19(5), 348–360. of feeding. Physiology and Behavior, 63(4), 553–558.
Honma, K., Honma, S., et al. (1986). Disorganization of the rat Kalsbeek, A., Yi, C. X., et al. (2010). The hypothalamic clock
activity rhythm by chronic treatment with methamphet- and its control of glucose homeostasis. Trends in Endocri-
amine. Physiology and Behavior, 38(5), 687–695. nology and Metabolism, 21(7), 402–410.
Honma, K., Honma, S., et al. (1987). Activity rhythms in the Kamphuis, W., Cailotto, C., et al. (2005). Circadian expression
circadian domain appear in suprachiasmatic nuclei lesioned of clock genes and clock-controlled genes in the rat retina.
rats given methamphetamine. Physiology and Behavior, 40 Biochemical and Biophysical Research Communications,
(6), 767–774. 330(1), 18–26.
178

Kiessling, S., Eichele, G., et al. (2010). Adrenal glucocorticoids Li, X., Sankrithi, N., et al. (2002). Transforming growth factor-
have a key role in circadian resynchronization in a mouse alpha is expressed in astrocytes of the suprachiasmatic
model of jet lag. The Journal of Clinical Investigation, 120 nucleus in hamster: Role of glial cells in circadian clocks.
(7), 2600–2609. Neuroreport, 13(16), 2143–2147.
Kohsaka, A., Laposky, A. D., et al. (2007). High-fat diet dis- Liu, A. C., Welsh, D. K., et al. (2007). Intercellular coupling
rupts behavioral and molecular circadian rhythms in mice. confers robustness against mutations in the SCN circadian
Cell Metabolism, 6(5), 414–421. clock network. Cell, 129(3), 605–616.
Kornmann, B., Schaad, O., et al. (2007a). System-driven and Lu, X. Y., Shieh, K. R., et al. (2002). Diurnal rhythm of agouti-
oscillator-dependent circadian transcription in mice with a related protein and its relation to corticosterone and food
conditionally active liver clock. PLoS Biology, 5(2), e34. intake. Endocrinology, 143(10), 3905–3915.
Kornmann, B., Schaad, O., et al. (2007b). Regulation of circa- Luiten, P. G., ter Horst, G. J., et al. (1987). The hypothalamus,
dian gene expression in liver by systemic signals and hepato- intrinsic connections and outflow pathways to the endocrine
cyte oscillators. Cold Spring Harbor Symposia on system in relation to the control of feeding and metabolism.
Quantitative Biology, 72, 319–330. Progress in Neurobiology, 28(1), 1–54.
Kramer, A., Yang, F. C., et al. (2001). Regulation of daily Marcheva, B., Ramsey, K. M., et al. (2010). Disruption of the
locomotor activity and sleep by hypothalamic EGF receptor clock components CLOCK and BMAL1 leads to hypo-
signaling. Science, 294(5551), 2511–2515. insulinaemia and diabetes. Nature, 466(7306), 627–631.
Kraves, S., & Weitz, C. J. (2006). A role for cardiotrophin-like Masubuchi, S., Honma, S., et al. (2000). Clock genes outside
cytokine in the circadian control of mammalian locomotor the suprachiasmatic nucleus involved in manifestation of
activity. Nature Neuroscience, 9(2), 212–219. locomotor activity rhythm in rats. European Journal of Neu-
La Fleur, S. E. (2003). Daily rhythms in glucose metabolism: roscience, 12(12), 4206–4214.
Suprachiasmatic nucleus output to peripheral tissue. Journal Masubuchi, S., Honma, S., et al. (2001). Circadian activity
of Neuroendocrinology, 15(3), 315–322. rhythm in methamphetamine-treated Clock mutant mice.
La Fleur, S. E., Kalsbeek, A., et al. (1999). A suprachiasmatic European Journal of Neuroscience, 14(7), 1177–1180.
nucleus generated rhythm in basal glucose concentrations. McCarthy, J. J., Andrews, J. L., et al. (2007). Identification of
Journal of Neuroendocrinology, 11(8), 643–652. the circadian transcriptome in adult mouse skeletal muscle.
La Fleur, S. E., Kalsbeek, A., et al. (2000). Polysynaptic neural Physiological Genomics, 31(1), 86–95.
pathways between the hypothalamus, including the McClung, C. A., Sidiropoulou, K., et al. (2005). Regulation of
suprachiasmatic nucleus, and the liver. Brain Research, 871 dopaminergic transmission and cocaine reward by the Clock
(1), 50–56. gene. Proceedings of the National Academy of Sciences of
La Fleur, S. E., Kalsbeek, A., et al. (2001). A daily rhythm in the United States of America, 102(26), 9377–9381.
glucose tolerance: A role for the suprachiasmatic nucleus. McNamara, P., Seo, S. B., et al. (2001). Regulation of CLOCK
Diabetes, 50(6), 1237–1243. and MOP4 by nuclear hormone receptors in the vasculature:
Lamia, K. A., Sachdeva, U. M., et al. (2009). AMPK regulates A humoral mechanism to reset a peripheral clock. Cell, 105
the circadian clock by cryptochrome phosphorylation and (7), 877–889.
degradation. Science, 326(5951), 437–440. Mihaylova, M. M., & Shaw, R. J. (2011). The AMPK signalling
Lamia, K. A., Storch, K.-F., et al. (2008). Physiological signifi- pathway coordinates cell growth, autophagy and metabo-
cance of a peripheral tissue circadian clock. Proceedings of lism. Nature Cell Biology, 13(9), 1016–1023.
the National Academy of Sciences, 105(39), 15172–15177. Mistlberger, R. E. (1994). Circadian food-anticipatory activity:
Lamont, E. W., Robinson, B., et al. (2005). The central and Formal models and physiological mechanisms. Neuroscience
basolateral nuclei of the amygdala exhibit opposite diurnal and Biobehavioral Reviews, 18(2), 171–195.
rhythms of expression of the clock protein Period2. Pro- Moore, R. Y., & Eichler, V. B. (1972). Loss of a circadian
ceedings of the National Academy of Sciences of the United adrenal corticosterone rhythm following suprachiasmatic
States of America, 102(11), 4180–4184. lesions in the rat. Brain Research, 42(1), 201–206.
Le Minh, N., Damiola, F., et al. (2001). Glucocorticoid hor- Muroya, S., Funahashi, H., et al. (2004). Orexins (hypocretins)
mones inhibit food-induced phase-shifting of peripheral directly interact with neuropeptide Y, POMC and glucose-
circadian oscillators. EMBO Journal, 20(24), 7128–7136. responsive neurons to regulate Ca2 þ signaling in a recipro-
Lee, C., Etchegaray, J. P., et al. (2001). Posttranslational cal manner to leptin: Orexigenic neuronal pathways in the
mechanisms regulate the mammalian circadian clock. Cell, mediobasal hypothalamus. European Journal of Neurosci-
107(7), 855–867. ence, 19(6), 1524–1534.
Levitan, R. D. (2007). The chronobiology and neurobiology of Muroya, S., Yada, T., et al. (1999). Glucose-sensitive neurons in
winter seasonal affective disorder. Dialogues in Clinical the rat arcuate nucleus contain neuropeptide Y. Neuroscience
Neuroscience, 9(3), 315–324. Letters, 264(1–3), 113–116.
179

Nader, N., Chrousos, G. P., et al. (2009). Circadian rhythm Reddy, A. B., & O’Neill, J. S. (2010). Healthy clocks, healthy
transcription factor CLOCK regulates the transcriptional body, healthy mind. Trends in Cell Biology, 20(1), 36–44.
activity of the glucocorticoid receptor by acetylating its Reick, M., Garcia, J. A., et al. (2001). NPAS2: An analog of
hinge region lysine cluster: Potential physiological clock operative in the mammalian forebrain. Science, 293
implications. The FASEB Journal, 23(5), 1572–1583. (5529), 506–509.
Nagoshi, E., Saini, C., et al. (2004). Circadian gene expression Ripperger, J. A., & Schibler, U. (2006). Rhythmic CLOCK-
in individual fibroblasts: Cell-autonomous and self-sustained BMAL1 binding to multiple E-box motifs drives circadian
oscillators pass time to daughter cells. Cell, 119(5), 693–705. Dbp transcription and chromatin transitions. Nature Genetics,
Nakahata, Y., Kaluzova, M., et al. (2008). The NADþ- 38(3), 369–374.
dependent deacetylase SIRT1 modulates CLOCK-mediated Ripperger, J. A., Shearman, L. P., et al. (2000). CLOCK, an
chromatin remodeling and circadian control. Cell, 134(2), essential pacemaker component, controls expression of the
329–340. circadian transcription factor DBP. Genes and Development,
Nakahata, Y., Sahar, S., et al. (2009). Circadian control of 14(6), 679–689.
the NADþ salvage pathway by CLOCK-SIRT1. Science, Rosenfeld, P., Van Eekelen, J. A., et al. (1988). Ontogeny
324(5927), 654–657. of the type 2 glucocorticoid receptor in discrete rat brain
Oishi, K., Amagai, N., et al. (2005). Genome-wide expression regions: An immunocytochemical study. Brain Research,
analysis reveals 100 adrenal gland-dependent circadian 470(1), 119–127.
genes in the mouse liver. DNA Research, 12(3), 191–202. Rosenwasser, A. M., Fecteau, M. E., et al. (2005). Effects of
Oishi, K., Miyazaki, K., et al. (2003). Genome-wide expression ethanol intake and ethanol withdrawal on free-running cir-
analysis of mouse liver reveals CLOCK-regulated circadian cadian activity rhythms in rats. Physiology and Behavior,
output genes. Journal of Biological Chemistry, 278(42), 84(4), 537–542.
41519–41527. Roybal, K., Theobold, D., et al. (2007). Mania-like behavior
Oster, H., Damerow, S., et al. (2006a). Transcriptional induced by disruption of CLOCK. Proceedings of the
profiling in the adrenal gland reveals circadian regulation National Academy of Sciences of the United States of Amer-
of hormone biosynthesis genes and nucleosome assembly ica, 104(15), 6406–6411.
genes. Journal of Biological Rhythms, 21(5), 350–361. Rudic, R., McNamara, P., et al. (2004). BMAL1 and CLOCK,
Oster, H., Damerow, S., et al. (2006b). The circadian rhythm two essential components of the circadian clock, are
of glucocorticoids is regulated by a gating mechanism resid- involved in glucose homeostasis. PLoS Biology, 2(11), e377.
ing in the adrenal cortical clock. Cell Metabolism, 4(2), Ruis, J. F., Buys, J. P., et al. (1990). Effects of T cycles of light/
163–173. darkness and periodic forced activity on methamphetamine-
Panda, S., Antoch, M. P., et al. (2002). Coordinated transcrip- induced rhythms in intact and SCN-lesioned rats: Explana-
tion of key pathways in the mouse by the circadian clock. tion by an hourglass-clock model. Physiology and Behavior,
Cell, 109(3), 307–320. 47(5), 917–929.
Phillipson, O. T., & Griffiths, A. C. (1985). The topographic Rutter, J., Reick, M., et al. (2002). Metabolism and the control
order of inputs to nucleus accumbens in the rat. Neurosci- of circadian rhythms. Annual Review of Biochemistry, 71,
ence, 16(2), 275–296. 307–331.
Preitner, N., Damiola, F., et al. (2002). The orphan nuclear Sadacca, L., Lamia, K., et al. (2011). An intrinsic circadian clock
receptor REV-ERBalpha controls circadian transcription of the pancreas is required for normal insulin release and
within the positive limb of the mammalian circadian oscilla- glucose homeostasis in mice. Diabetologia, 54(1), 120–124.
tor. Cell, 110(2), 251–260. Saeb-Parsy, K., Lombardelli, S., et al. (2000). Neural
Puschel, G. P. (2004). Control of hepatocyte metabolism by connections of hypothalamic neuroendocrine nuclei in the
sympathetic and parasympathetic hepatic nerves. The rat. Journal of Neuroendocrinology, 12(7), 635–648.
Anatomical Record. Part A, Discoveries in Molecular, Cellu- Sakoda, H., Ogihara, T., et al. (2000). Dexamethasone-
lar, and Evolutionary Biology, 280(1), 854–867. induced insulin resistance in 3T3-L1 adipocytes is due to
Ralph, M. R., Foster, R. G., et al. (1990). Transplanted inhibition of glucose transport rather than insulin signal
suprachiasmatic nucleus determines circadian period. Sci- transduction. Diabetes, 49(10), 1700–1708.
ence, 247(4945), 975–978. Sasaki, K., Cripe, T. P., et al. (1984). Multihormonal regulation
Ramsey, K. M., Yoshino, J., et al. (2009). Circadian clock feed- of phosphoenolpyruvate carboxykinase gene transcription.
back cycle through NAMPT-mediated NADþ biosynthesis. The dominant role of insulin. Journal of Biological Chemis-
Science, 324(5927), 651–654. try, 259(24), 15242–15251.
Reddy, A. B., Maywood, E. S., et al. (2007). Glucocorticoid Segall, L. A., Perrin, J. S., et al. (2006). Glucocorticoid
signaling synchronizes the liver circadian transcriptome. rhythms control the rhythm of expression of the clock pro-
Hepatology, 45(6), 1478–1488. tein, Period2, in oval nucleus of the bed nucleus of the stria
180

terminalis and central nucleus of the amygdala in rats. Tosini, G., & Menaker, M. (1996). Circadian rhythms in
Neuroscience, 140(3), 753–757. cultured mammalian retina. Science, 272(5260), 419–421.
Sellix, M. T., Egli, M., et al. (2006). Anatomical and functional Turek, F. W., Joshu, C., et al. (2005). Obesity and metabolic
characterization of clock gene expression in neuroendocrine syndrome in circadian clock mutant mice. Science, 308
dopaminergic neurons. American Journal of Physiology. (5724), 1043–1045.
Regulatory, Integrative and Comparative Physiology, 290 Ueda, H. R., Chen, W., et al. (2002). A transcription factor
(5), R1309–R1323. response element for gene expression during circadian
Sellix, M. T., & Freeman, M. E. (2003). Circadian rhythms of night. Nature, 418(6897), 534–539.
neuroendocrine dopaminergic neuronal activity in ovariec- Um, J. H., Yang, S., et al. (2007). Activation of 50 -AMP-
tomized rats. Neuroendocrinology, 77(1), 59–70. activated kinase with diabetes drug metformin induces
Shibata, S., Oomura, Y., et al. (1982). Circadian rhythmic changes casein kinase Iepsilon (CKIepsilon)-dependent degradation
of neuronal activity in the suprachiasmatic nucleus of the rat of clock protein mPer2. Journal of Biological Chemistry,
hypothalamic slice. Brain Research, 247(1), 154–158. 282(29), 20794–20798.
Shibley, H. L., Malcolm, R. J., et al. (2008). Adolescents with Uz, T., Ahmed, R., et al. (2005). Effect of fluoxetine and
insomnia and substance abuse: Consequences and com- cocaine on the expression of clock genes in the mouse hip-
orbidities. Journal of Psychiatric Practice, 14(3), 146–153. pocampus and striatum. Neuroscience, 134(4), 1309–1316.
Silver, R., LeSauter, J., et al. (1996). A diffusible coupling signal Vander Kooi, B. T., Onuma, H., et al. (2005). The glucose-6-
from the transplanted suprachiasmatic nucleus controlling phosphatase catalytic subunit gene promoter contains both
circadian locomotor rhythms. Nature, 382(6594), 810–813. positive and negative glucocorticoid response elements.
Son, G. H., Chung, S., et al. (2008). Adrenal peripheral clock Molecular Endocrinology, 19(12), 3001–3022.
controls the autonomous circadian rhythm of glucocorticoid Vicentic, A., Lakatos, A., et al. (2005). CART peptide diurnal
by causing rhythmic steroid production. Proceedings of the rhythm in brain and effect of fasting. Brain Research, 1032
National Academy of Sciences, 105(52), 20970–20975. (1–2), 111–115.
Spanagel, R., Pendyala, G., et al. (2005). The clock gene Per2 Watts, A. G., & Swanson, L. W. (1987). Efferent projections of
influences the glutamatergic system and modulates alcohol the suprachiasmatic nucleus: II. Studies using retrograde
consumption. Nature Medicine, 11(1), 35–42. transport of fluorescent dyes and simultaneous peptide
Stephan, F. K. (2002). The “other” circadian system: Food as immunohistochemistry in the rat. The Journal of Compara-
a zeitgeber. Journal of Biological Rhythms, 17(4), 284–292. tive Neurology, 258(2), 230–252.
Stephan, F. K., Swann, J. M., et al. (1979). Entrainment of cir- Watts, A. G., Swanson, L. W., et al. (1987). Efferent
cadian rhythms by feeding schedules in rats with projections of the suprachiasmatic nucleus: I. Studies using
suprachiasmatic lesions. Behavioral and Neural Biology, 25 anterograde transport of Phaseolus vulgaris leucoagglutinin
(4), 545–554. in the rat. The Journal of Comparative Neurology, 258(2),
Stephan, F. K., & Zucker, I. (1972). Circadian rhythms in drink- 204–229.
ing behavior and locomotor activity of rats are eliminated by Webb, I. C., Baltazar, R. M., et al. (2009). Bidirectional
hypothalamic lesions. Proceedings of the National Academy interactions between the circadian and reward systems: Is
of Sciences of the United States of America, 69(6), 1583–1586. restricted food access a unique zeitgeber? European Journal
Storch, K. F., Lipan, O., et al. (2002). Extensive and divergent of Neuroscience, 30(9), 1739–1748.
circadian gene expression in liver and heart. Nature, 417 Weinstein, S. P., Wilson, C. M., et al. (1998). Dexamethasone
(6884), 78–83. inhibits insulin-stimulated recruitment of GLUT4 to the cell
Storch, K. F., Paz, C., et al. (2007). Intrinsic circadian clock of surface in rat skeletal muscle. Metabolism, 47(1), 3–6.
the mammalian retina: Importance for retinal processing of Welsh, D. K., Logothetis, D. E., et al. (1995). Individual
visual information. Cell, 130(4), 730–741. neurons dissociated from rat suprachiasmatic nucleus
Storch, K. F., & Weitz, C. J. (2009). Daily rhythms of express independently phased circadian firing rhythms.
food-anticipatory behavioral activity do not require the Neuron, 14(4), 697–706.
known circadian clock. Proceedings of the National Acad- Welsh, D. K., Takahashi, J. S., et al. (2010). Suprachiasmatic
emy of Sciences of the United States of America, 106(16), nucleus: Cell autonomy and network properties. Annual
6808–6813. Review of Physiology, 72, 551–577.
Surjit, M., Ganti, Krishna P., et al. (2011). Widespread negative Westrich, L., & Sprouse, J. (2010). Circadian rhythm dys-
response elements mediate direct repression by agonist- regulation in bipolar disorder. Current Opinion in Investiga-
liganded glucocorticoid receptor. Cell, 145(2), 224–241. tional Drugs, 11(7), 779–787.
Szafarczyk, A., Ixart, G., et al. (1983). CNS control of the cir- Willie, J. T., Chemelli, R. M., et al. (2001). To eat or to sleep?
cadian adrenocortical rhythm. Journal of Steroid Biochemis- Orexin in the regulation of feeding and wakefulness. Annual
try, 19(1 Pt 3), 1009–1015. Review of Neuroscience, 24, 429–458.
181

Yamazaki, S., Numano, R., et al. (2000). Resetting central and Yuferov, V., Kroslak, T., et al. (2003). Differential gene
peripheral circadian oscillators in transgenic rats. Science, expression in the rat caudate putamen after “binge” cocaine
288(5466), 682–685. administration: Advantage of triplicate microarray analysis.
Yao, Z., DuBois, D., et al. (2006). Modeling circadian rhythms Synapse, 48(4), 157–169.
of glucocorticoid receptor and glutamine synthetase expres- Zhang, E. E., & Kay, S. A. (2010). Clocks not winding down:
sion in rat skeletal muscle. Pharmaceutical Research, 23(4), Unravelling circadian networks. Nature Reviews Molecular
670–679. Cell Biology, 11(11), 764–776.
Yi, C.-X., Serlie, M. J., et al. (2009). A major role for Zhang, E. E., Liu, Y., et al. (2010). Cryptochrome mediates
perifornical orexin neurons in the control of glucose metab- circadian regulation of cAMP signaling and hepatic gluco-
olism in rats. Diabetes, 58(9), 1998–2005. neogenesis. Nature Medicine, 16(10), 1152–1156.
Yoo, S. H., Yamazaki, S., et al. (2004). PERIOD2::LUCIFER- Zhang, S., Zeitzer, J. M., et al. (2004). Lesions of the
ASE real-time reporting of circadian dynamics reveals per- suprachiasmatic nucleus eliminate the daily rhythm of
sistent circadian oscillations in mouse peripheral tissues. hypocretin-1 release. Sleep, 27(4), 619–627.
Proceedings of the National Academy of Sciences of the
United States of America, 101(15), 5339–5346.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 11

Circadian rhythms in white adipose tissue

Rianne van der Spek{,*, Felix Kreier{, Eric Fliers{ and Andries Kalsbeek{,{

{
Department of Endocrinology and Metabolism, Academic Medical Center (AMC),
University of Amsterdam (UvA), Amsterdam, The Netherlands
{
Hypothalamic Integration Mechanisms, Netherlands Institute for Neuroscience, an Institute
of the Royal Netherlands Academy of Arts and Sciences, Amsterdam, The Netherlands

Abstract: Adipose tissue is an important endocrine organ. It is involved in the regulation of energy
metabolism by secreting factors (adipokines) that regulate appetite, food intake, glucose disposal,
and energy expenditure. Many of these adipokines display profound day/night rhythms, and
accumulating evidence links disruption of these rhythms to metabolic diseases such as obesity and type
2 diabetes. Here, we briefly present the circadian system, describe the development of white adipose
tissue (WAT) and its depot-specific characteristics and innervation, we discuss energy storage in WAT
and, lastly, review recent findings that link circadian rhythmicity to adipose tissue biology and obesity.

Keywords: white adipose tissue; circadian rhythm; SCN; suprachiasmatic nucleus; ANS; autonomic
nervous system; corticosterone; food intake; energy metabolism.

Introduction glucose and lipid metabolism. Once the amount


of fat has accumulated to such an extent that it
The prevalence of obesity is taking on enormous hinders tissue function, insulin resistance of
proportions. Obesity is defined by the World WAT may occur. Both human and rodent studies
Health Organisation as the accumulation of exces- point toward an important role for the circadian
sive fat tissue, to the point that it may impair timing system in energy metabolism. Disturbed
health (World Health Organisation, 2011). Fat, day/night rhythms are closely correlated with the
or white adipose tissue (WAT), is a metabolically development of obesity and type 2 diabetes
active tissue and a key player in the regulation of mellitus (T2DM). Here, we briefly present the cir-
cadian system and describe the development of
WAT and its depot-specific characteristics and
*Corresponding author.
innervation, we discuss energy storage in WAT
Tel.: þ31 20 566 6071; Fax: þ31 20 691 7682 and review recent findings that link circadian
E-mail: r.d.vanderspek@amc.uva.nl rhythmicity to adipose tissue biology and obesity.

http://dx.doi.org/10.1016/B978-0-444-59427-3.00011-3 183
184

Circadian rhythms cells. However, while within the SCN these


oscillators are synchronized via synaptic and para-
From an evolutionary perspective, circadian crine signals, they do not appear to communicate
(Latin; circa diem ¼ approximately a day) rhythms as much between cells in other tissues (Asher and
may have ensured continuous energy availability Schibler, 2011). Oscillators in non-SCN cells,
throughout the 24-h day for organisms relying known as peripheral clocks, are synchronized and
on photosynthesis (Hut and Beersma, 2011), entrained by the SCN via direct (neural or humoral
in addition to protecting the replicating DNA factors) and indirect (e.g., control of sleep–wake
from high ultraviolet radiation exposure during cycle, food intake, body temperature) cues.
daytime (Asher and Schibler, 2011). In mammals,
the circadian timing system allows the anticipation
of recurring events during the course of a day, Regulation of rhythms in peripheral tissues
along with temporal segregation of incompatible
processes such as sleep and wakefulness (Asher Since the discovery of the SCN in 1972, numerous
and Schibler, 2011). By definition, circadian neuroanatomical tracing studies have been per-
rhythms have a  24 h cycle, persist in the absence formed to reveal target areas for the circadian
of environmental cues (i.e., they are endogenous), information generated in the SCN. Generally,
adjust to match local time (so-called entrainment) these studies showed that the projection fibers
and maintain intrinsic periodicity over a range from the SCN are surprisingly limited and mainly
of physiological temperatures (i.e., temperature restricted to a few hypothalamic nuclei (Watts
compensation) (O’Neill and Reddy, 2011). and Swanson, 1987; Watts et al., 1987). In these
In mammals, the biological master clock resides neuroanatomical tracing studies, the hypotha-
in the suprachiasmatic nuclei (SCN) in the anterior lamic paraventricular nucleus (PVN) showed up
hypothalamus. The endogenous rhythmicity of as an important target area of the SCN. Although
mammals is thus generated within the SCN and the PVN is considered to be the neuroendocrine
distributed to the rest of the brain and to the body. “headquarters” of the hypothalamus, it also
To maintain synchrony with the outside world, the contains a prominent population of neurons
SCN integrate entraining signals (Zeitgebers) of projecting to the autonomic nervous system
which light, food intake, and locomotor activity (ANS), i.e., preautonomic neurons.
are the most important ones. At least three major Within the preautonomic neurons in the PVN,
input systems feed information to the SCN. The there is a clear separation of neurons projecting
first consists of melanopsin-containing ganglion to the sympathetic and the parasympathetic
cells within the retina projecting via the ret- branch of the ANS (Buijs et al., 2003; Kreier
inohypothalamic tract directly to the SCN using et al., 2006). Via these ANS projections, the
glutamate and pituitary adenylate cyclase-activating SCN are able to convey rhythmic information to
polypeptide (PACAP) as neurotransmitters. The various peripheral tissues. For example, the
second input system runs from the retina to the rhythmic release of melatonin by the pineal gland
intergeniculate leaflet (IGL) and from there via is controlled by SCN input to the preautonomic
the geniculohypothalamic tract to the SCN, using neurons in the PVN that regulate the sympathetic
neuropeptide Y (NPY), enkephalin, and GABA input to the pineal gland (Moore, 1996). Besides
as its main neurotransmitters. The third input sys- this circadian control, the ANS may play a
tem projects from the medial and dorsal raphe role in resetting peripheral clocks after phase
nuclei to the SCN and the IGL and uses serotonin shift-inducing light exposure (Cailotto et al.,
(5-HT) as its main transmitter (Dibner et al., 2010). 2009).
Circadian oscillators have virtually the same However, peripheral rhythmicity is not solely
molecular make-up in SCN neurons and non-SCN entrained by the ANS. Although a complete
185

autonomic hepatic denervation (both the sympa- the PVN and the ANS—on the adrenals and affects
thetic and parasympathetic branch) abolished their sensitivity to ACTH (Kalsbeek et al.,
the daily rhythm in plasma glucose concentrations, 2012; Lilley et al., 2012). Adrenal denervation
it had no overall effects on the molecular rhythms abolishes the circadian corticosterone rhythm, the
in the liver (Cailotto et al., 2005). Catecholamines light-induced corticosterone secretion, and the
like epinephrine and nor-epinephrine can modulate daily change in ACTH sensitivity (Engeland
clock function in vitro (Reilly et al., 2008), however and Arnhold, 2005; Ishida et al., 2005; Kalsbeek
dopamine b-hydroxylase (Dbh, the enzyme that et al., 2012).
converts dopamine into nor-epinephrine) knock- Next to innervation and hormonal control regu-
out mice displayed no disrupted molecular rhythms lation of the feeding–fasting cycle provides
in peripheral tissues (Reilly et al., 2008). Moreover, several indirect pathways for the SCN to entrain
Dbh/ mice treated with a- and b-adrenergic peripheral tissues. These pathways include daily
receptor antagonists showed undisturbed molecu- rhythms in hormones secreted upon feeding and
lar rhythms in heart, aorta, liver, and brown fasting (e.g., cholecystokinin, peptide YY, ghrelin,
adipose tissue (BAT; Reilly et al., 2008), thus leptin), changes in plasma concentrations of food
excluding the possible compensatory effect of metabolites (glucose, cholesterol, fatty acids,
dopamine. heme), postprandial temperature elevations, and
Parabiosis experiments have shown that non- intracellular changes in redox state (see section
neural signals suffice for the maintenance of circa- on nongenetic clock) (Dibner et al., 2010).
dian rhythms of clock gene expression in liver and At present, it is unclear to what extent each of
kidney, but not in heart, spleen, or skeletal mus- these Zeitgebers contributes to the regulation
cle (Guo et al., 2005). Moreover, in the submaxil- of circadian rhythms in peripheral clock systems,
lary glands, sympathetic input modulated but did in general, or to WAT in particular. Figure 1
not determine the phase of peripheral circadian summarizes the different pathways via which the
oscillators (Vujovic et al., 2008). Together, these SCN could relay its timing information to WAT.
data indicate that peripheral oscillators may
receive multiple signals that contribute to their
phase of entrainment.
Another example of this multiple signaling princi- Cellular timekeeping
ple is found in the hypothalamo–pituitary–adrenal
(HPA) axis, where the SCN use a two-stage mecha- Transcriptional–translational feedback model
nism to control the daily rhythm in plasma cortico-
sterone concentrations. Plasma glucocorticoid Cellular timekeeping is regulated on several
hormone levels exhibit robust daily oscillations, levels. On a genetic level, messenger RNA
and glucocorticoids are potent Zeitgebers, both (mRNA) is synthesized from DNA (transcrip-
in vitro and in vivo (Balsalobre et al., 2000), by tion). This mRNA may contain coding regions
binding to the glucocorticoids response element (exons) and non-coding regions (introns). To
in the promotor region of Period1 (Per1) and form a protein, introns are spliced out, so the
possibly Period2 (Per2) (Balsalobre et al., 2000; ribosome can translate the mRNA into an amino
So et al., 2009; Yamamoto et al., 2005). On the acid sequence, thereby producing a polypeptide
one hand, the SCN act on the corticotrophin- (translation). Subsequent posttranslational modi-
releasing hormone (CRH)-containing neuroendo- fication attaches other biochemical functional
crine motorneurons in the PVN to bring about the groups (e.g., phosphates) to the polypeptide, to
release of adrenocorticotrophic hormone (ACTH) activate or inactivate the protein, while protein
from the pituitary, while on the other hand, it folding gives the polypeptide a 3D structure,
also acts—through the preautonomic neurons in enabling interaction with other proteins. Changes
186

CRYs
P

PERs +
SCN

Lipolysis
Clock Bmal1 Lipogenesis
Glucose transport

Autonomic
Behavior innervation RevErba +

RORs

Hormonal Fig. 2. Simplified version of the molecular core clock


mechanism. The core loop is formed by Clock:Bmal1 and
Period 1-3 (Per1–3) and Cryptochrome 1 and 2 (Cry1-2). The
Clock:Bmal1 heterodimer stimulates the transcription of Per1–3
and Cry1–2. Subsequently, Pers and Crys heterodimerize,
translocate to the nucleus, and inhibit Clock:Bmal1 activity. As
a consequence, Clock:Bmal1 transcriptional activity drops,
which reduces the transcription of Per and Cry genes, thereby
activating Clock:Bmal1 again. Additional loops formed by
RevErbs and RORs enhance the robustness of the core loop.
Components of the core clock regulate energy metabolism, for
example, lipolysis, lipogenesis, and glucose transport.

Grip strength Adipokines Glucose


Clock:Bmal1 again. Figure 2 shows a simplified
Fig. 1. Three pathways via which the SCN could relay timing model of this molecular clock mechanism.
information to peripheral tissues. Additional loops enhance the robustness of the
core loop. The retinoic acid-related orphan
nuclear receptors, RevErbs and RORs, represent
additional regulatory loops by binding to retinoic
in the transcriptome do thus not necessarily acid-related orphan receptor response elements
reflect altered protein expression or function, on the Bmal1 promotor (Ko and Takahashi,
and vice versa, as has been shown for the liver 2006).
(Reddy et al., 2006).
In the transcriptional–translational feedback Nongenetic clock
model, the molecular clock is based on a core
loop formed by Clock:Bmal1 and Period 1–3 Cycles of transcription and translation feedback
(Per1–3) and Cryptochrome 1 and 2 (Cry1–2). generally operate on a timescale of up to a few
The Clock:Bmal1 heterodimer stimulates the hours and maintaining an oscillatory rhythm on
transcription of Per1–3 and Cry1–2, and Pers a daily basis thus requires a significant delay
and Crys subsequently heterodimerize, translo- between activation and repression of transcrip-
cate to the nucleus, and inhibit Clock:Bmal1 tion. Posttranslational modifications such as phos-
activity. As a consequence, Clock:Bmal1 tran- phorylation, histone acetylation, methylation,
scriptional activity drops, which reduces the tran- sumoylation, and ubiquitination affect the stabil-
scription of Per and Cry genes, thereby activating ity and nuclear translocation of core clock
187

proteins, thereby delaying the cycle to  24 h Adenosine monophosphate-activated protein


(Cardone et al., 2005; Gallego and Virshup, kinase (AMPK) activation has an impact on clock
2007; Ko and Takahashi, 2006; Reppert and function through various mechanisms. AMPK
Weaver, 2001). A more extended review of the senses the increase in AMP/ATP ratio in
genetic clock can be found in Ko and Takahashi conditions that deplete energy, such as hypoxia,
(2006) and Gallego and Virshup (2007). ischemia, and glucose deprivation. During these
Some observations cannot be explained suffi- conditions, AMPK may activate pathways that
ciently by this transcriptional–translational feed- generate ATP and suppress ATP-consuming pro-
back model. In the past few decades, several cesses. For instance, in the hypothalamus, AMPK
nongenetic pathways were found to play a role activity stimulates food intake, whereas in the
in cellular timekeeping. periphery AMPK activity stimulates uptake and
Nicotinamide adenine dinucleotide (NAD) is oxidation of fatty acids in addition to glucose
involved in reduction–oxidation (redox) reactions; uptake (Um et al., 2011). AMPK is part of the
it can accept (becoming NADH), carry, and core clock mechanism by degrading Per2 (via
donate electrons (returning to NAD). In vitro CKIe), phosphorylating Cry1 and thereby
experiments have shown that the redox state of decreasing its stability. In line with this concept,
NAD can regulate DNA binding activity of the studies in mouse embryonic stem cells lacking
Clock:Bmal1 heterodimer. Interestingly, in vivo both isoforms of the catalytic subunit (AMPKa1/
NAD levels are subjected to daily variations, a2 /) and in mice lacking either AMPKa1 or
thereby giving rhythmic input to the genetic clock AMPKa2 show altered free-running periods,
(Asher and Schibler, 2011; Merrow and indicating that AMPK plays a fundamental role
Roenneberg, 2001; Rutter et al., 2001), but there in the core clock (Um et al., 2011).
are also several indirect pathways via which Cyclic adenosine monophosphate (cAMP) is a
the redox state can be linked to the clock. The further example of an acute signaling pathway
enzymes silent information regulator protein that is tightly intertwined with the core clock.
(SIRT) and poly (ADP-ribose) polymerase 1 cAMP is a second messenger derived from ATP
(PARP-1) are both NAD-dependent enzymes. and is used for signal transduction in the cell, for
SIRT is expressed rhythmically and interacts with example, transferring the postsynaptic effects of
Clock:Bmal1 heterodimers, leading to rhythmic hormones that cannot pass the cell membrane.
deacetylation of Clock:Bmal1, histone H3, and In a series of elegant experiments, O’Neill et al.
Per2 (Asher and Schibler, 2011). When subjected showed that cAMP is not solely an output of the
to daytime feeding the liver of Parp-1/ mice SCN, but an integral component of the SCN pace-
showed a significantly delayed phase inversion of maker, regulating transcriptional cycles (O’Neill
clock gene expression, when compared with wild et al., 2008).
type mice, suggesting that PARP-1 activity is Heme has various biological functions; its most
indeed implicated in the phase entrainment of common function is as a component of hemoglo-
peripheral oscillators. However, PARP-1 activity bin, the red pigment in blood that transports
is probably driven by feeding rhythms rather oxygen and carbon dioxide. However, heme also
than by local circadian clocks (Asher and functions as a ligand for the—formerly orphan—
Schibler, 2011). RevErba and -b receptors. Binding of heme to
The energy status of a cell influences the redox the RevErbs enables the receptors to recruit
state, implying that via this pathway, food intake NCoR, their corepressor, and subsequently to
might be able to entrain the circadian system repress the transcription of target genes (e.g.,
(Arble et al., 2010; Maury et al., 2010; Merrow Bmal1), thereby regulating genes that are essen-
and Roenneberg, 2001). tial to the core clock. Furthermore, Per2 contains
188

heme as a prosthetic group (Burris, 2008), and investigated. It appears to be related to the con-
Bmal1, Per1, and Per2 regulate the expression trol of hematopoiesis and osteoblastogenesis by
of aminolevulinic acid synthase 1 (Alas1), a rate- acting as an energy store, but also via the release
limiting enzyme in heme production. Moreover, of paracrine factors (Casteilla et al., 2008). BAT
PGC1a (a transcription factor that mediates tran- and WAT share the ability to store lipids as
sition toward FA oxidation and gluconeogenesis triglycerides, but use them for different purposes.
when glucose is low) directly regulates the expres- BAT produces heat and plays an important role
sion of Alas1, identifying heme as another signal- in nonshivering thermogenesis. WAT, besides
ing molecule via which nutritional status can functioning as mechanical and thermal protection
influence the core clock mechanism. of vital organs and as an important long-term
Intriguingly, circadian rhythmicity has recently energy store, secretes several proteins that influence
been found in red blood cells, which have no processes as diverse as hemostasis, blood pressure,
nucleus or indeed any other organelles, and immune function, angiogenesis, and energy balance
therefore do not undergo transcription or trans- (Christodoulides et al., 2009). Besides adipocytes,
lation. Peroxiredoxins (PRX) are highly con- WAT comprises a stromavascular fraction including
served antioxidant proteins that scavenge macrophages, mesenchymal stem cells (MSCs),
cellular reactive oxygen species, and in red preadipocytes, endothelial and epithelial cells.
blood cells, PRX oxidation shows a clear circa-
dian rhythm (O’Neill and Reddy, 2011). This
illustrates very clearly that, in the absence of WAT development
transcription, circadian rhythms remain observ-
able in basic biochemical reactions. However, Adipocytes derive from MSCs cells residing in
when PRX rhythms were measured in an the adipose tissue stroma. The differentiation of
arrhythmic mouse model (Cry1/2 /), PRX MSCs into mature adipocytes (adipogenesis)
rhythms differed from controls, implying that, occurs in two phases. The first phase, known
although purely biochemical mechanisms are as determination, involves the commitment of a
able to sustain 24-h rhythms, they must normally pluripotent stem cell to the adipocyte lineage.
reciprocally interact with gene expression cycles Determination—an as yet poorly characterized
(Reddy and O’Neill, 2011). process—converts the stem cell into a
Taken together, these observations indicate preadipocyte. The second phase is terminal dif-
nongenetic pathways that have a fundamental ferentiation, in which the preadipocyte takes on
role in the core clock mechanism and represent the characteristics of the mature adipocyte: it
molecular routes via which nutritional status acquires the machinery that is necessary for
may entrain the clock. lipid transport and synthesis, insulin sensitivity,
and the secretion of adipocyte specific proteins.
Terminal differentiation involves a cascade of
White adipose tissue transcriptional events, mainly regulated by
CCAAT/enhancer-binding proteins and peroxi-
Adipose tissue is one of the largest organs in the some proliferator-activated receptor gamma
body, making up percentages of bodyweight from (PPARg) (Christodoulides et al., 2009; Rosen
5% in lean men to over 50% in the morbidly and MacDougald, 2006). Each fat depot (see
obese. In mammals, three functionally different the section “WAT depots”) has a unique pattern
types of adipose tissue have been described: of developmental gene expression and these
BAT, WAT, and bone marrow adipose tissue differences are largely independent of the func-
(BMAT). The role of BMAT is poorly tional state (Gesta et al., 2007).
189

WAT depots males and postmenopausal females, and a pear-


shaped gynoid (subcutaneous) fat distribution in
WAT is abundantly present throughout the body. premenopausal females (Elbers et al., 1999).
The main depots in rodents can be divided in those
found underneath the skin (subcutaneous; cervical, WAT innervation
dorsal, lumbar, abdominal, mammary, and glu-
teofemoral) and those found in the thoracic A role for the ANS in the regulation of fat metab-
and abdominal cavities. These internal fat olism was proposed as early as the 1920s, as
depots may be further subdivided in visceral and reviewed by Kreier and Swaab (Kreier and
nonvisceral. Visceral depots are found intratho- Swaab, 2010). In 1921, the neurologist Klien pres-
racic, intraperitoneal (omental, mesenteric, and ented a case of lipodystrophy (abnormal body fat
umbilical), and extraperitoneal (retroperitoneal, distribution) and suggested a role for the hypothal-
perirenal, gonadal and urogenital). The non- amus in the development of this syndrome. More-
visceral depots are mainly found in muscle tissue over, he hypothesized that dedicated regions of
(intramuscular and perimuscular; a.o. orbital). the hypothalamus would be involved in the accu-
Among white adipocytes, cells from different mulation of fat tissue, while other regions would
depots may have distinct molecular and physiologi- be involved in the mobilization of energy stores
cal properties (Poulos et al., 2010; Rosen and (Klien, 1921). It took nearly a century to provide
MacDougald, 2006). A striking example of this is some neuroanatomical evidence for this concept.
illustrated by the fact that, in humans, accumulation The sympathetic branch of the ANS regulates
of excess fat in the visceral compartment carries lipolysis (Bamshad et al., 1998). However, the pro-
increased metabolic risk, even if subcutaneous fat posed role for the parasympathetic innervation of
mass is normal (Yamamoto et al., 2010). Regional WAT in the storage of lipids is still debated
differences result from variations in molecular cha- (Giordano et al., 2006; Kreier et al., 2006).
racteristics, such as expression of developmental Sympathetic nerve fibers entering WAT mainly
genes, hormone receptor distribution, adipokine innervate blood vessels; only 2–3% of the
expression pattern, and secretory profile, as well as adipocytes receive parenchymal innervation
from depot-specific innervation and vascularization (Slavin and Ballard, 1978). Nevertheless, these
(Fruhbeck, 2008; Turer et al., 2011; Yamamoto sparse nerve-endings might be able to affect many
et al., 2010). For instance, adipocytes in visceral surrounding cells, and thereby potentially account
depots are sensitive to lipolytic stimuli, whereas for a functionally significant contribution to WAT
adipocytes from nonvisceral depots do not release regulation (Bartness et al., 2010a). Neuronal
stored lipids easily (Rosen and MacDougald, tracing methods have been a valuable tool in
2006). However, the proteome of visceral and unraveling neuronal connections between the
subcutaneous WAT in humans displays more brain and WAT. The Bartness lab reported the
differences in the stromavascular cells than in connection from brain to WAT via the sympathetic
isolated adipocytes, suggesting that besides intrinsic nervous system (SNS) in hamsters using pseudo
molecular characteristics in adipocytes, also inner- rabies virus (PRV), which is a neuroinvasive retro-
vation and vascularization, and cellular heterogene- grade viral tracer that can be transported between
ity in WAT are important to explain physiological functionally connected neurons (Bamshad et al.,
differences between depots (Fain et al., 2004). Fur- 1998). Using two PRV strains with a different
thermore, WAT distribution is influenced by age, reporter, Kreier et al. identified separate neuronal
sex, and endocrine determinants such as growth pathways innervating different fat compartments
hormone, cortisol, and sex steroids, leading to an (Fig. 3) (Kreier et al., 2006). These separate path-
apple-shaped android (visceral) fat distribution in ways allow a differential ANS drive to internal
190

(a)
MPO d
Hypothalamus:
c
3
PVN
b
Spinal cord:
IML
SCN
a

CeA

Subcutaneous
Intraabdominal body compartment

(b)
a b

c d e

Fig. 3. Separate neuron sets in the hypothalamus, amygdala, and spinal cord project to different body compartments. (A) Five
microliters of PRV B80 and PRV GFP were injected into parasympathetically denervated intraabdominal adipose tissue and in
subcutaneous adipose tissue. The intermediolateral (IML) cell column of the spinal cord shows a separate control of the
compartments; therefore, the survival time of the animals was chosen so that either only second-order or third-order neurons
were labeled. (B) In an upstream direction, the PVN of the hypothalamus shows specialized sets of neurons projecting to only
one compartment. The same specialization can be seen in the MPO (c), the central biological clock of the hypothalamus (SCN,
d), and the amygdala (e). (Bar in SCN: IML/PVN ¼ 50 mm; MPO/CeA/SCN ¼ 100 mm) (Kreier et al., 2006).
191

and subcutaneous fat compartments. Thus, these glucose uptake (Liu et al., 2004; Yang et al.,
tracing data may provide a neuroanatomical basis 2009). Moreover, pharmacological observations
for depot-specific differences in lipolytic rate in human subcutaneous adipose tissue micro-
among WAT depots, as seen in, for example, dialysis studies showed cholinergic effects on lipol-
Siberian hamsters exposed to short winter-like ysis (Andersson and Arner, 2001). In addition,
days (Bartness, 1995), or in patients treated with functional studies provided evidence for the ana-
pharmacological doses of glucocorticoids (Peckett bolic function of parasympathetic innervation of
et al., 2011). adipose tissue, as euglycemic hyperinsulinemic
Parasympathetic innervation of WAT was clamp studies revealed a > 30% reduction in the
reported after injecting a retrograde PRV tracer insulin-mediated uptake of glucose and FFAs in
in retroperitoneal WAT, in combination with sur- rat adipose tissue after selective removal of its
gical sympathetic denervation of the same WAT parasympathetic input. Furthermore, the activity
depot, inducing intense neuronal labeling in the of the catabolic enzyme hormone sensitive lipase
dorsal motor nucleus of the vagus in rats (HSL) increased by 51% in the denervated adi-
(DMV) (Kreier et al., 2002). The DMV is the pose tissue (Kreier et al., 2002).
main cranial motor nucleus for the vagal nerve, As well as receiving information from the CNS,
which is the major parasympathetic nerve, and WAT also provides neural feedback to the brain
thus neuronal labeling in this nucleus implies via sensory nerves. In 1987, Fishman and Dark
parasympathetic innervation of retroperitoneal showed that application of the anterograde tract
WAT. However, an experiment using chemical tracer True Blue in rat inguinal WAT (iWAT)
denervation in a different fat pad and in another or dorsosubcutaneous WAT stained neurons in
species did not replicate this finding (Giordano the dorsal root ganglia (Fishman and Dark,
et al., 2006), while immunohistochemical markers 1987). Substance P and calcitonin gene-related
of parasympathetic nerves were shown to be peptide were identified as neurotransmitters in
absent in various fat pads of several animal these neurons (Bartness et al., 2010a), and by
models (Giordano et al., 2006). using the H129 strain of herpes simplex virus-1,
In order to confirm the existence of parasympa- a viral anterograde transneuronal tract tracer
thetic innervation of WAT, evidence of the pres- injected into iWAT and eWAT of Siberian
ence of parasympathetic ganglia within the hamsters, Song and colleagues identified the spi-
adipose tissue, the presence of biochemical nal cord and brain sites that receive this sensory
indicators of PNS (peripheral nervous system) input (Song et al., 2009). Earlier, labeling in the
innervation or the presence of neurochemical gracile nucleus—the sensory nucleus in the brain
markers of PNS innervation are needed (Bartness stem involved in the perception of information
et al., 2010a). Such histological or biochemical from the lower body—had been found using chol-
evidence is currently not available. However, sev- era toxin B, which is a monosynaptic anterograde
eral physiological studies suggest a role for the neuronal tracer (Kreier et al., 2006). This obser-
PNS in WAT metabolism. In cultured rat eWAT vation was later confirmed with the H129 antero-
adipocytes, acetylcholine decreased insulin- grade tracer (Song et al., 2009).
stimulated glucose uptake in a dose-dependent The function of these sensory nerves is not fully
manner. This effect was blocked completely understood; however, studies by Song et al. sug-
by atropine (nonselective muscarinic receptor gest that sensory nerves could be informing the
agonist), and partially by 4-diphenylacetoxy- brain about lipid reserves, which might be the
N-methylpiperdine methiodide (4-DAMP), a M3 neural equivalent of the humoral information that
muscarinic receptor antagonist. This suggests a leptin provides to the brain (Bartness et al.,
role for the M3 receptor in insulin-mediated 2010b; Song et al., 2009).
192

Regulation of energy storage termed lipotoxicity, has been implicated in b-cell


loss during the progression of type 2 diabetes,
Lipogenesis and in the pathogenesis of diabetic complications
through loss of cardiomyocytes, hepatocytes,
The adipocyte is unique among cells in that one renal parenchymal cells, and endothelial cells
organelle, the lipid droplet, can encompass over (Brookheart et al., 2009).
95% of the entire cell body. This lipid droplet
serves as a storage vessel for triglycerides pro- Lipolysis
duced by lipogenesis and released through lipoly-
sis (Trujillo and Scherer, 2006). Lipogenesis is Lipolysis is the catabolic process leading to the
the anabolic process by which simple sugars breakdown of triacylglycerols (TAGs) into FFAs
together with glycerol are converted to fats. Sim- and glycerol. After release into the blood, FFAs
ple sugars such as glucose are converted to pyru- are transported and taken up by other tissues to be
vate by glycolysis, and subsequently pyruvate utilized for b-oxidation and subsequent ATP gener-
dehydrogenase converts pyruvate into acetyl- ation. Some FFAs do not leave the fat cell and are
CoA. Following the formation of acetyl-CoA, fatty reesterified into intracellular TAG. During lipolysis,
acid synthesis combines glycerol with three acetyl- intracellular TAG undergoes hydrolysis through
CoAs to form triglycerides, which are transported the action of three major lipases: adipose triglycer-
from the liver as very low-density lipoproteins, to ide lipase (ATGL/desnutrin/phospholipase A2z),
be stored in the adipose tissues. HSL, and monoacylglycerol (MGL) lipase. ATGL
Bar the proposed role of the PNS, lipogenesis is hydrolyses TAGs into diacylglycerol (DAG) and
regulated by insulin. Insulin plasma levels increase one FA, followed by HSL converting DAG into
proportionally with rising blood glucose levels. monoacylglycerol (MAG) plus one FA, MGL then
Thus, high insulin levels are associated with the hydrolyses MAG to produce glycerol and a third
fed state and might therefore be expected to FA (Ahmadian et al., 2010; Lafontan and Langin,
increase the storage of energy. Insulin stimulates 2009).
lipogenesis in two main ways, namely, by (1) the Lipolysis is regulated by the ANS (Bartness
upregulation of pyruvate dehydrogenase and (2) et al., 2010a) and by several humoral factors, such
the dephosphorylation (activation) of acetyl-coA as catecholamines (phosphorylation of HSL),
carboxylase. glucocorticoids (upregulation of ATGL), natri-
Much of what we know about lipogenesis uretic peptides, and growth hormone (Ahmadian
has been elucidated through the study of et al., 2010; Lafontan and Langin, 2009). While
thiazolidinediones (TZDs), a class of antidiabetic systemic regulation of lipolysis has been relatively
drugs that improve insulin sensitivity. TZDs well characterized, much remains to be
increase the expression of PPARs and thereby investigated regarding the local regulation of
promote energy storage. Upregulation of PPARs lipolysis in adipocytes by (autocrine/paracrine)
in the adipocyte results in upregulation of factors. Adipocytes secrete several factors able
lipoprotein lipase, fatty acid transporter protein, to regulate lipolysis locally, such as TNFa, which
adipocyte fatty acid binding protein, malic stimulates lipolysis, and adenosine, which inhibits
enzyme, glucokinase, and the GLUT4 glucose lipolysis (Ahmadian et al., 2010).
transporter (Trujillo and Scherer, 2006).
In obese patients, lipids may be stored ectopi- Circadian rhythms in WAT
cally in nonadipose tissues, including the pan-
creas, heart, liver, kidney, and blood vessel wall. Like most other tissues, WAT gene expression
This process and its deleterious consequences, shows circadian rhythmicity (Ando et al., 2005;
193

Ptitsyn et al., 2006). Although diurnal variations proteins besides cytokines are being produced,
in adipose tissue metabolism are undisputedly the term adipokine now incorporates all proteins
regulated by neurohumoral factors, the circadian secreted from adipose tissue (Stryjecki and
clock within the adipocyte probably plays a signif- Mutch, 2011; Wang et al., 2008).
icant role as well, by altering the sensitivity of the Extensive reviews on the metabolic functions of
adipocyte to specific stimuli throughout the day adipokines can be found in Halberg et al. (2008),
(Bray and Young, 2007). Indeed, several rodent Maury and Brichard (2010), Poulos et al. (2010),
models with alterations in whole body clock gene and Trujillo and Scherer (2006).
expression display disturbed lipid metabolism
(Gimble et al., 2011; Grimaldi et al., 2010; Shimba
et al., 2005). For instance, serum leptin levels Leptin
increased during the light phase in ClockD19
mutant mice fed a regular diet, and this increase Leptin is a hormone secreted by adipose tissue in
was enhanced in mice fed a high-fat diet (Turek proportion to body fat amount and relays fat stor-
et al., 2005). Moreover, both in vitro and in vivo, age information to the brain. High levels of leptin
lack of Bmal1 results in reduced differentiation of signal satiety and reduce food intake, whereas
adipocytes and reduced lipid storage in the low levels of leptin stimulate food intake (Schwartz
adipocytes (Bunger et al., 2005; Kondratov et al., et al., 2000). The discovery that leptin could regu-
2006; Shimba et al., 2005, 2011). ClockD19 mutant late body weight through effects on food intake
mice on a C57BL/6J background have hypercho- and energy expenditure represented a major
lesterolemia, hypertriglyceridemia, hyperglycemia, breakthrough in our understanding of the neuro-
and hypoinsulinemia. However, the severity of anatomical and molecular components of the sys-
the ClockD19 phenotype is dependent on the tems involved in energy homeostasis (Farooqi,
genetic background of the mice (Kennaway et al., 2011). The prevalence of mutations in this system
2012; Oishi et al., 2006). Furthermore, Per2- range from 0.5% to 1% in adult obesity up to
deficient mice display altered lipid metabolism 6% in subjects with severe obesity starting in child-
with a drastic reduction of total TAG and nones- hood (Farooqi, 2007; Larsen et al., 2005). For the
terified fatty acids. Per2 exerts its inhibitory func- discovery of this previously unknown endocrine
tion by blocking PPARy recruitment to its target system, Coleman and Friedman received the
promoters and thereby its transcriptional activa- Albert Lasker Award for Basic Medical Research
tion (Grimaldi et al., 2010). Yet RORa mutant in 2010 (Flier and Maratos-Flier, 2010).
mice resist obesity when placed on a high-fat diet Plasma levels of leptin are regulated by the
(Lau et al., 2008). In sum, these observations dem- biological clock, leading to a clear day/night
onstrate a tight relationship between genes of the rhythm. Furthermore, WAT mass and feeding reg-
core clock mechanism and lipid metabolism. ulate leptin plasma levels; long periods of fasting
WAT plays a central role in the regulation of eliminate the leptin rhythm (Elimam and Marcus,
energy metabolism, mainly via the secretion of 2002). However, under constant and continuous
factors (adipokines) that regulate appetite, food feeding conditions, a circadian rhythm in leptin
intake, glucose disposal, and energy expenditure persists, indicating a role for the circadian clock
(Wang et al., 2008). Adipokines are secreted by in regulating leptin levels during fed conditions
adipocytes and/or the stromavascular fraction of (Kalsbeek et al., 2001; Simon et al., 1998). In
WAT. Originally, the term adipokine described healthy volunteers, misalignment between behav-
cytokines secreted specifically from adipocytes; ior and endogenous circadian timing leads to
however, as many cell types in adipose tissue lower overall leptin levels (Scheer et al., 2009),
have been found to secrete proteins, and other suggesting that leptin responds to the endogenous
194

circadian clock, independent of behavioral factors In addition to being regulated by the clock, lep-
such as feeding. Although SCN lesions eliminate tin serves as an input factor for the biological
leptin circadian rhythmicity (Kalsbeek et al., clock, as the leptin receptor is expressed in SCN
2001), cultured adipocytes still show rhythmic lep- cells, and in vitro leptin can advance the SCN
tin mRNA expression, implying regulation by an (Prosser and Bergeron, 2003).
endogenous clock within the adipocytes (Otway In sum, leptin is a pivotal factor in the interplay
et al., 2009). Figure 4 shows plasma leptin levels between feeding cues, metabolic state, and circa-
in rats under varying conditions. dian timing.

(a) Other adipokines


4
Besides leptin several other adipokines exhibit sig-
3
nificant day/night rhythmicity. Adiponectin is an
adipokine that is involved in glucose and lipid
Leptin (ng/ml)

metabolism by increasing fatty acid oxidation and


2 potentiating insulin-mediated inhibition of hepatic
gluconeogenesis, thus promoting insulin sensitivity
(Barnea et al., 2010). Interestingly, although
1
adiponectin is produced by adipose tissue, its serum
levels and WAT gene expression decrease in obe-
sity and in animals fed a high-fat diet (Barnea
0 6 12 18 24 et al., 2010; Boucher et al., 2005; Turer et al., 2011).
(b) Both in vitro and in vivo, adiponectin has a sig-
8
nificant day/night rhythm (Barnea et al., 2010;
Garaulet et al., 2011; Gavrila et al., 2003; Otway
6 et al., 2009; Scheer et al., 2009, 2010), with a
trough at night for humans, and a trough during
Leptin (ng/ml)

daytime for rats (Oliver et al., 2006; Scheer


4 et al., 2010). In lean men, this rhythm is not
driven by the feeding/fasting cycle (Scheer et al.,
2
2010). ClockD19 mutant mice that retain melato-
nin rhythmicity (ClockD19þ MEL) show increased
eWAT adiponectin gene expression, which may
contribute to the improved insulin resistance
0 6 12 18 24
found in ClockD19þ MEL mice compared to
Zeitgeber time (h)
ClockD19 mice (Kennaway et al., 2012).
Fig. 4. Diurnal plasma leptin profiles in different experimental Resistin is a cytokine that is produced in WAT
conditions. The diurnal plasma leptin profile (mean  SEM) of (adipocytes in rodents, macrophages in human)
nonoperated control animals (n ¼ 7; ○) is compared with that and is a potential mediator of T2DM and cardio-
in SCN-lesioned (n ¼ 9; ; a) or ADXþCort (n ¼ 8; △) and vascular disease (Ando et al., 2005; Fain et al.,
regular-fed (n ¼ 6; ▪) animals (b). Asterisks indicate that 2003; Oliver et al., 2006; Schwartz and Lazar,
plasma leptin values are significantly different (P < 0.01)
from trough values between ZT0 and ZT6. (b) The shaded 2011), with higher expression rates in omental
area in (b) indicates the mean  SEM for the nonoperated versus subcutaneous WAT of obese female
control animals (Kalsbeek et al., 2001). subjects (Fain et al., 2003). Resistin mRNA
195

expression is rhythmic in several WAT com- Due to limited sampling possibilities in human
partments in rats, with a peak in the late dark/ experiments, most of the research on the relation-
early light phase (Oliver et al., 2006). Resistin is ship between obesity and WAT rhythms has been
downregulated by fasting and upregulated by done in animal models. Below, we describe 3 rodent
(re-)feeding (Oliver et al., 2006). However, models with obesity due to altered leptin function.
WAT gene expression levels of resistin are
decreased in obese and high-fat diet-fed mice
(Boucher et al., 2005). Rotating shift workers ob/ob mice
have elevated plasma levels of resistin compared
to day work controls (Burgueno et al., 2010). ob/ob mice exhibit a mutation in the leptin gene,
rendering leptin unable to bind to its receptors.
As a result, these mice become severely obese, with
WAT circadian rhythms in obesity hyperphagia, hyperglycemia, hyperinsulinemia,
high levels of corticosteroids, hypothyroidism,
Obesity and disturbed rhythmicity in WAT are dyslipidemia, decreased body temperature, de-
tightly correlated. For example, in humans, fective thermogenesis, and infertility due to
chronic desynchrony of internal circadian time hypogonadotropic hypogonadism. The broad and
with the external environment (e.g., during severe behavioral, metabolic and neuroendocrine
shiftwork) correlates with increased incidence changes in these animals represent the classic star-
of obesity and type 2 diabetes (Pan et al., vation response designed to protect viability and
2011). Furthermore, circadian misalignment in can all be remedied by leptin administration (Arble
healthy subjects decreases leptin levels (Scheer et al., 2010; Robinson et al., 2000).
et al., 2009). In visceral and subcutaneous The mutated leptin gene in ob/ob mice results
WAT (sWAT) biopsies of severely obese in disturbed sleep and attenuates diurnal and
men, the expression level of clock genes was overall locomotor activity. However, reports are
correlated with waist circumference (Gomez- indecisive on the effect of the mutation on the
Abellan et al., 2008). However, when gene day/night pattern of food intake (Ando et al.,
expression was measured in human sWAT 2011; Laposky et al., 2006). Daily mRNA expres-
explants across a 24-h cycle, no differences in sion rhythms of clock genes in the SCN are unaf-
expression rhythm of clock genes were observed fected; yet the daily rhythms of these genes were
between overweight study participants and type substantially damped down in the liver and
2 diabetes participants or lean controls. The eWAT of ob/ob mice (Ando et al., 2011). Daily
authors suggest this might be due to the early leptin injections modestly but significantly
stage of the disease and the high level of improved the mRNA expression rhythms of sub-
glycemic control in the group of participants. sets of clock genes in liver and eWAT as they
However, in cultured WAT of severely obese did the activity levels during the dark phase. At
women, slight differences between visceral and sub- present, it is unclear whether the increased noc-
cutaneous fat compartments in acrophase and turnal activity is a prerequisite for the improved
amplitude of the expression of several clock and rhythms in liver and eWAT clock genes.
metabolic genes have been found (Garaulet et al.,
2011; Martinez-Agustin et al., 2010). Together,
these results indicate that circadian WAT rhythms db/db mice
may be disturbed in obesity, and that results from
one WAT compartment cannot be downright db/db mice have a mutation in the long form of
extrapolated to another WAT compartment. the leptin receptor, leading to ineffective leptin
196

signaling and thereby to a phenotype similar to Zucker rats, but not in the SCN, mesenteric
the ob/ob mutation. These mice are characterized WAT, and heart. The authors suggest that leptin
by impaired regulation of core components of the may play a role in the regulation of the clock in
clock mechanism in WAT, possibly due to the liver (Motosugi et al., 2011).
impaired AMPK and SIRT1 function (Caton
et al., 2011). Clock, Bmal1 and Per2 mRNA,
Conclusion
and/or protein levels are decreased; treatment
with the antidiabetic drug metformin reversed
The role of the circadian timing system in adi-
these decreases (Caton et al., 2011).
pose biology represents an exciting new field
of study that will give us a greater insight into
the pathogenesis of obesity and its health con-
Obese Zucker rat
sequences. The discovery of the ob/ob gene
and its product leptin, along with the other
A third rodent model manifesting altered leptin
adipokines, revealed an intriguing endocrine sys-
signaling is the Zucker rat. These rats are rela- tem that regulates feeding and metabolism via
tively insensitive to leptin, again due to a muta-
communication between adipose tissue and the
tion in the long form of the leptin receptor and
biological clock in the brain. The monogenetic
consequently have a similar phenotype to ob/ob
mutations in this system leading to obesity have
and db/db mice. As early as 1977, Becker and
taught us valuable lessons on the regulation of
Grinker observed that the normal pattern of pre-
energy metabolism. Moreover, there seems to
dominantly nocturnal feeding was absent in the
be a clear link between circadian misalignment
Zucker rat (Becker and Grinker, 1977). A more
and metabolic disorders, such as obesity and
detailed study 10 years later found no significant type 2 diabetes. In view of the recent development
differences between obese (fa/fa) and lean (Fa/–)
of a 24/7 h society, combined with globalization
Zucker rats in light and dark feeding, expressed
with its increased intertime zone travelling, it will
as a percentage of 24 h intake. The increased food
be increasingly important to expand our knowl-
intake of the Zucker was mainly due to an increase
edge on the health consequences of circadian
in meal size (Alingh et al., 1986; Fukagawa et al.,
misalignment.
1988). Nevertheless, some circadian disturbance
was noted: only in Zucker rats was a significant dif-
Acknowledgments
ference observed between meal sizes in the light
and dark phase, and Zucker rats ate fewer but
We thank Wilma Verweij for correction of the
larger meals during the first half of the dark phase.
manuscript. Rianne van der Spek was financially
Body temperature, activity, and feeding rhythms
supported by the Netherlands Organization for
were phase-advanced in obese Zucker rats, proba-
Scientific Research Foundation (ZonMw
bly because of altered SCN resetting in response
TOP91207036).
to light (Fukagawa et al., 1992; Mistlberger et al.,
1998). All in all, the conclusion after almost 20 years
of research was that, although the circadian ampli- References
tude of daily temperature and activity rhythms in
obese Zucker is depressed, obese rats do exhibit Ahmadian, M., Wang, Y., & Sul, H. S. (2010). Lipolysis in
normal entrainment and pacemaker functions in adipocytes. The International Journal of Biochemistry & Cell
Biology, 42(5), 555–559.
the circadian timing system (Murakami et al., 1995).
Alingh, P. A., de Jong-Nagelsmit, A., Keijser, J., &
Moreover, more recently, it was shown that, Strubbe, J. H. (1986). Daily rhythms of feeding in the genet-
compared to control animals, clock gene expres- ically obese and lean Zucker rats. Physiology & Behavior,
sion rhythms were damped down in the liver of 38(3), 423–426.
197

Andersson, K., & Arner, P. (2001). Systemic nicotine Buijs, R. M., la Fleur, S. E., Wortel, J., Van, H. C.,
stimulates human adipose tissue lipolysis through local cho- Zuiddam, L., Mettenleiter, T. C., et al. (2003). The
linergic and catecholaminergic receptors. International Jour- suprachiasmatic nucleus balances sympathetic and parasym-
nal of Obesity and Related Metabolic Disorders, 25(8), pathetic output to peripheral organs through separate pre-
1225–1232. autonomic neurons. The Journal of Comparative
Ando, H., Kumazaki, M., Motosugi, Y., Ushijima, K., Neurology, 464(1), 36–48.
Maekawa, T., Ishikawa, E., et al. (2011). Impairment of Bunger, M. K., Walisser, J. A., Sullivan, R., Manley, P. A.,
peripheral circadian clocks precedes metabolic abnormalities Moran, S. M., Kalscheur, V. L., et al. (2005). Progressive
in ob/ob mice. Endocrinology, 152(4), 1347–1354. arthropathy in mice with a targeted disruption of the
Ando, H., Yanagihara, H., Hayashi, Y., Obi, Y., Tsuruoka, S., Mop3/Bmal-1 locus. Genesis, 41(3), 122–132.
Takamura, T., et al. (2005). Rhythmic messenger ribonucleic Burgueno, A., Gemma, C., Gianotti, T. F., Sookoian, S., &
acid expression of clock genes and adipocytokines in mouse Pirola, C. J. (2010). Increased levels of resistin in rotating
visceral adipose tissue. Endocrinology, 146(12), 5631–5636. shift workers: A potential mediator of cardiovascular risk
Arble, D. M., Ramsey, K. M., Bass, J., & Turek, F. W. (2010). associated with circadian misalignment. Atherosclerosis,
Circadian disruption and metabolic disease: Findings from 210(2), 625–629.
animal models. Best Practice & Research. Clinical Endocri- Burris, T. P. (2008). Nuclear hormone receptors for heme:
nology & Metabolism, 24(5), 785–800. REV-ERBalpha and REV-ERBbeta are ligand-regulated
Asher, G., & Schibler, U. (2011). Crosstalk between compo- components of the mammalian clock. Molecular Endocrinol-
nents of circadian and metabolic cycles in mammals. Cell ogy, 22(7), 1509–1520.
Metabolism, 13(2), 125–137. Cailotto, C., la Fleur, S. E., Van, H. C., Wortel, J.,
Balsalobre, A., Brown, S. A., Marcacci, L., Tronche, F., Kalsbeek, A., Feenstra, M., et al. (2005). The
Kellendonk, C., Reichardt, H. M., et al. (2000). Resetting suprachiasmatic nucleus controls the daily variation of
of circadian time in peripheral tissues by glucocorticoid sig- plasma glucose via the autonomic output to the liver: Are
naling. Science, 289(5488), 2344–2347. the clock genes involved? The European Journal of Neuro-
Bamshad, M., Aoki, V. T., Adkison, M. G., Warren, W. S., & science, 22(10), 2531–2540.
Bartness, T. J. (1998). Central nervous system origins of the Cailotto, C., Lei, J., van der Vliet, J., van Heijningen, C.,
sympathetic nervous system outflow to white adipose tissue. van Eden, C. G., Kalsbeek, A., et al. (2009). Effects of noc-
The American Journal of Physiology, 275(1 Pt 2), turnal light on (clock) gene expression in peripheral organs:
R291–R299. A role for the autonomic innervation of the liver. PLoS
Barnea, M., Madar, Z., & Froy, O. (2010). High-fat diet One, 4(5), e5650.
followed by fasting disrupts circadian expression of Cardone, L., Hirayama, J., Giordano, F., Tamaru, T.,
adiponectin signaling pathway in muscle and adipose tissue. Palvimo, J. J., & Sassone-Corsi, P. (2005). Circadian clock
Obesity (Silver Spring), 18(2), 230–238. control by SUMOylation of BMAL1. Science, 309(5739),
Bartness, T. J. (1995). Short day-induced depletion of lipid 1390–1394.
stores is fat pad- and gender-specific in Siberian hamsters. Casteilla, L., Penicaud, L., Cousin, B., & Calise, D. (2008).
Physiology & Behavior, 58(3), 539–550. Choosing an adipose tissue depot for sampling: Factors in
Bartness, T. J., Shrestha, Y. B., Vaughan, C. H., selection and depot specificity. Methods in Molecular Biol-
Schwartz, G. J., & Song, C. K. (2010). Sensory and sympa- ogy, 456, 23–38.
thetic nervous system control of white adipose tissue lipoly- Caton, P. W., Kieswich, J., Yaqoob, M. M., Holness, M. J., &
sis. Molecular and Cellular Endocrinology, 318(1–2), 34–43. Sugden, M. C. (2011). Metformin opposes impaired AMPK
Bartness, T. J., Vaughan, C. H., & Song, C. K. (2010). Sympa- and SIRT1 function and deleterious changes in core clock pro-
thetic and sensory innervation of brown adipose tissue. tein expression in white adipose tissue of genetically-obese db/
International Journal of Obesity, 34(Suppl. 1), S36–S42. db mice. Diabetes, Obesity & Metabolism, 13(12), 1097–1104.
Becker, E. E., & Grinker, J. A. (1977). Meal patterns in the Christodoulides, C., Lagathu, C., Sethi, J. K., & Vidal-Puig, A.
genetically obese Zucker rat. Physiology & Behavior, 18 (2009). Adipogenesis and WNT signalling. Trends in Endo-
(4), 685–692. crinology and Metabolism, 20(1), 16–24.
Boucher, J., Castan-Laurell, I., Daviaud, D., Guigne, C., Dibner, C., Schibler, U., & Albrecht, U. (2010). The mamma-
Buleon, M., Carpene, C., et al. (2005). Adipokine expres- lian circadian timing system: Organization and coordination
sion profile in adipocytes of different mouse models of obe- of central and peripheral clocks. Annual Review of Physiol-
sity. Hormone and Metabolic Research, 37(12), 761–767. ogy, 72, 517–549.
Bray, M. S., & Young, M. E. (2007). Circadian rhythms in the Elbers, J. M., Asscheman, H., Seidell, J. C., & Gooren, L. J.
development of obesity: Potential role for the circadian (1999). Effects of sex steroid hormones on regional fat
clock within the adipocyte. Obesity Reviews, 8(2), 169–181. depots as assessed by magnetic resonance imaging in trans-
Brookheart, R. T., Michel, C. I., & Schaffer, J. E. (2009). As a sexuals. The American Journal of Physiology, 276(2 Pt 1),
matter of fat. Cell Metabolism, 10(1), 9–12. E317–E325.
198

Elimam, A., & Marcus, C. (2002). Meal timing, fasting and Gesta, S., Tseng, Y. H., & Kahn, C. R. (2007). Developmental
glucocorticoids interplay in serum leptin concentrations origin of fat: Tracking obesity to its source. Cell, 131(2),
and diurnal profile. European Journal of Endocrinology, 242–256.
147(2), 181–188. Gimble, J. M., Sutton, G. M., Bunnell, B. A., Ptitsyn, A. A., &
Engeland, W. C., & Arnhold, M. M. (2005). Neural circuitry in Floyd, Z. E. (2011). Prospective influences of circadian
the regulation of adrenal corticosterone rhythmicity. Endo- clocks in adipose tissue and metabolism. Nature Reviews.
crine, 28(3), 325–332. Endocrinology, 7(2), 98–107.
Fain, J. N., Cheema, P. S., Bahouth, S. W., & Lloyd, H. M. Giordano, A., Song, C. K., Bowers, R. R., Ehlen, J. C.,
(2003). Resistin release by human adipose tissue explants Frontini, A., Cinti, S., et al. (2006). White adipose tissue
in primary culture. Biochemical and Biophysical Research lacks significant vagal innervation and immunohistochemical
Communications, 300(3), 674–678. evidence of parasympathetic innervation. American Journal
Fain, J. N., Madan, A. K., Hiler, M. L., Cheema, P., & of Physiology. Regulatory, Integrative and Comparative
Bahouth, S. W. (2004). Comparison of the release of Physiology, 291(5), R1243–R1255.
adipokines by adipose tissue, adipose tissue matrix, and Gomez-Abellan, P., Hernandez-Morante, J. J., Lujan, J. A.,
adipocytes from visceral and subcutaneous abdominal adipose Madrid, J. A., & Garaulet, M. (2008). Clock genes are
tissues of obese humans. Endocrinology, 145(5), 2273–2282. implicated in the human metabolic syndrome. International
Farooqi, I. S. (2011). Genetic, molecular and physiological Journal of Obesity, 32, 121–128.
insights into human obesity. European Journal of Clinical Grimaldi, B., Bellet, M. M., Katada, S., Astarita, G.,
Investigation, 41(4), 451–455. Hirayama, J., Amin, R. H., et al. (2010). PER2 controls lipid
Farooqi, S. (2007). Insights from the genetics of severe child- metabolism by direct regulation of PPARgamma. Cell
hood obesity. Hormone Research, 68(Suppl. 5), 5–7. Metabolism, 12(5), 509–520.
Fishman, R. B., & Dark, J. (1987). Sensory innervation of Guo, H., Brewer, J. M., Champhekar, A., Harris, R. B., &
white adipose tissue. The American Journal of Physiology, Bittman, E. L. (2005). Differential control of peripheral cir-
253(6 Pt 2), R942–R944. cadian rhythms by suprachiasmatic-dependent neural
Flier, J. S., & Maratos-Flier, E. (2010). Lasker lauds leptin. signals. Proceedings of the National Academy of Sciences of
Cell Metabolism, 12(4), 317–320. the United States of America, 102(8), 3111–3116.
Fruhbeck, G. (2008). Overview of adipose tissue and its role in Halberg, N., Wernstedt-Asterholm, I., & Scherer, P. E. (2008).
obesity and metabolic disorders. Methods in Molecular Biol- The adipocyte as an endocrine cell. Endocrinology and
ogy, 456, 1–22. Metabolism Clinics of North America, 37(3), 753xi–768xi.
Fukagawa, K., Sakata, T., Yoshimatsu, H., Fujimoto, K., & Hut, R. A., & Beersma, D. G. (2011). Evolution of time-
Shiraishi, T. (1988). Disruption of light-dark cycle of feeding keeping mechanisms: Early emergence and adaptation to
and drinking behavior, and ambulatory activity induced by photoperiod. Philosophical Transactions of the Royal Soci-
development of obesity in the Zucker rat. International ety of London. Series B, Biological Sciences, 366(1574),
Journal of Obesity, 12(5), 481–490. 2141–2154.
Fukagawa, K., Sakata, T., Yoshimatsu, H., Fujimoto, K., Ishida, A., Mutoh, T., Ueyama, T., Bando, H., Masubuchi, S.,
Uchimura, K., & Asano, C. (1992). Advance shift of feeding Nakahara, D., et al. (2005). Light activates the adrenal
circadian rhythm induced by obesity progression in Zucker gland: Timing of gene expression and glucocorticoid release.
rats. The American Journal of Physiology, 263(6 Pt 2), Cell Metabolism, 2(5), 297–307.
R1169–R1175. Kalsbeek, A., Fliers, E., Romijn, J. A., la Fleur, S. E.,
Gallego, M., & Virshup, D. M. (2007). Post-translational Wortel, J., Bakker, O., et al. (2001). The suprachiasmatic
modifications regulate the ticking of the circadian clock. nucleus generates the diurnal changes in plasma leptin
Nature Reviews. Molecular Cell Biology, 8(2), 139–148. levels. Endocrinology, 142(6), 2677–2685.
Garaulet, M., Ordovas, J. M., Gomez-Abellan, P., Kalsbeek, A., van der Spek, R., Lei, J., Endert, E.,
Martinez, J. A., & Madrid, J. A. (2011). An approximation Buijs, R. M., & Fliers, E. (2012). Circadian rhythms in the
to the temporal order in endogenous circadian rhythms of hypothalamo-pituitary-adrenal (HPA) axis. Molecular and
genes implicated in human adipose tissue metabolism. Jour- Cellular Endocrinology, 349(1), 20–29.
nal of Cellular Physiology, 226(8), 2075–2080. Kennaway, D. J., Owens, J. A., Voultsios, A., & Wight, N.
Gavrila, A., Peng, C. K., Chan, J. L., Mietus, J. E., (2012). Adipokines and adipocyte function in clock mutant
Goldberger, A. L., & Mantzoros, C. S. (2003). Diurnal and mice that retain melatonin rhythmicity. Obesity (Silver
ultradian dynamics of serum adiponectin in healthy men: Spring), 20(2), 295–305.
Comparison with leptin, circulating soluble leptin receptor, Klien, H. (1921). Über Lipodystrophie nebst Mitteilung
and cortisol patterns. The Journal of Clinical Endocrinology eines Falles. Münchener Medizinische Wochenschrift, 7,
and Metabolism, 88(6), 2838–2843. 206–208.
199

Ko, C. H., & Takahashi, J. S. (2006). Molecular components of Maury, E., & Brichard, S. M. (2010). Adipokine dysregulation,
the mammalian circadian clock. Human Molecular Genetics, adipose tissue inflammation and metabolic syndrome.
15(Spec No. 2), R271–R277. Molecular and Cellular Endocrinology, 314(1), 1–16.
Kondratov, R. V., Kondratova, A. A., Gorbacheva, V. Y., Maury, E., Ramsey, K. M., & Bass, J. (2010). Circadian
Vykhovanets, O. V., & Antoch, M. P. (2006). Early aging rhythms and metabolic syndrome: From experimental
and age-related pathologies in mice deficient in BMAL1, genetics to human disease. Circulation Research, 106(3),
the core component of the circadian clock. Genes & Devel- 447–462.
opment, 20(14), 1868–1873. Merrow, M., & Roenneberg, T. (2001). Circadian clocks: Run-
Kreier, F., Fliers, E., Voshol, P. J., van Eden, C. G., ning on redox. Cell, 106(2), 141–143.
Havekes, L. M., Kalsbeek, A., et al. (2002). Selective para- Mistlberger, R. E., Lukman, H., & Nadeau, B. G. (1998). Cir-
sympathetic innervation of subcutaneous and intra-abdominal cadian rhythms in the Zucker obese rat: Assessment and
fat–functional implications. The Journal of Clinical Investiga- intervention. Appetite, 30(3), 255–267.
tion, 110(9), 1243–1250. Moore, R. Y. (1996). Neural control of the pineal gland.
Kreier, F., Kap, Y. S., Mettenleiter, T. C., van Heijningen, C., Behavioural Brain Research, 73(1–2), 125–130.
van der Vilet, J., Kalsbeek, A., et al. (2006). Tracing from Motosugi, Y., Ando, H., Ushijima, K., Maekawa, T.,
fat tissue, liver, and pancreas: A neuroanatomical frame- Ishikawa, E., Kumazaki, M., et al. (2011). Tissue-dependent
work for the role of the brain in type 2 diabetes. Endocrinol- alterations of the clock gene expression rhythms in leptin-
ogy, 147(3), 1140–1147. resistant Zucker diabetic fatty rats. Chronobiology Interna-
Kreier, F., & Swaab, D. F. (2010). Chapter 23: History of tional, 28(10), 968–972.
neuroendocrinology “the spring of primitive existence” Murakami, D. M., Horwitz, B. A., & Fuller, C. A. (1995). Cir-
Handbook of Clinical Neurology, 95, 335–360. cadian rhythms of temperature and activity in obese and
Lafontan, M., & Langin, D. (2009). Lipolysis and lipid mobili- lean Zucker rats. The American Journal of Physiology, 269
zation in human adipose tissue. Progress in Lipid Research, (5 Pt 2), R1038–R1043.
48(5), 275–297. O’Neill, J. S., Maywood, E. S., Chesham, J. E., Takahashi, J. S.,
Laposky, A. D., Shelton, J., Bass, J., Dugovic, C., Perrino, N., & & Hastings, M. H. (2008). cAMP-dependent signaling as a
Turek, F. W. (2006). Altered sleep regulation in leptin- core component of the mammalian circadian pacemaker.
deficient mice. American Journal of Physiology. Regulatory, Science, 320(5878), 949–953.
Integrative and Comparative Physiology, 290(4), R894–R903. O’Neill, J. S., & Reddy, A. B. (2011). Circadian clocks in
Larsen, L. H., Echwald, S. M., Sorensen, T. I., Andersen, T., human red blood cells. Nature, 469(7331), 498–503.
Wulff, B. S., & Pedersen, O. (2005). Prevalence of mutations Oishi, K., Atsumi, G., Sugiyama, S., Kodomari, I.,
and functional analyses of melanocortin 4 receptor variants Kasamatsu, M., Machida, K., et al. (2006). Disrupted fat
identified among 750 men with juvenile-onset obesity. The absorption attenuates obesity induced by a high-fat diet in
Journal of Clinical Endocrinology and Metabolism, 90(1), Clock mutant mice. FEBS Letters, 580(1), 127–130.
219–224. Oliver, P., Ribot, J., Rodriguez, A. M., Sanchez, J., Pico, C., &
Lau, P., Fitzsimmons, R. L., Raichur, S., Wang, S. C., Palou, A. (2006). Resistin as a putative modulator of insulin
Lechtken, A., & Muscat, G. E. (2008). The orphan nuclear action in the daily feeding/fasting rhythm. Pflügers Archiv,
receptor, RORalpha, regulates gene expression that con- 452(3), 260–267.
trols lipid metabolism: Staggerer (SG/SG) mice are resistant Otway, D. T., Frost, G., & Johnston, J. D. (2009). Circadian
to diet-induced obesity. The Journal of Biological Chemistry, rhythmicity in murine pre-adipocyte and adipocyte cells.
283(26), 18411–18421. Chronobiology International, 26(7), 1340–1354.
Lilley, T. R., Wotus, C., Taylor, D., Lee, J. M., & de la Pan, A., Schernhammer, E. S., Sun, Q., & Hu, F. B. (2011).
Iglesia, H. O. (2012). Circadian regulation of cortisol release Rotating night shift work and risk of type 2 diabetes: Two
in behaviorally split golden hamsters. Endocrinology, 153 prospective cohort studies in women. PLoS Medicine, 8
(2), 732–738. (12), e1001141.
Liu, R. H., Mizuta, M., & Matsukura, S. (2004). The expres- Peckett, A. J., Wright, D. C., & Riddell, M. C. (2011). The
sion and functional role of nicotinic acetylcholine receptors effects of glucocorticoids on adipose tissue lipid metabolism.
in rat adipocytes. The Journal of Pharmacology and Experi- Metabolism, 60(11), 1500–1510.
mental Therapeutics, 310(1), 52–58. Poulos, S. P., Hausman, D. B., & Hausman, G. J. (2010). The
Martinez-Agustin, O., Hernandez-Morante, J. J., Martinez- development and endocrine functions of adipose tissue.
Plata, E., Sanchez de Medina, F., & Garaulet, M. (2010). Molecular and Cellular Endocrinology, 323(1), 20–34.
Differences in AMPK expression between subcutaneous Prosser, R. A., & Bergeron, H. E. (2003). Leptin phase-
and visceral adipose tissue in morbid obesity. Regulatory advances the rat suprachiasmatic circadian clock in vitro.
Peptides, 163(1–3), 31–36. Neuroscience Letters, 336(3), 139–142.
200

Ptitsyn, A. A., Zvonic, S., Conrad, S. A., Scott, L. K., Simon, C., Gronfier, C., Schlienger, J. L., &
Mynatt, R. L., & Gimble, J. M. (2006). Circadian clocks Brandenberger, G. (1998). Circadian and ultradian
are resounding in peripheral tissues. PLoS Computational variations of leptin in normal man under continuous enteral
Biology, 2(3), e16. nutrition: Relationship to sleep and body temperature. The
Reddy, A. B., Karp, N. A., Maywood, E. S., Sage, E. A., Journal of Clinical Endocrinology and Metabolism, 83(6),
Deery, M., O’Neill, J. S., et al. (2006). Circadian orchestration 1893–1899.
of the hepatic proteome. Current Biology, 16(11), 1107–1115. Slavin, B. G., & Ballard, K. W. (1978). Morphological studies
Reddy, A. B., & O’Neill, J. S. (2011). Metaclocks. EMBO on the adrenergic innervation of white adipose tissue. The
Reports, 12(7), 612. Anatomical Record, 191(3), 377–389.
Reilly, D. F., Curtis, A. M., Cheng, Y., Westgate, E. J., So, A. Y., Bernal, T. U., Pillsbury, M. L., Yamamoto, K. R., &
Rudic, R. D., Paschos, G., et al. (2008). Peripheral circadian Feldman, B. J. (2009). Glucocorticoid regulation of the cir-
clock rhythmicity is retained in the absence of adrenergic cadian clock modulates glucose homeostasis. Proceedings
signaling. Arteriosclerosis, Thrombosis, and Vascular Biol- of the National Academy of Sciences of the United States of
ogy, 28(1), 121–126. America, 106(41), 17582–17587.
Reppert, S. M., & Weaver, D. R. (2001). Molecular analysis of Song, C. K., Schwartz, G. J., & Bartness, T. J. (2009). Antero-
mammalian circadian rhythms. Annual Review of Physiol- grade transneuronal viral tract tracing reveals central sen-
ogy, 63, 647–676. sory circuits from white adipose tissue. American Journal
Robinson, S. W., Dinulescu, D. M., & Cone, R. D. (2000). of Physiology. Regulatory, Integrative and Comparative
Genetic models of obesity and energy balance in the mouse. Physiology, 296(3), R501–R511.
Annual Review of Genetics, 34, 687–745. Stryjecki, C., & Mutch, D. M. (2011). Fatty acid-gene
Rosen, E. D., & MacDougald, O. A. (2006). Adipocyte differ- interactions, adipokines and obesity. European Journal of
entiation from the inside out. Nature Reviews. Molecular Clinical Nutrition, 65(3), 285–297.
Cell Biology, 7(12), 885–896. Trujillo, M. E., & Scherer, P. E. (2006). Adipose tissue-derived
Rutter, J., Reick, M., Wu, L. C., & McKnight, S. L. (2001). factors: Impact on health and disease. Endocrine Reviews,
Regulation of clock and NPAS2 DNA binding by the redox 27(7), 762–778.
state of NAD cofactors. Science, 293(5529), 510–514. Turek, F. W., Joshu, C., Kohsaka, A., Lin, E., Ivanova, G.,
Scheer, F. A., Chan, J. L., Fargnoli, J., Chamberland, J., McDearmon, E., et al. (2005). Obesity and metabolic syn-
Arampatzi, K., Shea, S. A., et al. (2010). Day/night drome in circadian Clock mutant mice. Science, 308(5724),
variations of high-molecular-weight adiponectin and 1043–1045.
lipocalin-2 in healthy men studied under fed and fasted Turer, A. T., Khera, A., Ayers, C. R., Turer, C. B.,
conditions. Diabetologia, 53(11), 2401–2405. Grundy, S. M., Vega, G. L., et al. (2011). Adipose tissue
Scheer, F. A., Hilton, M. F., Mantzoros, C. S., & Shea, S. A. mass and location affect circulating adiponectin levels.
(2009). Adverse metabolic and cardiovascular consequences Diabetologia, 54(10), 2515–2524.
of circadian misalignment. Proceedings of the National Um, J. H., Pendergast, J. S., Springer, D. A., Foretz, M.,
Academy of Sciences of the United States of America, 106 Viollet, B., Brown, A., et al. (2011). AMPK regulates circa-
(11), 4453–4458. dian rhythms in a tissue- and isoform-specific manner. PLoS
Schwartz, D. R., & Lazar, M. A. (2011). Human resistin: One, 6(3), e18450.
Found in translation from mouse to man. Trends in Endocri- Vujovic, N., Davidson, A. J., & Menaker, M. (2008). Sympa-
nology and Metabolism, 22(7), 259–265. thetic input modulates, but does not determine, phase of
Schwartz, M. W., Woods, S. C., Porte, D., Jr., Seeley, R. J., & peripheral circadian oscillators. American Journal of Physi-
Baskin, D. G. (2000). Central nervous system control of ology. Regulatory, Integrative and Comparative Physiology,
food intake. Nature, 404(6778), 661–671. 295(1), R355–R360.
Shimba, S., Ishii, N., Ohta, Y., Ohno, T., Watabe, Y., Wang, P., Mariman, E., Renes, J., & Keijer, J. (2008). The
Hayashi, M., et al. (2005). Brain and muscle Arnt-like pro- secretory function of adipocytes in the physiology of
tein-1 (BMAL1), a component of the molecular clock, white adipose tissue. Journal of Cellular Physiology, 216
regulates adipogenesis. Proceedings of the National Acad- (1), 3–13.
emy of Sciences of the United States of America, 102(34), Watts, A. G., & Swanson, L. W. (1987). Efferent projections of
12071–12076. the suprachiasmatic nucleus: II. Studies using retrograde
Shimba, S., Ogawa, T., Hitosugi, S., Ichihashi, Y., transport of fluorescent dyes and simultaneous peptide
Nakadaira, Y., Kobayashi, M., et al. (2011). Deficient of a immunohistochemistry in the rat. The Journal of Compara-
clock gene, brain and muscle Arnt-like protein-1 (BMAL1), tive Neurology, 258(2), 230–252.
induces dyslipidemia and ectopic fat formation. PLoS One, Watts, A. G., Swanson, L. W., & Sanchez-Watts, G. (1987).
6(9), e25231. Efferent projections of the suprachiasmatic nucleus: I. Studies
201

using anterograde transport of Phaseolus vulgaris Yamamoto, T., Nakahata, Y., Tanaka, M., Yoshida, M.,
leucoagglutinin in the rat. The Journal of Comparative Neu- Soma, H., Shinohara, K., et al. (2005). Acute physical stress
rology, 258(2), 204–229. elevates mouse period1 mRNA expression in mouse periph-
World Health Organisation, (2011). WHO Fact sheet N311— eral tissues via a glucocorticoid-responsive element. The
Obesity and overweight.. . Journal of Biological Chemistry, 280(51), 42036–42043.
Yamamoto, Y., Gesta, S., Lee, K. Y., Tran, T. T., Saadatirad, P., Yang, T. T., Chang, C. K., Tsao, C. W., Hsu, Y. M., Hsu, C. T.,
& Kahn, C. R. (2010). Adipose depots possess unique devel- & Cheng, J. T. (2009). Activation of muscarinic M-3 recep-
opmental gene signatures. Obesity (Silver Spring), 18(5), tor may decrease glucose uptake and lipolysis in adipose tis-
872–878. sue of rats. Neuroscience Letters, 451(1), 57–59.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 12

Circadian modulation of sleep in rodents

Roman Yasenkov1 and Tom Deboer*

Laboratory for Neurophysiology, Department of Molecular Cell Biology, Leiden University Medical Center,
Leiden, The Netherlands

Abstract: Sleep is regulated by circadian and homeostatic processes. The sleep homeostat keeps track of
the duration of prior sleep and waking and determines the intensity of sleep. In mammals, the homeostatic
process is reflected by the slow waves in the non-rapid eye movement (NREM) sleep electroencephalogram
(EEG). The circadian process is controlled by a pacemaker located in the suprachiasmatic nucleus of the
hypothalamus and provides the sleep homeostat with a circadian framework.
This review summarizes the changes in sleep obtained after different chronobiological interventions
(changes in photoperiod, light availability, and running wheel availability), the influence of mutations
or lesions in clock genes on sleep, and research on the interaction between sleep homeostasis and the
circadian clock.
Research in humans shows that the period of consolidated waking during the day is a consequence of
the interaction between an increasing homeostatic sleep drive and a circadian signal, which promotes
waking during the day and sleep during the night. In the rat, it was shown that, under constant
homeostatic sleep pressure, with similar levels of slow waves in the NREM sleep EEG at all time
points of the circadian cycle, still a small circadian modulation of the duration of waking and NREM
sleep episodes was observed. Under similar conditions, humans show a clear circadian modulation in
REM sleep, whereas in the rat, a circadian modulation in REM sleep was not present. Therefore, in
the rat, the sleep homeostatic modulation in phase with the circadian clock seems to amplify the
relatively weak circadian changes in sleep induced by the circadian clock.
Knowledge about the interaction between sleep and the circadian clock and the circadian modulation of
sleep in other species than humans is important to better understand the underlying regulatory mechanisms.
Keywords: circadian; electroencephalogram; hamster; mouse; rat; rodent; photoperiod; REM sleep; SCN
lesion; sleep; sleep homeostasis.

*Corresponding author.
Tel.: þ31 71 526 9771; Fax: þ31 71 526 8270
E-mail: tom.de_boer@lumc.nl

1
Present address: Faculty of Health and Medical Sciences,
University of Surrey, Guildford, Surrey, United Kingdom.

http://dx.doi.org/10.1016/B978-0-444-59427-3.00012-5 203
204

Introduction and the (C57BL/6) mouse were recorded by sev-


eral more authors than listed here. In Table 1,
Sleep is regulated by homeostatic and circadian the “doubles” are included which are also
processes (Achermann and Borbely, 2011). discussed in Tables 2 and 3.
In mammals, homeostatic sleep pressure is Many rodent species recorded in the laboratory
reflected by electroencephalogram (EEG) slow- are night active and therefore have their
wave activity (SWA, EEG power density main sleep phase during the light period (rat,
between  1 and 4 Hz) in undisturbed non-rapid mouse, Syrian hamster, Djungarian hamster, chin-
eye movement (NREM) sleep. In all mammalian chilla), others are day active (13-lined ground
species investigated until now, SWA increases squirrel, European ground squirrel, Siberian chip-
after sleep deprivation, and in several species, a munk, golden-mantled ground squirrel), and a
dose–response relationship between waking dura- few species show crepuscular or cathemeral
tion and subsequent SWA was established behavior, with no preference for activity in the
(Deboer and Tobler, 2003; Dijk et al., 1987; light or dark period. The latter is confirmed by
Huber et al., 2000a; Lancel et al., 1991; Strijkstra the ratio in total sleep time (TST) between light
and Daan, 1998; Tobler and Borbely, 1986). and dark being close to 1 and the percentage of
A study involving an afternoon nap showed that TST in the rest period being close to 50 (naked
also predictable decreases in sleep homeostasis mole rat, guinea pig, degu, Mongolian gerbil).
can be observed in the NREM sleep EEG SWA These sleep patterns are relatively fixed when
(Werth et al., 1996b). Mathematical models, the animals are left undisturbed in the laboratory
simulating the homeostatic response, have been (but see Table 3 for some exceptions), whereas in
applied successfully in humans (Achermann the wild, rest–activity patterns seem to be more
et al., 1993), rats (Franken et al., 1991), and mice flexible and may be very dependent on environ-
(Franken et al., 2001; Huber et al., 2000a). mental influences (Daan et al., 2011; Gattermann
The circadian process is controlled by an endoge- et al., 2008; Levy et al., 2007).
nous pacemaker, which is located in the In contrast to humans, rodents show a strong
suprachiasmatic nucleus (SCN) of the hypothala- ultradian component in their sleep, and even in
mus (Meijer and Rietveld, 1989). This pacemaker species with a clear preference for activity in the
is thought to provide the sleep homeostat with a light or dark phase, no single period of con-
circadian framework. solidated waking is seen during the active period
or consolidated sleep during the rest period.
Despite these differences, sleep homeostatic the-
Normal sleep in rodents ory, as it predicts the depth of sleep, seems to
apply just as well to polyphasic sleep in rodents
Rodents show a large variety of sleep patterns. as it does to monophasic sleep in humans. How-
In Table 1, all rodent species are listed in which ever, due to the polyphasic nature of their sleep,
at least one full 24 h cycle of EEG and EMG rodents are more flexible in their response to
was recorded under a 12 h:12 h L:D cycle environmental influences on sleep. This may
(Table 1). We do not claim completeness and in explain the relatively large differences in 24-h
addition realize that, particularly in mice, a broad TST found in the same species between different
range of strains and knockout models is available laboratories in the Sprague–Dawley rat, Syrian
that could also have been included. For most spe- hamster, and C57BL/6 mice (Table 1). The range
cies, only one publication is available, but for of TST in the C57BL/6 mouse is between 40.6%
some species (mouse, rat, and hamster), several and 48.6%, which is a difference of almost 2 h
studies have been published and not all of them over 24 h. Since these mice are supposed to be
are covered in this review. Particularly, the rat genetically identical, this indicates that the
205

Table 1. List of rodent species and their electroencephalographic confirmed total sleep time under 12 h:12 h light:dark conditions

% TST % TSTL % TSTD Ratio % TST in rest period References

Diurnal species
Spermophilus citellus
European ground squirrel 57.5 46.6 68.3 1.47 59.4 Strijkstra (1999)
Eutamias sibericus
Siberian chipmunk 52.8 29.1 74.7 2.57 70.7 Dijk and Daan (1989)
Citellus lateralis
Golden-mantled ground
squirrel 60.6 40.9 80.4 1.97 66.3 Haskell et al. (1979)
Citellus tridecemlineatus
Thirteen-lined ground squirrel 57.5 54.8 60.2 1.10 52.3 Van Twyver (1969)a
Nocturnal species
Mesocricetus auratus
Syrian hamster 64.7 78.2 51.2 1.53 60.5 Tobler and Jaggi (1987)
60.1 76.0 44.2 1.72 63.2 Van Twyver (1969)a
Mus musculus
Mouse CRI 54.8 64.5 45.1 1.43 58.9 Van Twyver (1969)a
Mouse C57BL/6 48.6 64.1 33.1 1.94 65.9 Huber et al. (2000a)
40.6 56.4 24.7 2.28 69.5 Franken et al. (1999)
46.6 63.4 29.8 2.13 68.0 Deboer et al. (2007b)
Rattus norvegicus
Sprague–Dawley rat 47.9 68.9 26.9 2.56 71.9 Franken et al. (1995)
49.2 68.0 30.4 2.24 69.1 Tobler et al. (1994)
57.1 78.1 36.0 2.17 68.4 Mistlberger et al. (1983)
Long Evans rat 55.2 65.7 44.7 1.47 59.5 Van Twyver (1969)a
Chinchilla langer
Chinchilla 52.2 57.6 46.8 1.23 55.2 Van Twyver (1969)a
No clear diurnal preference
Cavia porcellus
Guinea pig 32.2 35.7 28.9 1.24 55.4 Tobler and Franken (1993)
Octodon degus
Degus 37.6 36.7 38.5 1.05 48.8 Kas and Edgar (1998)
Meriones unguiculatus
Mongolian gerbil 54.5 57.1 51.8 1.10 52.4 Susic and Masirevic (1986)
Spalax ehrenbergi
Mole rat 51.8 48.8 54.7 1.12 52.8 Tobler and Deboer (2001)

TST, total sleep time over 24 h; TSTL, total sleep time during the light period; TSTD, total sleep time during the dark period; Ratio, ratio between
TSTL and TSTD, a measure for diurnality or nocturnality.
a
TSTL and TSTD values calculated from the LD ratio.

differences in TST are caused by (in most cases, Table 1. However, this ratio also depends on the
unknown) differences in environmental conditions total amount of sleep (low 24-h TST results in
between the laboratories. The same variability higher ratios). Another index for the amplitude,
may also be responsible for the difference in the the percentage of 24-h TST spent in the main
ratio of TST between light (TSTL) and dark rest period, is found in a much closer range
(TSTD; an index for diurnality/nocturnality of (65.9–69.5%), suggesting that this variable may
the sleep–wake rhythm) which ranges from 1.94 be less susceptible to differences in environmental
to 2.28 in the examples for C57BL/6 mice in conditions.
206

Table 2. Comparison of mouse strains recorded in the same laboratory

% TST % TSTL % TSTD Ratio % TST in rest period

Huber et al. (2000a)


129/Ola 51.2 63.6 38.7 1.64 62.1
129/SvJ 58.6 66.9 50.3 1.33 57.1
C57BL/6J 48.6 64.1 33.1 1.94 65.9
Franken et al. (1999)
AKR/J 48.0 63.9 32.1 1.99 66.6
BALB/cByJ 43.9 52.2 35.6 1.47 59.4
C57BL/6J 40.6 56.4 24.7 2.28 69.5
C57BR/cdJ 39.9 59.1 20.2 2.93 74.1
DBA/2J 36.4 55.3 17.5 3.16 76.0
129/Ola 40.7 58.3 23.0 2.53 71.6

TST, total sleep time over 24 h; TSTL, total sleep time during the light period; TSTD, total sleep time during the dark period; Ratio, ratio between
TSTL and TSTD, a measure for diurnality or nocturnality.

Table 3. Effects of SCN lesion, photoperiod, and wheel availability

Condition % TST % TSTL % TSTD Ratio % TST in rest period References

LD/DD
Mouse C57BL/6 LD 46.6 63.4 29.8 2.13 68.0 Deboer et al. (2007b)
DD 47.8 65.0 30.0 2.17 78.7
Sprague–Dawley rat LD 49.2 68.0 30.4 2.24 69.1 Tobler et al. (1994)
DD 50.3 67.5 33.0 2.05 67.1
Wheel availability
Degu No wheel 37.6 36.7 38.5 1.05 51.2 Kas and Edgar (1998)
Wheel 38.2 39.2 36.2 1.08 51.3
Mouse C57BL/6 No wheel 43.8 63.3 24.2 2.62 72.7 Vyazovskiy et al. (2006)
Wheel 39.6 68.3 10.8 6.32 86.2
SCN lesion
Sprague–Dawley rat SCN 57.1 78.1 36.0 2.17 68.3 Mistlberger et al. (1983)
SCNx 60.7 60.8 60.7 1.00 50.1
Rata SCN 51.2 67.9 38.2 1.78 66.3 Coindet et al. (1975)
SCNx 52.5 59.2 45.8 1.29 56.4
Photoperiod
Siberian chipmunk LD 18/6 48.7 39.3 77.6 1.97 39.8 Dijk and Daan (1989)
LD 12/12 52.8 29.1 74.7 2.57 70.7
LD 6/18 53.0 17.2 65.7 3.82 82.6
Djungarian hamster LD 16/8 59.2 67.2 42.8 1.57 87.3 Deboer and Tobler (1996)
LD 8/16 61.8 71.7 56.7 1.26 38.7
Rat LD 16/8 52.7 64.4 28.3 2.28 94.0 Franken et al. (1995)
LD 12/12 47.9 68.9 26.9 2.56 71.9
LD 8/16 55.8 73.3 45.7 1.60 43.8

TST, total sleep time over 24 h; TSTL, total sleep time during the light period; TSTD, total sleep time during the dark period; Ratio, ratio between
TSTL and TSTD, a measure for diurnality or nocturnality.
a
Strain not specified.
207

Another aspect that becomes clear from the endogenous circadian pacemaker. Changes in
data in Table 1 is that some species which, from these factors result in significant changes in overt
the behavioral point of view, are considered to rest–activity behavior and sleep.
be clearly diurnal or nocturnal do not show a Mice and rats, acutely transferred from a
large difference in TSTL and TSTD. The Euro- 12 h:12 h L:D cycle into constant darkness (DD),
pean and 13-lined ground squirrel are considered do not show large changes in their TST over
to be day active, and the Long Evans rat and the 24 h (Table 3). In addition, no clear changes in
chinchilla are known as night active species. They the preferences for sleep in the rest or active
spent, however, less than 60% of their 24-h TST phase were observed. In albino Sprague–Dawley
in their main rest period. The guinea pig is, on rats, the distribution of sleep and waking across
the basis of its day–night difference in sleep, not the circadian day did show a small increase in
different from these four species. Yet, when ana- REM sleep and a clear increase in SWA in
lyzing the hourly distribution of activity and NREM sleep in the rest period (Tobler et al.,
TST, it is considered to be a species with no pref- 1994). In contrast, in pigmented C57BL/6 mice,
erence for sleep in the light or dark period. It is no changes were observed from LD to DD
the detailed distribution of sleep and waking dur- (Deboer et al., 2007b). These differences between
ing the active phase, and particularly the expres- rats and mice may be related to the difference in
sion of a relatively long consolidated waking pigmentation between the animals but may also
period and the overt behavior during that active be based on species-specific differences.
period (i.e., increased activity), which determines It may be obvious that removal of the central
whether an animal is considered to have a clear pacemaker will influence the distribution of sleep.
diurnal or nocturnal preference. After SCN lesions, the difference in TST ratio
When comparing the variation between mouse between the light and the dark period decreases,
strains, it seems that at least part of the diurnal but sometimes a preference for sleep in the light
modulation of sleep has a genetic basis. Different period remains (Coindet et al., 1975). The latter
mouse strains recorded under the same cir- may be a masking effect of light due to a suppres-
cumstances in the same laboratory have different sion of activity (Deboer and Tobler, 2000) and
amounts of TST over 24 h (ranging from 40% to direct induction of sleep by light (Alfoldi et al.,
50%,  2.5 h difference) and show also strong 1991; Benca et al., 1998; Deboer et al., 2007b).
differences in the light–dark ratios (Table 2). In most SCN lesion experiments in mice and rats,
Strains like 129/SV and BALBc have relatively performed under constant dark conditions, the
low ratios, whereas C57 strains, 129/Ola, and par- amount of sleep over 24 h did not change,
ticularly DBA show large differences between indicating that the endogenous clock does not
light and dark. However, quantitative trait loci determine the amount of sleep (Ibuka et al.,
analysis did not result in a putative genetic locus 1980; Mistlberger et al., 1983; Tobler et al.,
responsible for diurnal variation in TST (Tafti 1983), but also here exceptions have been
et al., 1997). reported (Easton et al., 2004).
An increase of sleep in the rest phase was seen
in mice which were provided with a running
Clocks, photoperiod, running wheels, and light wheel (Vyazovskiy et al., 2006). Here, TST over
24 h hardly changed, but the preference for sleep
Chronobiologically relevant factors that may in the light period became stronger. Without a
influence sleep are light and darkness, the photo- running wheel, the Degu is a crepuscular species
period (day length), the presence or absence of a with a slight preference for sleep in the dark
running wheel, and the presence or absence of an period. The availability of a running wheel
208

reversed this preference, making the animals they will spend a substantial amount of time dur-
slightly nocturnal (Kas and Edgar, 1998). These ing this active period also in sleep. In general, the
data suggest that the distribution of sleep, and data show that rodents are very flexible in their
possibly also the amount of sleep, is subject to choice of the time of day for their sleep.
time management “decisions,” just like other
behavioral activities observed in mice (de Visser
et al., 2005). Circadian–homeostatic interactions
Under different photoperiods, TST over 24 h
does not change systematically in rodents As mentioned in the section “Introduction,” sleep
(Table 3); however, the ratio of the difference in is regulated by homeostatic and circadian
TST between light and dark changes. In the diur- mechanisms. This begs the question whether
nal Siberian chipmunk, the amplitude increased these two mechanisms regulate sleep indepen-
(Dijk and Daan, 1989), whereas in the nocturnal dently or whether there is an interaction between
rat and Djungarian hamster, the amplitude the two. Experiments in rodents have contributed
decreased (Deboer and Tobler, 1996; Franken extensively to answer this question. From sleep
et al., 1995) when photoperiod shortened. This deprivation experiments in SCN-lesioned rats
suggests that diurnal and nocturnal rodents may (Mistlberger et al., 1983; Tobler et al., 1983;
show an opposite response to photoperiodic Trachsel et al., 1992), it became clear that sleep
changes. The distribution of sleep of the rodent homeostatic mechanisms function without the
species confirms this (Fig. 1). In humans, the presence of an intact circadian clock. Phase-
effect of photoperiodic changes seems to differ shifting experiments in humans showed that sleep
slightly from the changes in the Siberian chip- homeostatic regulation does not follow the
munk as the percentage of TSTD decreased circadian timing system (Dijk et al., 1989). There-
(Wehr et al., 1993), reducing the daily amplitude fore, it was assumed that sleep homeostasis
in the short photoperiod. What is also clear from and circadian regulation of sleep function inde-
these data is that when the duration of the oppor- pendently. However, different types of continu-
tunity to sleep in the preferred phase of the day ous regulatory interactions were also proposed
(light or dark) is reduced, rodents decrease the (Achermann and Borbely, 1994; Edgar et al.,
relative amount of TST in this period without 1993; Mistlberger, 2005).
changing TST over 24 h. This is in contrast with The notion that changes in sleep pressure can
the available human data, where in general, no influence the functioning of the circadian clock
sleep is observed during the light period. The lat- came from data in mice and hamsters, which
ter may be because most experimental designs in showed that sleep deprivation attenuates phase
human sleep research do not allow sleep in the shifts induced by light (Challet et al., 2001; Mis-
light period, but it may also be because human tlberger et al., 1997). It was also shown that sleep
sleep is more disturbed by light. The flexibility deprivation can change circadian phase and FOS
in the timing of sleep in rodents results, under expression in the SCN (Antle and Mistlberger,
certain photoperiods, in the rather extreme con- 2000). Recently, it was shown that partial sleep
dition that the Siberian chipmunk and the deprivation reduces phase advances to light in
Djungarian hamster spent less than 40% of their humans as well (Burgess, 2010).
TST over 24 h in what is considered to be their When sleep–wake behavior and SCN neuronal
main “sleep” or rest period. The remaining sleep activity are recorded simultaneously, the influ-
is fulfilled in their main “waking” or active ence of sleep on SCN neuronal activity becomes
period. This means that, although the animals very clear. At the onset of NREM sleep, SCN
are more active compared to the main rest period, neuronal activity decreases, whereas at the
209

100 100
Human
75 75

50 50

25 25

0 0

100 100
Chipmunk
75 75

50 50

25 25
Sleep (%)

0 0

100 100
Rat
75 75

50 50

25 25

0 0

100 100
Djungarian hamster
75 75

50 50

25 25

0 0
0 6 12 18 24 0 6 12 18 24
Time since mid-active period (h)

Fig. 1. Mean 24-h profiles total sleep time in four different species in two different photoperiods. Mean 24-h profiles of EEG
confirmed sleep in humans, Siberian chipmunks, rats, and Djungarian hamsters in a short (left) and long (right) main sleep
period. Human data adapted from Wehr et al. (1993), Siberian chipmunk data adapted from Dijk and Daan (1989), rat data
adapted from Franken et al. (1995), and Djungarian hamster data adapted from Deboer and Tobler (1996).

transition from NREM sleep to REM sleep or in NREM sleep was found, suggesting that deep
waking, the activity increases again (Deboer sleep reduces SCN activity. Further experiments
et al., 2003). In addition, a significant negative cor- showed that SCN neuronal activity is attenuated
relation between SWA and SCN neuronal activity after sleep deprivation for approximately the same
210

duration as SWA in NREM sleep is increased dark period (Franken et al., 1991). In an experiment
(Deboer et al., 2007a). Taken together, the data where the effect of a 6-h sleep deprivation was
show that information about vigilance state tested in the light and in the dark period, SWA
changes and changes in sleep pressure reach the was lower than S in the dark period (Vyazovskiy
SCN and influence its neuronal firing pattern. et al., 2007). Even in DD conditions, a circadian
Recent experiments applying functional magnetic modulation of the decreasing time constant was nec-
resonance imaging show similar effects in humans essary to obtain an optimal fit, suggesting that there
(Schmidt et al., 2009). The data suggest that sleep is an endogenous circadian influence on the expres-
homeostatic feedback on the output of the endog- sion of SWA in NREM sleep (Deboer, 2009).
enous clock may be continuous instead of being Detailed analysis of the behavior of SWA revealed
restricted to distinct time points in the circadian that the rate of increase of SWA at the onset of a
cycle. Recently, a model which integrates the NREM sleep episode showed a circadian modula-
traditional two process model of sleep regulation tion, which may explain part of the systematic circa-
with these recent findings was proposed (Dijk dian deviations observed in the simulations of
and Archer, 2010). In this conceptual model, Process S in the rat (Deboer, 2009).
sleep homeostasis and the circadian clock
function independently, but sleep homeostatic
pressure feeds back on the circadian clock, Clock genes and sleep
influencing its output.
Whether information on circadian time also The discovery of clock genes has brought new per-
influences sleep homeostasis, other than intervening spective to the relationship between the circadian
with the timing of sleep, remains unclear. However, clock and sleep. The search for clock genes has
some evidence that the circadian clock may influ- been highly successful, resulting in five to seven
ence sleep homeostatic mechanisms comes from central clock genes (Cry1, Cry2, BMAL1, Clock,
mathematical simulations of sleep homeostasis in Per1, Per2, and Per3, reviewed by Albrecht,
rodents. In these simulations, the time course of 2002). Knocking out one or more of these genes
the sleep homeostatic process (Process S) is pre- results in animals with a changed or disabled circa-
dicted on the basis of prior sleep–wake history and dian clock in which sleep regulation can be
correlated with SWA in NREM sleep. In rodents, investigated, and in some cases, obstructing normal
the time course of Process S is determined itera- clock functioning also seemed to change sleep
tively on the basis of the vigilance states. During homeostasis.
waking and REM sleep, S increases, whereas during For instance, Cry1 and Cry2 double-knockout
NREM sleep, S decreases (Deboer, 2009; Franken mice showed increased NREM sleep time and
et al., 1991; Huber et al., 2000a; Vyazovskiy et al., consolidation, and increased EEG SWA in
2007). Both increase and decrease of Process S NREM sleep (Wisor et al., 2002), which are all
occur with corresponding fixed time constant signs of high homeostatic NREM sleep pressure.
(Deboer, 2009; Franken et al., 1991; Huber et al., The BMAL1 mutation exhibits more sleep and
2000a; Vyazovskiy et al., 2007). For mice, this an attenuation of the homeostatic response to
approach was sufficient to reach an optimal fit in sleep deprivation (Laposky et al., 2005). Also
most strains (Franken et al., 2001; Huber et al., the Clock mutation seems to alter sleep homeo-
2000a). However, in the rat, an additional modula- stasis, since the mutant mice sleep less than
tion of the decreasing time constant (determining wild-type control animals (Naylor et al., 2000).
the speed of decrease of Process S during NREM All three clock gene mutations induce arrhyth-
sleep) was needed, as SWA was consistently higher mic behavior (Bunger et al., 2000; van der Horst
than S in the light period and lower than S in the et al., 1999; Vitaterna et al., 1994), similar to
211

SCN-lesioned animals. However, the difference is which are not present in other mammals (Jenkins
that SCN lesions in rodents generally (for et al., 2005).
exceptions, see Easton et al., 2004) do not change The homeostatic response after sleep depriva-
the time spent in sleep (Ibuka et al., 1980; Mis- tion is intact in most mentioned clock gene
tlberger et al., 1983; Tobler et al., 1983; Trachsel mutants, which again supports the notion that
et al., 1992). NPAS2-deficient mice also show a sleep homeostasis can function independently
reduction in sleep time (Dudley et al., 2003) and from the circadian clock.
an attenuated response after sleep deprivation
(Franken et al., 2006). Modeling sleep homeostasis
in Cry1,2 double knockout mice showed that the Separating circadian and homeostatic processes
parameters describing the increase in sleep pres-
sure during waking are altered such that sleep pres- Experimental “separation” of circadian- and sleep-
sure builds up faster compared to wild type (Wisor dependent processes can help in understanding the
et al., 2002). The results indicate that Cry, BMAL1, basic mechanisms in the functioning of these two
NPAS2, and Clock are not only involved in major regulating systems. This separation is com-
generating circadian rhythms but may also play plicated by the synchronous progression of circa-
a role in sleep homeostasis, probably in brain dian time and changes in sleep and wakefulness.
areas outside the SCN. However, these genes are In humans, several protocols (ultrashort day,
knocked out in the entire organism from the begin- forced desynchrony) have been established to
ning of development. It is therefore unclear which uncouple sleep homeostasis from the output of
part of these results is caused by changes in circa- the circadian pacemaker (Carskadon and Dement,
dian regulation in the SCN, and which by loss 1975; Dijk and Czeisler, 1995; Lavie and Scherson,
and/or compensation of gene function in other 1981). In animals, research in this direction was
regions of the brain. Until studies become available done either after SCN lesions (Mistlberger et al.,
where the mice underwent SCN lesions or the 1983; Tobler et al., 1983; Trachsel et al., 1992) or
clock genes are mutated/knocked out locally in in spontaneously arrhythmic animals (Deboer
the SCN, the exact meaning of these results may and Tobler, 2003; Larkin et al., 2004; Tobler
not be resolved. and Franken, 1993), mainly showing that sleep
The sleep data obtained in mPer1 and mPer2 homeostasis can function independently from the
mutant mice are better understood and suggest a circadian clock.
role for these genes in the circadian timing of sleep. The basic problem is to differentiate daily
Particularly, the mPer2 mutant shows a clear phe- changes in observed variables, caused by the
notype with advanced sleep reflecting the phase central clock itself (endogenously), from environ-
advance of motor activity onset in the mutant mental influences of light, temperature (exoge-
(Kopp et al., 2002; Shiromani et al., 2004), which nous), and/or masking effect of behavior, such as
corresponds with similar observations in humans, sleep, rest, and activity, or food and water intake,
with a mutation in the Per2 gene in a familial which are only in part or indirect clock driven.
form of advanced sleep phase syndrome (Toh As early as 1938, Nathanial Kleitman per-
et al., 2001). formed in a pilot study a forced desynchrony pro-
The correspondence between mPer3 mutant tocol where the subjects were forced to sleep on a
mice and human polymorphism in the Per3 gene, daily schedule outside of the reach of the endoge-
associated with delayed sleep phase syndrome, nous circadian pacemaker. Unfortunately, only
is less clear. An explanation for this discrepancy one of the two participants in this experiment
may be that in primates the Per3 gene contains was able to follow the forced schedule of a 21-
a number of tandem repeat polymorphisms or a 28-h day (Kleitman, 1963). Other attempts
212

were made, but it took until the 1990s before the alternations (Lavie and Scherson, 1981), and the
protocol was applied systematically in humans 10/20 min ultrashort sleep–wake schedule (Liu
(Dijk and Czeisler, 1995). This research brought et al., 2000; Tagaya et al., 2002). Also here a
many new insights into the interaction between strong waking peak was found around 10 p.m.
sleep homeostasis and the circadian clock in and a peak in REM sleep in the early morning.
humans. It showed that in humans, waking is Particularly, the inability to sleep properly around
enforced very strongly by the endogenous 10 p.m. made that subjects lost 30% of TST over
clock, particularly at the end of the day (around 24 h compared to normal baseline conditions
10 p.m.), and that also REM sleep has a strong (Carskadon and Dement, 1975, 1980).
circadian component, peaking in the early morn- A similar approach, to specifically separate the
ing. In rodents, a forced desynchrony protocol with circadian- and sleep-dependent components of
repeated sleep deprivations and rest according to sleep regulation, was performed recently in freely
a 20 h:20 h schedule was applied only once, but behaving rats. Here the animals were exposed to
unfortunately no sleep was recorded (Strijkstra a 2 h/2 h short-sleep deprivation protocol in
et al., 1999). constant darkness (Yasenkov and Deboer, 2010,
In another type of protocol in rats, a 22-h dim 2011). The protocol aims to stabilize the homeo-
light–dark cycle is applied which results in a split static sleep pressure over the circadian day,
of the rest–activity rhythm into two components. without reducing the opportunity to sleep. In the
One component follows the 22-h light–dark cycle, case of the rat, this means that the animals should
whereas the other component ( 25 h) follows the be able to sleep approximately 50% of 24 h. For
endogenous circadian pacemaker (de la Iglesia that reason, the sleep deprivation paradigm con-
et al., 2004). After releasing into DD, the pattern sisted of 2-h sleep deprivation periods alternated
immediately disappears and only a single circa- with 2-h periods of rest (2 h/2 h protocol). This
dian period remains. With this protocol, it was schedule was maintained for at least 48 h to
possible to induce a 22-h rhythm in NREM sleep ensure an equilibrium can be reached. Another
and waking, but not REM sleep, next to the nor- crucial part of the experimental design was the
mal endogenous 25-h circadian rhythm in sleep constant DD condition which guaranteed that
and waking (Cambras et al., 2007). Unfortu- any physiological changes in the circadian range
nately, the authors did not separate sleep homeo- observed during the experimental procedures
stasis from the 25-h rhythm induced by the are due to intrinsic properties of the endogenous
circadian clock, as was done in the forced circadian clock, and not caused by the changes
desynchrony experiment in humans (Dijk and of the light regime.
Czeisler, 1995). This limits the interpretation of The protocol was successful in stabilizing
the data concerning the relative role of the circa- homeostatic sleep pressure as SWA in NREM
dian clock and the sleep homeostat in sleep regu- sleep was equally distributed across the circadian
lation in the rat. As a result, the clock-driven day (Yasenkov and Deboer, 2010, 2011). Despite
circadian component of sleep regulation is proba- constant DD conditions and constant sleep pres-
bly overestimated in this protocol. sure, the animals still showed a circadian rhythm
Several experiments were carried out in in NREM sleep, but with a lower amplitude,
humans using the short “90-min day” protocol whereas REM sleep was equally distributed
(Carskadon and Dement, 1975, 1980), in which across the circadian day (Fig. 2, top panels). The
30 min of rest and 60 min of enforced wakefulness remaining circadian rhythm in NREM sleep was
are alternated for several days. Variations on caused by a significant modulation of NREM sleep
this protocol were the “ultrashort sleep–wake episode duration (Fig. 2). NREM sleep episode fre-
schedule,” with 15 min of waking–5 min sleep quency, and REM sleep episode frequency and
213

NREM sleep REM sleep

100 20

75 15
Amount
of sleep
(%)
50 10

25 5

0 0
8
of sleep eposides

6
(episodes/h)

7
Frequency

6 4

5 2
4
0
9
of sleep episodes

8
7 1.5
Duration

(min)

6
1.0
5
4 0.5
3
2 0
0 12 24/0 12 24/0 0 12 24/0 12 24/0
Circadian time (h)

2 h/2 h Baseline

Fig. 2. Circadian profiles of NREM and REM sleep (amount, frequency, and duration) in baseline and during constant sleep
pressure in the rat. Top panels—amount of NREM sleep and REM sleep plotted as percentage of recording time. Middle
panels—sleep episode frequencies (events per hour). Bottom panels—sleep episode duration (min) for corresponding vigilance
states. All data computed for the undisturbed 2-h periods of the 2 h/2 h protocol and corresponding baseline intervals. Each data
point represents 2-h mean values (n ¼ 8) for 2 h/2 h protocol (dots) and corresponding baseline intervals (circles). Data are
double plotted to visualize the circadian modulation. Top horizontal bars indicate the periods of subjective rest (gray) and
activity (black). *p < 0.05 and **p < 0.01: two-tailed paired t-test after significant ANOVA or rANOVA, factor “day.” Data
adapted from Yasenkov and Deboer (2010).

duration did not show a circadian rhythm during humans, where under similar conditions sleep
the 2 h/2 h protocol. These findings are of interest dropped to less than 5 h per 24 h (Carskadon
because this is much harder, if not impossible, to and Dement, 1975; Dantz et al., 1994). This differ-
see such changes in species in which sleep episodes ence between humans and rodents is probably
are much longer, like humans. caused by the polyphasic nature of sleep in
During the 2 h/2 h protocol, the animals lost rodents. From these experiments, it can be con-
only 7% of their TST over 24 h, which was caused cluded that, in contrast to humans, the daily mod-
by a decrease in NREM sleep, as the amount of ulation of REM sleep in rats is less pronounced
REM sleep did not change (Yasenkov and and possibly more sensitive to changes in sleep
Deboer, 2010). This is in strong contrast with homeostatic pressure. In contrast, the amount of
214

NREM sleep and waking still shows a clear influ- Mistlberger, 2000), but it can also be that changes
ence of the circadian clock in the rat. In addition, in sleep homeostatic pressure merely change the
the experimental conditions resulted in a clear output of the SCN.
separation of circadian- and sleep-dependent
components between the baseline day and the
2 h/2 h protocol, and explicit sleep- and circadian- Protocols and techniques
dependent EEG frequencies in both NREM and
REM sleep were identified (Yasenkov and An application of new protocols and techniques
Deboer, 2011). The data indicate that in the rat may elucidate some of the questions. For the
the circadian changes in homeostatic sleep pres- understanding of the interaction between the cir-
sure, which accompany the circadian changes in cadian and sleep homeostatic component in
waking, strongly amplify the circadian pattern in rodents, the 2 h/2 h protocol may be an attractive
NREM and REM sleep. Therefore, despite the tool. We have shown that in the protocol, any
larger flexibility in sleep timing in rodents, still a changes obtained in the course of the circadian
clear circadian modulation of sleep can be day are clearly not induced by the sleep
observed when the two are in phase under undis- homeostat, but probably by the circadian clock.
turbed conditions. The protocol can be applied on several different
experimental levels from gene expression to
behavior and cognition.
Future perspectives However, the constant sleep pressure in the
protocol can also be a disadvantage. A true
There are still several open questions related to forced desynchrony protocol in a rodent, where
sleep regulation. The EEG SWA data obtained sleep homeostatic changes occur out of phase
during NREM sleep suggest the presence of some with the circadian clock, will probably elucidate
kind of homeostatic process. However there is no more complex interactions. These type of exper-
clear evidence where the homeostat is located in iments are time consuming, as they are in
the brain. This question has become increasingly humans, and difficult to perform in rodents, as
pressing after the discovery of cortical differences rodents are not as good as humans in following
in local sleep homeostasis, where time constants instructions from the experimenter.
of homeostatic increase and decrease of EEG
SWA (and therefore of Process S) differed
between cortical areas (Huber et al., 2000b; Werth Modeling
et al., 1996a). Also the discovery of a cortical sleep
homeostat in rats (Vyazovskiy et al., 2009) or the The modeling of sleep homeostasis has seen many
finding that learned tasks influence the level of new developments since its implementation in the
EEG SWA in subsequent NREM sleep in the cor- two process model of sleep regulation, and the
tical area related to this task (Huber et al., 2004; time course of SWA in the course of a sleep epi-
Vyazovskiy et al., 2004) adds to the complexity. It sode can be predicted very precise (Achermann
is also unknown whether there is a neuroanatomi- et al., 1993). Unfortunately, simulation of sleep
cal substrate for the interaction between sleep homeostasis in rodents did not reach this level
homeostasis and the circadian clock. It seems that of detail. In both humans and rodents, the circa-
the interaction may take place in the SCN, as dian process is not incorporated routinely in
changes in sleep pressure influence SCN neuronal present-day modeling. There are, however, some
firing (Deboer et al., 2003, 2007a) and can influ- indications that the circadian clock may influence
ence circadian phase resetting (Antle and sleep homeostatic time constants and that an
215

effort should be made to incorporate circadian REM rapid eye movement sleep
time as a variable in the modeling of sleep sleep
homeostasis. For instance, when modeling sleep SCN suprachiasmatic nucleus
homeostasis in the rat, time-of-day-dependent SWA slow-wave activity
deviations are observed in the fit between model TST total sleep time
and data, even under constant dark conditions TSTD total sleep time in the dark
(Deboer, 2009; Franken et al., 1991). In addition, TSTL total sleep time in the light
it was also shown that, in animals lacking the Cry1
and Cry2 genes, the buildup rate of the homeo-
static process was significantly shortened (Wisor Glossary of sleep-related terms
et al., 2002). Discrepancies were observed
between the data and the simulation when the Electroencephalogram Electrical activity in the
homeostatic sleep process was simulated for sleep brain recorded from scalp electrodes (in
scheduled outside the normal sleep phase, like an humans) or epidural electrodes (in animals).
afternoon nap (Werth et al., 1996b), sleep in the Non-rapid eye movement sleep One of the two
early evening, or extended sleep (Achermann basic sleep states (see also rapid eye movement
et al., 1993). From the data presented in this sleep) characterized in animals by slow waves
review, it is clear that the circadian clock does (< 5 Hz) in the electroencephalogram and low
probably more than mark the onset and end of muscle activity (see also slow waves).
the rest and activity phase, and this should be Rapid eye movement sleep One of the two basic
implemented into the models. In addition, more sleep states (see also non-rapid eye movement
data of different species (and not only rodents) sleep) characterized by rapid eye movements,
are needed to be able to eventually come up with muscle atonia, muscle twitches, and low-amplitude
a model that can be applied to general mamma- fast-frequency electroencephalogram.
lian sleep. Sleep intensity Depth of sleep.
Sleep homeostasis A balancing process that
maintains the amount and intensity of sleep
Acknowledgment within certain boundaries. Slow waves in undis-
turbed non-rapid eye movement sleep are
This research was supported by the Netherlands thought to be a reflection of the level of the
Organization for Scientific Research (NWO, sleep homeostatic process.
Grant 818.02.016 to T. D.). Slow waves Oscillations in the sleep electroen-
cephalogram below 5 Hz which are a character-
istic of non-rapid eye movement sleep.
Abbreviations Slow-wave activity spectral power in the slow-
wave frequency range of the electroencephalo-
2 h/2 h the repeated short-sleep depriva- gram, which is a measure of non-rapid eye
tion protocol movement sleep intensity.
CT circadian time
DD constant darkness
EEG electroencephalogram
References
EMG electromyogram
LD light/dark cycle Achermann, P., & Borbely, A. A. (1994). Simulation of day-
NREM non-rapid eye movement sleep time vigilance by the additive interaction of a homeostatic
sleep and a circadian process. Biological Cybernetics, 71, 115–121.
216

Achermann, P., & Borbely, A. A. (2011). Sleep homeostasis Dantz, B., Edgar, D. M., & Dement, W. C. (1994). Circadian
and models of sleep regulation. In M. Kryger, T. Roth & rhythms in narcolepsy: Studies on a 90 minute day. Electro-
W. C. Dement (Eds.), Principles and practice of sleep medi- encephalography and Clinical Neurophysiology, 90, 24–35.
cine. (5th ed., pp. 431–444). St. Louis: Elsevier Saunders. de la Iglesia, H. O., Cambras, T., Schwartz, W. J., &
Achermann, P., Dijk, D. J., Brunner, D. P., & Borbely, A. A. Diez-Noguera, A. (2004). Forced desynchronization of dual
(1993). A model of human sleep homeostasis based on EEG circadian oscillators within the rat suprachiasmatic nucleus.
slow-wave activity: Quantitative comparison of data and Current Biology, 14, 796–800.
simulations. Brain Research Bulletin, 31, 97–113. de Visser, L., van den Bos, R., & Spruijt, B. M. (2005).
Albrecht, U. (2002). Invited review: Regulation of mammalian Automated home cage observations as a tool to measure
circadian clock genes. Journal of Applied Physiology, 92, the effects of wheel running on cage floor locomotion.
1348–1355. Behavioural Brain Research, 160, 382–388.
Alfoldi, P., Franken, P., Tobler, I., & Borbely, A. A. (1991). Deboer, T. (2009). Sleep and sleep homeostasis in constant
Short light-dark cycles influence sleep stages and EEG darkness in the rat. Journal of Sleep Research, 18,
power spectra in the rat. Behavioural Brain Research, 43, 357–364.
125–131. Deboer, T., Detari, L., & Meijer, J. H. (2007a). Long term
Antle, M. C., & Mistlberger, R. E. (2000). Circadian clock effects of sleep deprivation on the mammalian circadian
resetting by sleep deprivation without exercise in the Syrian pacemaker. Sleep, 30, 257–262.
hamster. The Journal of Neuroscience, 20, 9326–9332. Deboer, T., Ruijgrok, G., & Meijer, J. H. (2007b). Short light-
Benca, R. M., Gilliland, M. A., & Obermeyer, W. H. (1998). dark cycles affect sleep in mice. European Journal of Neuro-
Effects of lighting conditions on sleep and wakefulness in science, 26, 3518–3523.
albino Lewis and pigmented Brown Norway rats. Sleep, 21, Deboer, T., & Tobler, I. (1996). Shortening of the photoperiod
451–460. affects sleep distribution, EEG and cortical temperature in
Bunger, M. K., Wilsbacher, L. D., Moran, S. M., the Djungarian hamster. Journal of Comparative Physiology
Clendenin, C., Radcliffe, L. A., Hogenesch, J. B., et al. A, 179, 483–492.
(2000). Mop3 is an essential component of the master circa- Deboer, T., & Tobler, I. (2000). Running wheel size influences
dian pacemaker in mammals. Cell, 103, 1009–1017. circadian rhythm period and its phase shift in mice. Journal
Burgess, H. J. (2010). Partial sleep deprivation reduces phase of Comparative Physiology A, 186, 969–973.
advances to light in humans. Journal of Biological Rhythms, Deboer, T., & Tobler, I. (2003). Sleep regulation in the
25, 460–468. Djungarian hamster: Comparison of the dynamics leading
Cambras, T., Weller, J. R., Angles-Pujoras, M., Lee, M. L., to the slow-wave activity increase after sleep deprivation
Christopher, A., Diez-Noguera, A., et al. (2007). Circa- and daily torpor. Sleep, 26, 567–572.
dian desynchronization of core body temperature and Deboer, T., Vansteensel, M. J., Detari, L., & Meijer, J. H.
sleep stages in the rat. Proceedings of the National Acad- (2003). Sleep states alter activity of suprachiasmatic nucleus
emy of Sciences of the United States of America, 104, neurons. Nature Neuroscience, 6, 1086–1090.
7634–7639. Dijk, D. J., & Archer, S. N. (2010). PERIOD3, circadian
Carskadon, M. A., & Dement, W. C. (1975). Sleep studies on a phenotypes, and sleep homeostasis. Sleep Medicine Reviews,
90-minute day. Electroencephalography and Clinical Neuro- 14, 151–160.
physiology, 39, 145–155. Dijk, D. J., Beersma, D. G., & Daan, S. (1987). EEG power
Carskadon, M. A., & Dement, W. C. (1980). Distribution of density during nap sleep: Reflection of an hourglass measur-
REM sleep on a 90 minute sleep-wake schedule. Sleep, 2, ing the duration of prior wakefulness. Journal of Biological
309–317. Rhythms, 2, 207–219.
Challet, E., Turek, F. W., Laute, M., & Van Reeth, O. (2001). Dijk, D. J., Beersma, D. G., Daan, S., & Lewy, A. J. (1989).
Sleep deprivation decreases phase-shift responses of circa- Bright morning light advances the human circadian system
dian rhythms to light in the mouse: Role of serotonergic without affecting NREM sleep homeostasis. American Jour-
and metabolic signals. Brain Research, 909, 81–91. nal of Physiology, 256, R106–R111.
Coindet, J., Chouvet, G., & Mouret, J. (1975). Effects of Dijk, D. J., & Czeisler, C. A. (1995). Contribution of the circa-
lesions of the suprachiasmatic nuclei on paradoxical sleep dian pacemaker and the sleep homeostat to sleep propen-
and slow wave sleep circadian rhythms in the rat. Neurosci- sity, sleep structure, electroencephalographic slow waves,
ence Letters, 1, 243–247. and sleep spindle activity in humans. The Journal of Neuro-
Daan, S., Spoelstra, K., Albrecht, U., Schmutz, I., Daan, M., science, 15, 3526–3538.
Daan, B., et al. (2011). Lab mice in the field: Unorthodox Dijk, D. J., & Daan, S. (1989). Sleep EEG spectral analysis in
daily activity and effects of a dysfunctional circadian clock a diurnal rodent: Eutamias sibiricus. Journal of Comparative
allele. Journal of Biological Rhythms, 26, 118–129. Physiology A, 165, 205–215.
217

Dudley, C. A., Erbel-Sieler, C., Estill, S. J., Reick, M., Kas, M. J., & Edgar, D. M. (1998). Crepuscular rhythms of
Franken, P., Pitts, S., et al. (2003). Altered patterns of sleep EEG sleep-wake in a hystricomorph rodent, Octodon degus.
and behavioral adaptability in NPAS2-deficient mice. Journal of Biological Rhythms, 13, 9–17.
Science, 301, 379–383. Kleitman, N. (1963). Sleep and wakefulness. Chicago &
Easton, A., Meerlo, P., Bergmann, B., & Turek, F. W. (2004). London: The University of Chicago Press.
The suprachiasmatic nucleus regulates sleep timing and Kopp, C., Albrecht, U., Zheng, B., & Tobler, I. (2002).
amount in mice. Sleep, 27, 1307–1318. Homeostatic sleep regulation is preserved in mPer1 and
Edgar, D. M., Dement, W. C., & Fuller, C. A. (1993). Effect of mPer2 mutant mice. European Journal of Neuroscience, 16,
SCN lesions on sleep in squirrel monkeys: Evidence for 1099–1106.
opponent processes in sleep-wake regulation. The Journal Lancel, M., van Riezen, H., & Glatt, A. (1991). Effects of cir-
of Neuroscience, 13, 1065–1079. cadian phase and duration of sleep deprivation on sleep and
Franken, P., Chollet, D., & Tafti, M. (2001). The homeostatic EEG power spectra in the cat. Brain Research, 548, 206–214.
regulation of sleep need is under genetic control. The Jour- Laposky, A., Easton, A., Dugovic, C., Walisser, J.,
nal of Neuroscience, 21, 2610–2621. Bradfield, C., & Turek, F. (2005). Deletion of the mamma-
Franken, P., Dudley, C. A., Estill, S. J., Barakat, M., lian circadian clock gene BMAL1/Mop3 alters baseline
Thomason, R., O’Hara, B. F., et al. (2006). NPAS2 as a sleep architecture and the response to sleep deprivation.
transcriptional regulator of non-rapid eye movement sleep: Sleep, 28, 395–409.
Genotype and sex interactions. Proceedings of the National Larkin, J. E., Yokogawa, T., Heller, H. C., Franken, P., &
Academy of Sciences of the United States of America, 103, Ruby, N. F. (2004). Homeostatic regulation of sleep in
7118–7123. arrhythmic Siberian hamsters. American Journal of Physiol-
Franken, P., Malafosse, A., & Tafti, M. (1999). Genetic ogy. Regulatory, Integrative and Comparative Physiology,
determinants of sleep regulation in inbred mice. Sleep, 22, 287, R104–R111.
155–169. Lavie, P., & Scherson, A. (1981). Ultrashort sleep-walking
Franken, P., Tobler, I., & Borbely, A. A. (1991). Sleep homeo- schedule. I. Evidence of ultradian rhythmicity in
stasis in the rat: Simulation of the time course of EEG slow- “sleepability” Electroencephalography and Clinical Neuro-
wave activity. Neuroscience Letters, 130, 141–144. physiology, 52, 163–174.
Franken, P., Tobler, I., & Borbely, A. A. (1995). Varying Levy, O., Dayan, T., & Kronfeld-Schor, N. (2007). The rela-
photoperiod in the laboratory rat: Profound effect on 24-h tionship between the golden spiny mouse circadian system
sleep pattern but no effect on sleep homeostasis. American and its diurnal activity: An experimental field enclosures
Journal of Physiology, 269, R691–R701. and laboratory study. Chronobiology International, 24,
Gattermann, R., Johnston, R. E., Yigit, N., Fritzsche, P., 599–613.
Larimer, S., Ozkurt, S., et al. (2008). Golden hamsters are Liu, X., Uchiyama, M., Shibui, K., Kim, K., Kudo, Y.,
nocturnal in captivity but diurnal in nature. Biology Letters, Tagaya, H., et al. (2000). Diurnal preference, sleep habits,
4, 253–255. circadian sleep propensity and melatonin rhythm in healthy
Haskell, E. H., Walker, J. M., & Berger, R. J. (1979). Effects human subjects. Neuroscience Letters, 280, 199–202.
of cold stress on sleep of an hibernator, the golden-mantled Meijer, J. H., & Rietveld, W. J. (1989). Neurophysiology of the
ground squirrel (C. lateralis). Physiology and Behavior, 23, suprachiasmatic circadian pacemaker in rodents. Physiolog-
1119–1121. ical Reviews, 69, 671–707.
Huber, R., Deboer, T., & Tobler, I. (2000a). Effects of sleep Mistlberger, R. E. (2005). Circadian regulation of sleep in
deprivation on sleep and sleep EEG in three mouse strains: mammals: Role of the suprachiasmatic nucleus. Brain
Empirical data and simulations. Brain Research, 857, 8–19. Research. Brain Research Reviews, 49, 429–454.
Huber, R., Deboer, T., & Tobler, I. (2000b). Topography of Mistlberger, R. E., Bergmann, B. M., Waldenar, W., &
EEG dynamics after sleep deprivation in mice. Journal of Rechtschaffen, A. (1983). Recovery sleep following sleep
Neurophysiology, 84, 1888–1893. deprivation in intact and suprachiasmatic nuclei-lesioned
Huber, R., Ghilardi, M. F., Massimini, M., & Tononi, G. rats. Sleep, 6, 217–233.
(2004). Local sleep and learning. Nature, 430, 78–81. Mistlberger, R. E., Landry, G. J., & Marchant, E. G. (1997).
Ibuka, N., Nihonmatsu, I., & Sekiguchi, S. (1980). Sleep- Sleep deprivation can attenuate light-induced phase shifts
wakefulness rhythms in mice after suprachiasmatic nucleus of circadian rhythms in hamsters. Neuroscience Letters,
lesions. Waking and Sleeping, 4, 167–173. 238, 5–8.
Jenkins, A., Archer, S. N., & von Schantz, M. (2005). Expan- Naylor, E., Bergmann, B. M., Krauski, K., Zee, P. C.,
sion during primate radiation of a variable number tandem Takahashi, J. S., Vitaterna, M. H., et al. (2000). The circa-
repeat in the coding region of the circadian clock gene dian clock mutation alters sleep homeostasis in the mouse.
period3. Journal of Biological Rhythms, 20, 470–472. The Journal of Neuroscience, 20, 8138–8143.
218

Schmidt, C., Collette, F., Leclercq, Y., Sterpenich, V., Trachsel, L., Edgar, D. M., Seidel, W. F., Heller, H. C., &
Vandewalle, G., Berthomier, P., et al. (2009). Homeostatic Dement, W. C. (1992). Sleep homeostasis in suprachiasmatic
sleep pressure and responses to sustained attention in the nuclei-lesioned rats: Effects of sleep deprivation and
suprachiasmatic area. Science, 324, 516–519. triazolam administration. Brain Research, 589, 253–261.
Shiromani, P. J., Xu, M., Winston, E. M., Shiromani, S. N., van der Horst, G. T., Muijtjens, M., Kobayashi, K.,
Gerashchenko, D., & Weaver, D. R. (2004). Sleep rhythmic- Takano, R., Kanno, S., Takao, M., et al. (1999). Mammalian
ity and homeostasis in mice with targeted disruption of Cry1 and Cry2 are essential for maintenance of circadian
mPeriod genes. American Journal of Physiology. Regulatory, rhythms. Nature, 398, 627–630.
Integrative and Comparative Physiology, 287, R47–R57. Van Twyver, H. (1969). Sleep patterns in five rodent species.
Strijkstra, A. M. (1999). Periodic euthermy during hibernation Physiology and Behavior, 4, 901–905.
in the European ground squirrel: Causes and consequences. Vitaterna, M. H., King, D. P., Chang, A. M.,
Wageningen: Ponsen & Looiyen. Kornhauser, J. M., Lowrey, P. L., McDonald, J. D., et al.
Strijkstra, A. M., & Daan, S. (1998). Dissimilarity of slow-wave (1994). Mutagenesis and mapping of a mouse gene, Clock,
activity enhancement by torpor and sleep deprivation in essential for circadian behavior. Science, 264, 719–725.
a hibernator. American Journal of Physiology, 275, Vyazovskiy, V. V., Achermann, P., & Tobler, I. (2007). Sleep
R1110–R1117. homeostasis in the rat in the light and dark period. Brain
Strijkstra, A. M., Meerlo, P., & Beersma, D. G. (1999). Forced Research Bulletin, 74, 37–44.
desynchrony of circadian rhythms of body temperature and Vyazovskiy, V. V., Olcese, U., Lazimy, Y. M., Faraguna, U.,
activity in rats. Chronobiology International, 16, 431–440. Esser, S. K., Williams, J. C., et al. (2009). Cortical firing
Susic, V., & Masirevic, G. (1986). Sleep patterns in the Mon- and sleep homeostasis. Neuron, 63, 865–878.
golian gerbil, Meriones unguiculatus. Physiology and Behav- Vyazovskiy, V. V., Ruijgrok, G., Deboer, T., & Tobler, I.
ior, 37, 257–261. (2006). Running wheel accessibility affects the regional elec-
Tafti, M., Franken, P., Kitahama, K., Malafosse, A., troencephalogram during sleep in mice. Cerebral Cortex, 16,
Jouvet, M., & Valatx, J. L. (1997). Localization of candidate 328–336.
genomic regions influencing paradoxical sleep in mice. Vyazovskiy, V. V., Welker, E., Fritschy, J. M., & Tobler, I.
Neuroreport, 8, 3755–3758. (2004). Regional pattern of metabolic activation is reflected
Tagaya, H., Uchiyama, M., Shibui, K., Kim, K., Suzuki, H., in the sleep EEG after sleep deprivation combined with uni-
Kamei, Y., et al. (2002). Non-rapid-eye-movement sleep lateral whisker stimulation in mice. European Journal of
propensity after sleep deprivation in human subjects. Neuro- Neuroscience, 20, 1363–1370.
science Letters, 323, 17–20. Wehr, T. A., Moul, D. E., Barbato, G., Giesen, H. A.,
Tobler, I., & Borbely, A. A. (1986). Sleep EEG in the rat as a Seidel, J. A., Barker, C., et al. (1993). Conservation of pho-
function of prior waking. Electroencephalography and Clini- toperiod-responsive mechanisms in humans. American Jour-
cal Neurophysiology, 64, 74–76. nal of Physiology, 265, R846–R857.
Tobler, I., Borbely, A. A., & Groos, G. (1983). The effect of Werth, E., Achermann, P., & Borbely, A. A. (1996a). Brain
sleep deprivation on sleep in rats with suprachiasmatic topography of the human sleep EEG: Antero-posterior
lesions. Neuroscience Letters, 42, 49–54. shifts of spectral power. Neuroreport, 8, 123–127.
Tobler, I., & Deboer, T. (2001). Sleep in the blind mole rat Werth, E., Dijk, D. J., Achermann, P., & Borbely, A. A.
Spalax ehrenbergi. Sleep, 24, 147–154. (1996b). Dynamics of the sleep EEG after an early evening
Tobler, I., & Franken, P. (1993). Sleep homeostasis in the nap: Experimental data and simulations. American Journal
guinea pig: Similar response to sleep deprivation in the light of Physiology, 271, R501–R510.
and dark period. Neuroscience Letters, 164, 105–108. Wisor, J. P., O’Hara, B. F., Terao, A., Selby, C. P.,
Tobler, I., Franken, P., Alfoldi, P., & Borbely, A. A. (1994). Kilduff, T. S., Sancar, A., et al. (2002). A role for
Room light impairs sleep in the albino rat. Behavioural cryptochromes in sleep regulation. BMC Neuroscience, 3,
Brain Research, 63, 205–211. 20.
Tobler, I., & Jaggi, K. (1987). Sleep and EEG spectra in the Yasenkov, R., & Deboer, T. (2010). Circadian regulation of
Syrian hamster (Mesocricetus auratus) under baseline sleep and the sleep EEG under constant sleep pressure in
conditions and following sleep deprivation. Journal of Com- the rat. Sleep, 33, 631–641.
parative Physiology A, 161, 449–459. Yasenkov, R., & Deboer, T. (2011). Interrelations and circa-
Toh, K. L., Jones, C. R., He, Y., Eide, E. J., Hinz, W. A., dian changes of electroencephalogram frequencies under
Virshup, D. M., et al. (2001). An hPer2 phosphorylation site baseline conditions and constant sleep pressure in the rat.
mutation in familial advanced sleep phase syndrome. Sci- Neuroscience, 180, 212–221.
ence, 291, 1040–1043.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 13

Local aspects of sleep: Observations from


intracerebral recordings in humans

Lino Nobili{,#,*, Luigi De Gennaro{, Paola Proserpio{, Fabio Moroni{,},


Simone Sarasso||, Andrea Pigorini||, Fabrizio De Carli# and Michele Ferrara**

{
Centre of Epilepsy Surgery “C. Munari,” Center of Sleep Medicine, Niguarda Hospital, Milan, Italy
{
Department of Psychology, University of Rome “Sapienza,” Roma, Italy
}
Department of Psychology, University of Bologna, Bologna, Italy
||
Department of Clinical Sciences L. Sacco Università degli Studi di Milano, Milan, Italy
#
Institute of Bioimaging and Molecular Physiology, Genoa Unit, National Research Council, Genoa, Italy
** Department of Health Sciences, University of L’Aquila, L’Aquila, Italy

Abstract: Human sleep is considered a global phenomenon, orchestrated by central specialized neuronal
networks modulating the whole-brain activity. However, recent studies point to a local regulation of sleep.
Sleep disorders, such as sleepwalking, suggest that electroencephalographic (EEG) features of sleep and
wakefulness might be simultaneously present in different cerebral regions. Recently, intracranial EEG
recording techniques, mainly applied for the presurgical evaluation of drug-resistant epileptic patients,
have provided new and interesting information on the activity of different cortical and subcortical
structures during sleep in humans. In particular, it has been observed that the thalamus, during the
transition between wake and sleep undergoes a deactivation process that precedes the one occurring
within the cortex, with extensive cortical territories maintaining an activated pattern for several minutes
after the thalamic deactivation. Very recent intracerebral EEG studies have also shown that human
NREM sleep can be characterized by the coexistence of wake-like and sleep-like EEG patterns in
different cortical areas. Moreover, unit-firing recordings in multiple brain regions of neurosurgical
patients evidenced that most sleep slow waves and the underlying active and inactive neuronal states do
occur locally. These findings add a new dimension to the concept of local sleep regulation and opens new
perspectives in the interpretation of the substrates underlying behavioral states of vigilance. The
implications for sleep medicine are also discussed.

Keywords: local sleep; arousal; cortical activations; hippocampus; parasomnias; dissociated state; sleep
onset.

*Corresponding author.
Tel.: þ 39-0264442917-2918; Fax: þ 39-0264442868
E-mail: lino.nobili@ospedaleniguarda.it

http://dx.doi.org/10.1016/B978-0-444-59427-3.00013-7 219
220

Introduction Werth et al., 1996). Moreover, it has been demon-


strated that sleep EEG delta power can be locally
Human sleep is traditionally considered a global increased or reduced as a function of previous
phenomenon orchestrated by central specialized waking activity (Cantero et al., 2002; Huber et al.,
neuronal networks modulating whole-brain activ- 2004, 2006).
ity. Specifically, the reduced reticular formation Recently, the spread of intracranial EEG record-
tone is interpreted as the main determinant of ing techniques, mainly exploited for the presurgical
the electroencephalographic (EEG) synchroniza- evaluation of drug-resistant epileptic patients,
tion characterizing the sleep state (Fuller et al., allowed to overcome the intrinsic spatial limitations
2011; Moruzzi and Magoun, 1949). of scalp EEG techniques, providing detailed infor-
However, it has been recently proposed that mation on the activity of different cortical and sub-
sleep is local in nature, being a fundamental prop- cortical structures, both during wake and sleep
erty of small neuronal groups (Krueger and Obál, (Moroni et al., 2008, 2012).
1993; Krueger et al., 2008). According to this view, The aim of the present work is to review the avail-
global sleep (behaviorally and electroencephalo- able published data regarding local aspects of sleep
graphically defined) emerges when the altered supported by novel observations from our and
input–output state, that characterizes the sleep- other research groups, obtained using scalp high
like state at the local network level, involves a density and intracranial EEG techniques. Finally,
large and widespread number of cortical regions capitalizing on these observations, we will propose
(Krueger et al., 2008). novel insights for sleep medicine.
Moreover, experimental studies have shown
that sleep and wakefulness can be restricted to
small groups of neurons (Pigarev et al., 1997), to From wake to sleep—A behavioral transition
individual cortical columns, (Rector et al., 2005) through local brain changes
or to larger brain regions. For instance, some birds
and marine mammals can simultaneously exhibit A complex and widely distributed neural system,
sleep in one cerebral hemisphere and wakefulness with functionally distinct but integrated
in the other one in order to continue flying or swim- components, orchestrates sleep (Rosenwasser,
ming (Lyamin et al., 2008; Mukhametov, 1984). 2009). Schematically, there are three main sleep
To a broader extent, clinical evidence deriving regulatory processes: (1) the circadian process
from patients affected by sleep disorders, such as (Process C) that couples the timing for sleep and
sleepwalking, suggests that sleep and wakefulness wakefulness with the light–dark cycle; (2) the
might be simultaneously present in different homeostatic process (Process S), modulating sleep
cerebral regions indicating that the boundaries intensity and strongly dependent on wakefulness
between these different behavioral states are not duration and activity; and (3) the ultradian process,
strictly defined (Mahowald and Schenck, 2005). which regulates the intra-sleep NREM–REM
According to these observations, brain-sleep alternation (Borbély, 1982; Borbély and
state may be spatially non-uniform, and along Achermann, 1999; Ferrillo et al., 2007). In normal
the same line, sleep and wakefulness may not be conditions, the interaction of these processes
temporally discrete behavioral states. As an exam- allows an unambiguous separation between the
ple, scalp EEG studies have shown that large state of wakefulness and sleep, from both a behav-
regional frequency-specific EEG differences char- ioral and a neurophysiological perspective. Indeed,
acterize NREM sleep, with a frontal predominance during the aroused state, when we are fully respon-
of delta activity (1–4.5 Hz), the main index of sleep sive and alert, cortical and thalamocortical neurons
intensity (Ferrara et al., 2002; Finelli et al., 2001; display fast and complex patterns of activation;
221

when we fall asleep and become progressively less during recording sessions. These results indicated
responsive, this activated state is in turn replaced that, taking into account the neuronal firing pat-
by pervasive and synchronized low-frequency tern, sleep onset (SO) may occur in a strictly local
oscillations of neural activity (Steriade et al., manner; the animal can be behaviorally awake,
1993). This phenomenon is reflected in the interrup- but local cortical “islands of sleep” can appear.
tion of fast rhythms (alpha and beta activity) and the In a very recent experimental study, Vyazovskiy
appearance of slow waves on the EEG. These dra- et al. (2011) focused their analysis on the presence
matic changes in behavioral and electrophysiologi- of slow EEG activity during wakefulness. Slow
cal states, which in humans may take from few EEG rhythms (< 4 Hz) represent a typical feature
seconds to several minutes, require specific of NREM sleep reflected, at the cellular level, by
switching mechanisms in the brain, enabling rela- a bistable behavior of cortical neurons, alternating
tively rapid and complete state transitions (Fuller between two distinct states: one characterized by
et al., 2006). depolarized membrane potential with burst of
However, recent experimental evidence in action potentials (upstate) and the other associated
animals and humans demonstrated that this transi- with hyperpolarized membrane potential and neu-
tion mechanism might work with different timing ronal silence (downstate) (Contreras and Steriade,
for different brain areas, partially redefining the 1995; Crunelli and Hughes, 2010). Exploiting
classical distinction between wake and sleep as multiunit activity recordings, Vyazovskiy et al.
separate, incompatible states. (2011) showed that, in sleep-deprived but awake
rats, relatively small populations of neurons in
different cortical areas may suddenly manifest
Sleeping locally during wakefulness “downstates”—resembling those typical of NREM
sleep—associated with low-frequency ( 2 Hz)
Under extreme conditions, as is the case of pro- EEG activity and deficit in performance at a pellet
longed wakefulness, brain activity undergoes reaching task. Interestingly, these events occurred
dramatic changes. Classically, the EEG traces in a strictly local fashion, involving only a small
gradually show a progressive increase in the group of neurons at a time. In other words, subsets
amount of slow signal components (delta–theta of neurons may enter an “off” period in one corti-
activity) as a function of time spent awake, cal area but not in another and, even within
(Cajochen et al., 2002; De Gennaro et al., 2007; the same cortical area, some neurons may be
Finelli et al., 2000; Franken et al., 1991; Strijkstra “off” while others remain “on.” Finally, such local
et al., 2003) reflective of increased homeostatic “off periods” increased with the duration of wake-
sleep pressure (Process S). Experimental studies fulness, suggesting a relationship with a local
suggest that the progressive slowing of waking increase of sleep pressure.
EEG may be associated with the appearance of The authors interpret these findings as possibly
local sleep states during wakefulness. Indeed, it representing a form of neuronal tiredness due to
has been shown that in sleep-deprived monkeys use-dependent factors, such as synaptic overload
performing a visual discrimination task, some (Tononi and Cirelli, 2006). This view assumes
neurons in the extrastriate visual cortex display that local, use (and possibly learning)-dependent
the characteristic sleep firing pattern (bursting increase of synaptic strength is followed by a
mode), while neurons in the striate cortex show corresponding local increase in subsequent sleep
the waking pattern of activation (tonic firing) pressure, reflected in increased EEG slow-wave
(Pigarev et al., 1997). These episodes of “local activity (SWA). Along the same line, after
sleep” were much more frequent in the animal that showing that a homeostatic regulation of SWA
apparently experienced stronger sleep pressure is present also in subcortical structures such as
222

the hippocampus (Moroni et al., 2007), we posterior one observed during the presleep phase
recently performed intracerebral EEG recordings (De Gennaro et al., 2004, 2005). The local nature
of the hippocampal formation investigating the of EEG and corticographic changes during the
effects of both procedural and declarative wake-sleep transition and the existence of an
learning on the qualitative and quantitative antero-posterior gradient are coherent with the
measures of hippocampal sleep EEG (Moroni spatiotemporal dynamics of the < 1 Hz slow oscil-
et al., 2008). We reasoned that, if the hippocampus lation during full-fledged NREM sleep. This
is directly involved in the memory consolidation EEG activity originates more frequently at the
process, in agreement with the local use-dependent anterior cortical regions and propagates in an
theory, significant changes in the hippocampal antero-posterior direction, with a highly reproduc-
EEG slow rhythms should be expected. We indeed ible origin, speed, and direction (Massimini et al.,
showed that an intensive training on a sequential 2004).
finger-tapping task is followed by an increase in Electrocorticographic recordings of 13 epileptic
the amount of scalp recorded SWS, as well as patients undergoing presurgical assessment (Bódizs
in the hippocampal EEG power in the very low- et al., 2005) have confirmed these scalp observations.
frequency range (0.5–1.0 Hz) during the first During the wake–sleep transition, cortical delta
postlearning NREM period. The magnitude of activity showed an antero-posterior gradient with
procedural performance improvement was signifi- the frontal power peaking earlier than the temporal
cantly correlated with the subsequent EEG low- and the temporal earlier than the occipital. The
frequency power increase. This supports the time course of parahippocampal changes in delta
existence of a direct link between procedural activity during the transition paralleled those of the
memory consolidation and very low hippocampal frontal cortex (Hangya et al., 2011).
EEG rhythms. To our best knowledge, these are Although an asynchronous SO between differ-
the first reported subcortical (local) effects of ent cortical areas has been repeatedly demonstra-
learning on sleep EEG recorded from both scalp ted; only very recently it has been shown in
and intracerebral derivations. humans that the thalamus and the cortical mantle
are not strictly coupled during the wake–sleep
transition, as firmly believed in the past (Magnin
Falling asleep locally et al., 2010). Indeed, Magnin et al. (2010), using
intracerebral recordings in drug-resistant epileptic
So far, very few studies in humans have analyzed patients, showed that the thalamus, during physi-
the EEG changes during the wake–sleep transi- ological SO, undergoes a deactivation process
tion, and most of these were based on scalp that precedes the one occurring within the cortex.
recordings. They clearly differentiate topographi- In particular, the authors observed that extensive
cal EEG modifications showing, after SO, a cortical territories maintained an activated pat-
fronto-central predominance of delta/theta activ- tern for several minutes after the thalamic deacti-
ity, a centro-parietal maxima of the sigma activity, vation. Moreover, the study confirmed that the
and an anteriorization of the alpha activity (De cortical deactivation showed marked topographi-
Gennaro et al., 2001; Werth et al., 1996). The cal differences at SO both within and between
whole pattern indicates that the anterior areas patients.
are the first to show a synchronized EEG activity, Based on these findings, in order to explore the
while the occipital areas are the last ones (De possibility of transient decoupling between limbic
Gennaro et al., 2001). Moreover, SO seems to be and neocortical structures during wake–sleep
characterized by a posterior-to-anterior functional transition, we investigated simultaneous intrace-
cortical coupling replacing the anterior-to- rebral hippocampal and neocortical recordings in
223

five patients with refractory epilepsy. Our prelim- activity as a function of time spent awake in the
inary results (Sarasso et al., in preparation) show hippocampus compared to other neocortical
that several minutes before SO (detected on scalp areas.
EEG by the emergence of the first spindle or
K-complex), spindles appeared in the hippocam-
pus with the typical 4 s periodicity. Moreover, in Local cortical awakenings within the
one of our patients, we had the possibility to analyze sleeping brain
also the thalamic EEG activity. We observed that
during wakefulness the thalamus (in particular, the Since local “islands of sleep” can appear during
anterior thalamic nucleus) is characterized by a fast wakefulness, could also “local wakefulness” appear
rhythm, with a peak around 25 Hz (Fig. 1, upper during sleep?
panel). Few minutes before SO, periodic slow In order to probe the coexistence of dissociated
oscillations in the low-delta range (< 2 Hz) appeared (wake-like and sleep-like) electrophysiological
in the thalamic EEG recording, almost simulta- behaviors within the sleeping brain, we recently
neously with the emergence of hippocampal spindle analyzed intracerebral EEG activity drawn
activity (Fig. 1, middle panel). Quantitative EEG from sleep recordings of five patients with
analysis by fast fourier transform confirmed the pharmacoresistant focal epilepsy without sleep
observations derived from visual analysis. Figure 2 disturbances, who underwent presurgical intrace-
shows the time course of normalized absolute values rebral EEG investigation (Nobili et al., 2011).
of spectral power in the slow-delta (0.5–2.0 Hz) We observed that sleep can be characterized by
and sigma (12.0 –16.0 Hz) frequency range. Several the coexistence of wake-like and sleep-like EEG
minutes before SO, a progressive raise in the patterns in different cortical areas, as indicated
thalamic delta power is evident; when such an by a high number of local activations (abrupt
increase becomes steeper, also the hippocampal increases in EEG frequency, including alpha
delta power progressively increases. Interestingly, and/or beta rhythm) within the motor cortex
the appearance of spindles precedes the delta power (Mc), occurring synchronously with deep-sleep
raise in the hippocampal EEG recording. EEG patterns in the dorsolateral prefrontal cor-
Magnin et al. (2010) hypothesized that the tex (dlPFc) and scalp (Nobili et al., 2011). In par-
delay between thalamic and cortical deactivation ticular, we found that the physiological and
at SO could reflect a thalamus-driven process or progressive decay of SWA across NREM sleep
a higher sensitivity of the thalamus to firing cycles was comparable between the derivations
patterns of brainstem and hypothalamic afferents. (dlPFc, Mc, scalp) investigated. However, in spite
This hypothesis could explain also our findings of this similarity, the Mc showed a higher rate of
regarding a limbic structure, such as the hippo- local activations during NREM sleep compared
campus. In the future, it would be interesting to to the dlPFc and scalp. Moreover, the local
evaluate whether these features are common to activations in the Mc increased across sleep cycles
other archi-cortical regions. This would suggest and within each cycle toward the end of each
that the phylogenetic development of a cortical NREM period, being possibly associated, respec-
structure influences its sensitivity to the brainstem tively, with the decrease of SWA and the
and hypothalamic SO switch (Fuller et al., 2006). approach of REM sleep. Interestingly, recent
On the other hand, we cannot exclude that the observations derived from an intracerebral study
early hippocampal SO reflects an increased pres- conducted in epileptic patients showed that the
sure of the homeostatic process in this structure. spatial dynamics of cortical slow oscillations,
According to this alternative interpretation, one the hallmark of full-fledge NREM sleep, change
should find a higher relative low-frequency throughout the night, involving the entire cortical
Th

Hc

P cx

Scalp

EOG

Chin

EKG

Th

Hc

P cx

Scalp

EOG

Chin

EKG

Th

Hc

P cx

Scalp

EOG

Chin

EKG

1s

Fig. 1. Three 30-s epochs of intracerebral and scalp EEG recordings. Upper panel: patient behaviorally awake; all the EEG traces
show a typical waking pattern of activity. Middle panel: patient behaviorally awake, 10 min before sleep onset; appearance of
periodic slow oscillations in the low-delta range (<2 Hz) within the thalamic EEG trace and of sleep spindles in the
hippocampus. The parietal cortex and scalp EEG maintain a waking pattern of activity. Lower panel: stage 2 of NREM sleep; all
the EEG traces show a NREM sleep pattern of activity. Abbreviations: Th, thalamus; Hc, hippocampus; P cx, parietal cortex;
Scalp, scalp EEG (Fz-Cz); EOG, electroculogram; Chin, submental electromyogram; EKG, electrocardiogram.
225

250
Scalp EEG Thalamus Hippocampus Parietal cortex

12.0–16.0 Hz

Spectral power (normalized values)


200

150

100

50

0
200

180
0.5–2.0 Hz
Spectral power (normalized values)

160

140

120

100

80

60

40

20 N2
N1
W
0
0 5 10 15 20 25 30
Time (min)

Fig. 2. Time course of normalized absolute values of spectral power in the sigma (12.0–16.0 Hz, upper panel) and slow-delta range
(0.5–2.0 Hz, lower panel), in a patient with both intracerebral and scalp EEG derivations. Several minutes before the sleep onset (as
indicated by the start of N2 sleep), a progressive raise in the thalamic delta power is evident; when such an increase becomes
steeper, also the hippocampal delta power progressively increases. Abbreviations: W, wakefulness; N1, stage 1 of NREM sleep;
N2, stage 2 of NREM sleep.

mantle in early sleep and only limited cortical sometimes last up to 120 s and more than 50% of
areas toward the end of sleep (Nir et al., 2011). them were not accompanied by EMG activation.
In general, the Mc local activations (LAs) had an They could occur as isolated or cyclical events.
about  10 s duration (Fig. 3), but they could When LAs occur in a cyclical fashion, they strongly
226

(a) 250 mV
Mc
250 mV
dIPFc
100 mV
Fz-Cz

100 mV

Eog
100 mV

Chin

(b)
250 mV (c)
Mc
250 mV

dIPFc
100 mV

Fz-Cz
100 mV

Eog
100 mV

Chin

10 s

Fig. 3. Sample patterns of intracerebral and scalp EEG recordings. (a) NREM sleep: slow waves prevail in all the three EEG traces;
(b) and (c): NREM sleep: local activations appearing in the motor cortex accompanied by bursts of slow waves in the other two
EEG derivations. Abbreviations: Mc, motor cortex; dlPFc, dorsolateral prefrontal cortex; Fz-Cz, bipolar scalp EEG; EOG,
electroculogram; Chin, submental electromyogram.

resemble the scalp EEG pattern corresponding to the sensory-motor processing (McCormick and
the definition of the cyclic alternating pattern Feeser, 1990). Our data suggest that this state-
(CAP; Parrino et al., 2012; Fig. 4). switching can be confined to restricted thalamo-
Around most LAs, the time–frequency analysis cortical pathways.
of the EEG showed an increase of high-frequency On the basis of our results, we hypothesized
EEG activity (mainly in the 8–12 Hz frequency that a lower arousal threshold and a higher level
range) in the Mc, paralleled by an increase of low- of activation in the Mc during NREM sleep may
frequency EEG activities (peaking at 0.5–2 Hz) in have been evolutionary selected, because they
the dlPFc and scalp sites (Fig. 5). increase the probability of survival by facilitating
In the transition from sleep to the aroused motor behaviors in case of sudden awakenings
state, the ascending activating systems abolish (Nobili et al., 2011). Accordingly, the significant
the slow rhythmic burst firing mode in thalamo- enhancement of slow frequencies in the dlPFc
cortical circuits and induce the appearance of a immediately before and during the Mc LAs
high-frequency tonic firing pattern (Moruzzi and could be interpreted as a mechanism preventing
Magoun, 1949; Steriade et al., 1993). This allows a full awakening and allowing sleep as a global
the brain to recover full responsiveness enhancing process to continue uninterrupted, even when a
227

GC
dlPfcx
sF gyrus
mF gyrus
SMA
Mcx
Mcx
Pcx
Supr. gyrus
PC
PC
Intr. P sulcus
Is GC
Sup P lob
Precuneus
Scalp EEG
EOG
Chin
EKG

GC
dlPfcx
sF gyrus
mF gyrus
SMA
Mcx
Mcx
Pcx
Supr. gyrus
PC
PC
Intr. P sulcus
Is GC
Sup P lob
Precuneus
Scalp EEG
EOG
Chin
EKG

5s

Fig. 4. Sample patterns of intracerebral and scalp EEG recordings. Cyclical appearance of local activations in the supplementary
motor area, motor cortex, and parietal cortex. Abbreviations: GC, gyrus cinguli; dlPfcx, dorsolateral prefrontal cortex; sF gyrus,
superior frontal gyrus; mF gyrus, middle frontal gyrus; SMA, supplementary motor area; Mcx, motor cortex; Pcx, parietal cortex;
Supr. gyrus, supramarginal gyrus; PC, parietal cingulus; Intr. P sulcus, intraparietal sulcus; Is GC, isthmus of cingulate gyrus; Sup
P lob, superior parietal lobe; EOG, electroculogram; Chin, submental elecromyogram; EKG, electrocardiogram.

local activation appears. Our data are in accor- Local aspects of sleep—Implications for sleep
dance with results obtained from PET studies, medicine
which showed that the cortical deactivation
during NREM sleep is not a homogeneous phe- Considering sleep not only as a global phenome-
nomenon, with the associative cortices of the fron- non but also as a locally regulated process might
tal and parietal lobes being the most deactivated provide novel understanding also in the field of
areas, while the primary cortices are the least sleep medicine.
deactivated (Braun et al., 1997; Dang-Vu et al., It is well known that sleep deprivation can
2005). Finally, our findings are also coherent affect various aspects of performance (Bocca and
with the observation of similar cortical excitability Denise, 2006) reducing attention and vigilance,
of human primary Mc, as assessed by transcranial decision-making ability and memory functions
magnetic stimulation, in wake and NREM sleep (Curcio et al., 2006; Harrison and Horne, 2000;
(De Gennaro et al., 2007). Killgore et al., 2006).
228

16–32
1
8–12
Hz

2–4 0.5
0.5–1
0

16–32
1
8–12
Hz

2–4 0.5
0.5–1
0

16–32
1
8–12
Hz

2–4 0.5
0.5–1
0
-20 -16 -12 -8 -4 0 4 8 12 16 20
sec

Fig. 5. Time–frequency distribution of EEG patterns recorded in Motor cortex (upper part), dorsolateral prefrontal cortex (middle
part), and scalp (bottom part) in association with motor cortex local activations (LAs) in a single subject. A 40-s time window was
set around the start (zero time) of each Mc LA and signal power was calculated by wavelet transform for each time unit within a
time–frequency grid (0.125 s time resolution and frequency-dependent bandwidth). Time–frequency distribution was averaged
within each recording. The distribution of averaged signal amplitude (square root of the power) is presented for each brain
region by a color scale. The abrupt shift toward higher frequencies (8–12 Hz; 16–32 Hz) in the motor cortex is accompanied by
an increase of EEG activity in the low frequencies (0.5–2.0 Hz) in the other regions, in association with the onset of the Mc local
activation.

We might hypothesize that cognitive impairment of sleep on scalp EEG might explain specific mem-
and performance deficits induced by sleep depriva- ory alteration or amnesia for events preceding the
tion could be due to the occurrence of cortical and unequivocal occurrence of behavioral SO (Wyatt
subcortical local “islands of sleep” in behaviorally et al., 1994, 1997).
fully awake subjects. In the future, the application The different temporal dynamics of state-
of high-density EEG in sleep-deprived subjects synchronization in different cortical areas (as
may allow the detection of subtle signs of “neuronal observed in our data and by Magnin et al., 2010)
tiredness” expressed by local slow waves. Consider- could explain the frequent dissociations between
ing the fundamental involvement of the hippocam- EEG and behavioral or subjective measures of
pus in memory processing, the observation that this SO (Ogilvie and Wilkinson, 1988). Indeed, since
subcortical structure displays typical sleep features sleep does not simultaneously begin in all cortical
(i.e., spindles) many minutes before the occurrence and subcortical areas, it is not surprising that the
229

subjective definition of SO does not fit with scalp awareness though they can show automatic or
EEG electrophysiological observations. In the more organized motor behaviors. During these
near future, coregistering patients with intracere- events, scalp EEG generally shows high-amplitude
bral electrodes and high-density EEG, together slow waves, mainly expressed over the frontal
with the assessment of behavioral and vegetative regions, together with alpha and beta activity
variables (EKG, heart rate variability, respiration, (Schenck et al., 1989; Zadra et al., 2004). Con-
etc.), could allow extracting noninvasive parameters firming our prediction, we recently captured a
able to sensibly detect early signs of sleepiness even confusional arousal in a subject affected by drug-
in confined regions. This could be extremely useful resistant epilepsy who was undergoing an intrace-
for the prevention of accidents related to excessive rebral EEG investigation during presurgical
sleepiness in the general population. assessment. Intracerebral EEG showed the pres-
In the clinical practice, an impairment in the ence of local activations of the motor and cingulate
process of state-synchronization among different cortices associated with increased delta activity in
cortical regions, with cortical territories that the frontoparietal associative cortices (Terzaghi
remain activated for long time after the thalamic et al., 2009).
and hippocampus deactivation, could explain Currently, we have demonstrated that the
both long sleep latency in insomniac patients and appearance of local dissociated states is an intrin-
the mismatch between subjective and objective sic feature of NREM sleep (Nobili et al., 2011),
measures frequently observed in subjects with par- supporting the hypothesis that paradoxical phe-
adoxical insomnia (Manconi et al., 2010; Marzano nomena could be the result of a coexistence of
et al., 2008; Parrino et al., 2009). In particular, in different local states: an aroused state (very simi-
these patients, it has been observed that the per- lar to wakefulness) and sleep. Genetic factors or
ception of a nonrestorative sleep can be correlated external triggers such as sleep deprivation (Zadra
with a disproportionate increase of sleep instabil- et al., 2008), modifying the arousal threshold, may
ity, expressed by a high CAP rate also in the pres- favor the persistence of local dissociated activity
ence of normal sleep duration compared to healthy allowing the appearance of motor phenomena as
controls (Parrino et al., 2012; Stone et al., 2008; those observed in NREM parasomnias.
Terzano et al., 2003). We suggest that an increase Finally, if the aroused state can appear locally
in sleep instability, expressed by a disproportional during sleep, we cannot exclude that such a dis-
number of K-complexes and delta bursts, could sociation may also occur during awakenings (in
be related to the presence of unrecognized local this case with cortical areas showing local sleep
cortical activations characterized by higher EEG features). This could explain other phenomena
frequencies (Fig. 4). A high number of local corti- such as sleep inertia (Ferrara et al., 2006;
cal activations could alter the normal evolution of Marzano et al., 2011), the subjective feeling of
sleep playing an important role in the perception grogginess accompanied by decreased levels of
of a nonrestorative sleep or in the genesis of mis- cognitive and behavioral performance which typ-
perception insomnia. ically follows the awakening.
The appearance of localized simultaneous sleep- In conclusion, intracerebral EEG recordings
like and wake-like activity during NREM sleep offer a unique opportunity to investigate simulta-
may also explain paradoxical sleep phenomena neously several cortical and subcortical structures
such as sleepwalking and confusional arousal, during the entire vigilance spectrum, going from
which have been attributed to a breakdown of active wakefulness to deep sleep. Due to the inva-
the boundaries between wakefulness and NREM sive nature of the technique, it cannot be applied
sleep (Mahowald and Schenck, 2005). Indeed, dur- to study sleep physiology or sleep disorders in the
ing these phenomena the subjects lack of conscious general population.
230

In the future, by coupling intracerebral and De Gennaro, L., Marzano, C., Veniero, D., Moroni, F.,
scalp hd-EEG recordings, we can obtain novel Fratello, F., Curcio, G., et al. (2007). Neurophysiological
correlates of sleepiness: A combined TMS and EEG study.
and more precise interpretation of the nature NeuroImage, 36, 1277–1287.
of scalp recorded events, therefore overcoming De Gennaro, L., Vecchio, F., Ferrara, M., Curcio, G.,
the intrinsic spatial limitation of surface EEG Rossini, P. M., & Babiloni, C. (2004). Changes in fronto-
recordings. posterior functional coupling during sleep onset in humans.
Journal of Sleep Research, 13, 209–218.
De Gennaro, L., Vecchio, F., Ferrara, M., Curcio, G.,
References Rossini, P. M., & Babiloni, C. (2005). Antero-posterior func-
tional coupling at sleep onset: Changes as a function of
Bocca, M. L., & Denise, P. (2006). Total sleep deprivation increased sleep pressure. Brain Research Bulletin, 65, 133–140.
effect on disengagement of spatial attention as assessed Ferrara, M., Curcio, G., Fratello, F., Moroni, F., Marzano, C.,
by saccadic eye movements. Clinical Neurophysiology, 117, Pellicciari, M. C., et al. (2006). The electroencephalographic
894–899. substratum of the awakening. Behavioural Brain Research,
Bódizs, R., Sverteczki, M., Lázár, A. S., & Halász, P. (2005). 167, 237–244.
Human parahippocampal activity: Non-REM and REM Ferrara, M., De Gennaro, L., Curcio, G., Cristiani, R.,
elements in wake-sleep transition. Brain Research Bulletin, Corvasce, C., & Bertini, M. (2002). Regional differences of
65, 169–176. the human sleep electroencephalogram in response to selec-
Borbély, A. A. (1982). A two process model of sleep regula- tive slow-wave sleep deprivation. Cerebral Cortex, 12,
tion. Human Neurobiology, 1, 195–204. 737–748.
Borbély, A. A., & Achermann, P. (1999). Sleep homeostasis Ferrillo, F., Donadio, S., De Carli, F., Garbarino, S., &
and models of sleep regulation. Journal of Biological Nobili, L. (2007). A model-based approach to homeostatic
Rhythms, 14, 557–568. and ultradian aspects of nocturnal sleep structure in narco-
Braun, A. R., Balkin, T. J., Wesenten, N. J., Carson, R. E., lepsy. Sleep, 30, 157–165.
Varga, M., Baldwin, P., et al. (1997). Regional cerebral blood Finelli, L. A., Baumann, H., Borbély, A. A., & Achermann, P.
flow throughout the sleep–wake cycle. Brain, 120, 1173–1197. (2000). Dual electroencephalogram markers of human
Cajochen, C., Wyatt, J. K., Czeisler, C. A., & Dijk, D. J. (2002). sleep homeostasis: Correlation between theta activity in
Separation of circadian and wake duration-dependent modu- waking and slow-wave activity in sleep. Neuroscience, 101,
lation of EEG activation during wakefulness. Neuroscience, 523–529.
114, 1047–1060. Finelli, L. A., Borbely, A. A., & Achermann, P. (2001). Func-
Cantero, J. L., Atienza, M., Salas, R. M., & Dominguez- tional topography of the human nonREM sleep electroen-
Marin, E. (2002). Effects of prolonged waking-auditory cephalogram. European Journal of Neuroscience, 13,
stimulation on electroencephalogram synchronization and 2282–2290.
cortical coherence during subsequent slow-wave sleep. Jour- Franken, P., Dijk, D. J., Tobler, I., & Borbely, A. A. (1991).
nal of Neuroscience, 22, 4702–4708. Sleep deprivation in rats: Effects on EEG power spectra,
Contreras, D., & Steriade, M. (1995). Cellular basis of EEG vigilance states, and cortical temperature. American Journal
slow rhythms: A study of dynamic corticothalamic of Physiology, 261, R198–R208.
relationships. Journal of Neuroscience, 15, 604–622. Fuller, P. M., Gooley, J. J., & Saper, C. B. (2006). Neurobiol-
Crunelli, V., & Hughes, S. W. (2010). The slow (<1 Hz) ogy of the sleep-wake cycle: Sleep architecture, circadian
rhythm of non-REM sleep: A dialogue between three cardi- regulation, and regulatory feedback. Journal of Biological
nal oscillators. Nature Neuroscience, 13, 9–17. Rhythms, 21, 482–493.
Curcio, G., Ferrara, M., & De Gennaro, L. (2006). Sleep loss, Fuller, P., Sherman, D., Pedersen, N. P., Saper, C. B., & Lu, J.
learning capacity and academic performance. Sleep Medi- (2011). Reassessment of the structural basis of the ascending
cine Reviews, 10, 323–337. arousal system. The Journal of Comparative Neurology, 519,
Dang-Vu, T. T., Desseilles, M., Laureys, S., Degueldre, C., 933–956.
Perrin, F., Phillips, C., et al. (2005). Cerebral correlates of Hangya, B., Tihanyi, B. T., Entz, L., Fabó, D., Erfss, L.,
delta waves during non-REM sleep revisited. NeuroImage, Wittner, L., et al. (2011). Complex propagation patterns
28, 14–21. characterize human cortical activity during slow-wave sleep.
De Gennaro, L., Ferrara, M., Curcio, G., & Cristiani, R. Journal of Neuroscience, 31, 8770–8779.
(2001). Antero-posterior EEG changes during the Harrison, Y., & Horne, J. A. (2000). The impact of sleep dep-
wakefulness-sleep transition. Clinical Neurophysiology, rivation on decision making: A review. Journal of Experi-
112, 1901–1911. mental Psychology: Applied, 6, 236–249.
231

Huber, R., Ghilardi, M. F., Massimini, M., Ferrarelli, F., wakefulness in the human hippocampus. NeuroImage, 60,
Riedner, B. A., Peterson, M. J., et al. (2006). Arm immobi- 497–504.
lization causes cortical plastic changes and locally Moruzzi, G., & Magoun, H. W. (1949). Brain stem reticular
decreases sleep slow wave activity. Nature Neuroscience, formation and activation of the EEG. Electroencephalogra-
9, 1169–1176. phy and Clinical Neurophysiology, 1, 455–473.
Huber, R., Ghilardi, M. F., Massimini, M., & Tononi, G. Mukhametov, L. M. (1984). Sleep in marine mammals. Exper-
(2004). Local sleep and learning. Nature, 430, 78–81. imental Brain Research, 8, 227–238.
Killgore, W. D., Balkin, T. J., & Wesensten, N. J. (2006). Nir, Y., Staba, R. J., Andrillon, T., Vyazovskiy, V. V.,
Impaired decision making following 49 h of sleep depriva- Cirelli, C., Fried, I., et al. (2011). Regional slow waves and
tion. Journal of Sleep Research, 15, 7–13. spindles in human sleep. Neuron, 70, 153–169.
Krueger, J. M., & Obál, F. (1993). A neuronal group theory of Nobili, L., Ferrara, M., Moroni, F., De Gennaro, L., Lo
sleep function. Journal of Sleep Research, 2, 63–69. Russo, G., Campus, C., et al. (2011). Dissociated wake-like
Krueger, J. M., Rector, D. M., Roy, S., Van Dongen, H. P., and sleep-like electro-cortical activity during sleep.
Belenky, G., & Panksepp, J. (2008). Sleep as a fundamental NeuroImage, 58, 612–619.
property of neuronal assemblies. Nature Reviews Neurosci- Ogilvie, R. D., & Wilkinson, R. T. (1988). Behavioral versus
ence, 9, 910–919. EEG-based monitoring of all-night sleep/wake patterns.
Lyamin, O. I., Manger, P. R., Ridgway, S. H., Sleep, 11, 139–155.
Mukhametov, L. M., & Siegel, J. M. (2008). Cetacean sleep: Parrino, L., Ferri, R., Bruni, O., & Terzano, M. G. (2012).
An unusual form of mammalian sleep. Neuroscience and Cyclic alternating pattern (CAP): The marker of sleep insta-
Biobehavioral Review, 32, 1451–1484. bility. Sleep Medicine Reviews, 16, 27–45.
Magnin, M., Rey, M., Bastuji, H., Guillemant, P., Parrino, L., Milioli, G., De Paolis, F., Grassi, A., &
Mauguiere, F., & Garcia-Larrea, L. (2010). Thalamic deacti- Terzano, M. G. (2009). Paradoxical insomnia: The role of
vation at sleep onset precedes that of the cerebral cortex in CAP and arousals in sleep misperception. Sleep Medicine,
humans. Proceedings of the National Academy of Sciences 10, 1139–1145.
USA, 107, 3829–3833. Pigarev, I. N., Nothdurft, H. C., & Kastner, S. (1997).
Mahowald, M. W., & Schenck, C. H. (2005). Insights from Evidence for asynchronous development of sleep in cortical
studying human sleep disorders. Nature, 437, 1279–1285. areas. Neuroreport, 8, 2557–2560.
Manconi, M., Ferri, R., Sagrada, C., Punjabi, N. M., Rector, D. M., Topchiy, I. A., Carter, K. M., & Rojas, M. J.
Tettamanzi, E., Zucconi, M., et al. (2010). Measuring the (2005). Local functional state differences between rat corti-
error in sleep estimation in normal subjects and in patients cal columns. Brain Research, 1047, 45–55.
with insomnia. Journal of Sleep Research, 19, 478–486. Rosenwasser, A. M. (2009). Functional neuroanatomy of sleep
Marzano, C., Ferrara, M., Moroni, F., & De Gennaro, L. and circadian rhythms. Brain Research Reviews, 61,
(2011). Electroencephalographic sleep inertia of the awak- 281–306.
ening brain. Neuroscience, 10, 308–317. Schenck, C. H., Milner, D. M., Hurwitz, T. D., Bundlie, S. R.,
Marzano, C., Ferrara, M., Sforza, E., & De Gennaro, L. & Mahowald, M. W. (1989). A polysomnographic and
(2008). Quantitative electroencephalogram (EEG) in insom- clinical report on sleep-related injury in 100, adult patients.
nia: A new window on pathophysiological mechanisms. The American Journal of Psychiatry, 146, 1166–1173.
Current Pharmaceutical Design, 14, 3446–3455. Steriade, M., McCormick, D. A., & Sejnowski, T. J. (1993).
Massimini, M., Huber, R., Ferrarelli, F., Hill, S., & Tononi, G. Thalamocortical oscillations in the sleeping and aroused
(2004). The sleep slow oscillation as a traveling wave. brain. Science, 262, 679–685.
Journal of Neuroscience, 24, 6862–6870. Stone, K. C., Taylor, D. J., McCrae, C. S., Kalsekar, A., &
McCormick, D. A., & Feeser, H. R. (1990). Functional Lichstein, K. L. (2008). Non restorative sleep. Sleep Medi-
implications of burst firing and single spike activity in lateral cine Reviews, 12, 275–288.
geniculate relay neurons. Neuroscience, 39, 103–113. Strijkstra, A. M., Beersma, D. G., Drayer, B., Halbesma, N.,
Moroni, F., Nobili, L., Curcio, G., De Carli, F., Fratello, F., & Daan, S. (2003). Subjective sleepiness correlates nega-
Marzano, C., et al. (2007). Sleep in the human hippocampus: tively with global alpha (8–12 Hz) and positively with cen-
A stereo-EEG study. PLoS One, 2, e867. tral frontal theta (4–8 Hz) frequencies in the human resting
Moroni, F., Nobili, L., Curcio, G., De Carli, F., Tempesta, D., awake electroencephalogram. Neuroscience Letters, 340,
Marzano, C., et al. (2008). Procedural learning and sleep hippo- 17–20.
campal low frequencies in humans. NeuroImage, 42, 911–918. Terzaghi, M., Sartori, I., Tassi, L., Didato, G., Rustioni, V., Lo
Moroni, F., Nobili, L., De Carli, F., Massimini, M., Russo, G., et al. (2009). Evidence of dissociated arousal
Francione, S., Marzano, C., et al. (2012). Slow EEG rhythms states during NREM parasomnia from an intracerebral
and inter-hemispheric synchronization across sleep and neurophysiological study. Sleep, 32, 409–412.
232

Terzano, M. G., Parrino, L., Spaggiari, M. C., Palomba, V., Wyatt, J. K., Bootzin, R. R., Allen, J. J., & Anthony, J. L. (1997).
Rossi, M., & Smerieri, A. (2003). CAP variables and Mesograde amnesia during the sleep onset transition: Replica-
arousals as sleep electroencephalogram markers for primary tion and electrophysiological correlates. Sleep, 20, 512–522.
insomnia. Clinical Neurophysiology, 114, 1715–1723. Wyatt, J. K., Bootzin, R. R., Anthony, J., & Bazant, S. (1994).
Tononi, G., & Cirelli, C. (2006). Sleep function and synaptic Sleep onset is associated with retrograde and anterograde
homeostasis. Sleep Medicine Reviews, 10, 49–62. amnesia. Sleep, 17, 502–511.
Vyazovskiy, V. V., Olcese, U., Hanlon, E. C., Nir, Y., Zadra, A., Pilon, M., Joncas, S., Rompré, S., & Montplaisir, J.
Cirelli, C., & Tononi, G. (2011). Local sleep in awake rats. (2004). Analysis of postarousal EEG activity during somnam-
Nature, 472, 443–447. bulistic episodes. Journal of Sleep Research, 13, 279–284.
Werth, E., Achermann, P., & Borbély, A. A. (1996). Brain Zadra, A., Pilon, M., & Montplaisir, J. (2008). Poly-
topography of the human sleep EEG: Antero-posterior somnographic diagnosis of sleepwalking: Effects of sleep
shifts of spectral power. Neuroreport, 8, 123–127. deprivation. Annals of Neurology, 63, 513–519.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 14

The circadian clock component PERIOD2: From


molecular to cerebral functions

Jürgen A. Ripperger* and Urs Albrecht

Department of Biology, Unit of Biochemistry, University of Fribourg, Fribourg, Switzerland

Abstract: The circadian clock is based on a molecular oscillator, which simulates the external day within
nearly all of a body’s cells. This “internalized” day then defines activity and rest phases for the cells and the
organism by generating precise rhythms in the metabolism, physiology, and behavior. In its perfect state,
this timing system allows for the synchronization of an organism to its environment and this may optimize
energy handling and responses to daily recurring challenges. However, nowadays, we believe that
desynchronization of an organism due to its lifestyle or problems with its circadian clock not only causes
discomfort but also may aggravate conditions such as depression, metabolic syndrome, addiction, or cancer.
In this review, we focus on one simple cogwheel of the mammalian circadian clock, the PERIOD2
(PER2) protein. Originally identified as an integral part of the molecular mechanism that yields overt
rhythms of about 24 h, more recently multiple other functions have been identified. In essence, the
PER proteins, in addition to their important function within the molecular oscillator, can be seen not
only as integrators on the input side of the circadian clock but also as mediators of clock output. This
diversity in their function is possible, because the PER proteins can interact with a multitude of other
proteins transferring oscillator timing information to the latter. In this fashion, the circadian clock
synchronizes many rhythmic processes.

Keywords: metabolism; mood; synchronization; Per2.

Introduction (Dibner et al., 2010; Ko and Takahashi, 2006; Liu


et al., 2007). However, under normal conditions
The circadian clock generates a projection of the we may not become aware of the existence of such
external day into the individual cells of the body a timing device in us. The reason for this is that the
circadian clock is responsive to external timing
*Corresponding author. cues, the so-called Zeitgebers (german for “time-
Tel.: þ41-26-300-8674 giver”), such as the light/dark cycle. Hence, the
E-mail: juergenalexandereduard.ripperger@unifr.ch effects of the circadian clock are generally masked

http://dx.doi.org/10.1016/B978-0-444-59427-3.00014-9 233
234

by external factors. To observe the characteristic However, interesting similarities and common con-
properties of a circadian clock, it is necessary to place cepts can be found over the borders of the taxa, such
an organism into conditions without external as the occurrence of posttranslational circadian
Zeitgebers (e.g., constant darkness or constant light oscillations of the redox state of peroxiredoxins,
of constant light intensity). In these free-running which are observed in human red blood cells
conditions, solely the circadian clock governs the (O’Neill and Reddy, 2011) and also the green algae
rhythmic metabolism, physiology, and behavior of Ostreococcus tauri (O’Neill et al., 2011). The mam-
an organism. The advantage of such a process is malian circadian oscillator is related to the one
obvious. It provides a temporal network allowing found in the fruit fly Drosophila melanogaster
synchronization of many metabolic pathways that (Dunlap, 1999). The basic concept is quite simple
otherwise would run each on their own pace, just and is found in essentially all cells of an organism.
driven by substrate availability, for example. Daily At the center of the clock is a negative transcrip-
separation of metabolic pathways yielding or con- tional feedback loop. A pair of transcriptional
suming energy is already observed in simple activators increases the transcription of a set of
organisms such as cyanobacteria (Dong et al., 2010; repressors. The corresponding repressor proteins
Johnson et al., 2011). Experiments performed with accumulate until their concentration is sufficient
cyanobacteria also provide compelling evidence for to shut down the transcription of their own genes.
the biological advantage of possessing such timing Consequently, transcription and translation of the
systems: the growth and survival rate of strains repressors cease and the activity of the repressor
with mutations in their circadian clock are generally proteins declines. As soon as the transcription block
not affected unless these mutants are kept in is relieved, a new cycle of about 24 h restarts, yield-
coculture with another strain that is in better reso- ing robust oscillations with a periodicity of about
nance with the environmental light/dark cycle a day. Obviously, such a simple mechanism would
(Ouyang et al., 1998). Under the latter conditions, not be sufficient to produce all of the robust rhythms
the better adjusted one completely wipes out the less observed in the organisms. There would be the
adjusted strain. Hence, there is a substantial risk that the oscillations generated by the feedback
biological advantage to possess a well functioning loop dampen rapidly, nearly reaching equilibrium
timing system, and it is therefore not surprising that conditions (Ripperger and Brown, 2010). Hence,
circadian clocks are found throughout the bacteria, there exist regulatory posttranslational mechanisms
plant, and animal kingdoms (Hut and Beersma, and an associated network of other feedback loops
2011; Yerushalmi and Green, 2009). However, a that stabilizes and reinforces the circadian rhythms.
recent field study with wild type and circadian This network also provides input information to the
rhythm-affected Period2 (Per2) knockout mice did core circadian oscillator to allow synchronization of
not reveal a strong selective disadvantage of having the self-sustaining loop to the environment. On the
an impaired function of the circadian clock in other side, most of the primary output generated
mammals (Daan et al., 2011). Here, we would like by the circadian oscillator is mediated primarily by
to elaborate the basic mechanisms of the mammalian rhythmic transcriptional regulation of target genes.
circadian clock and then focus on the functions of
the circadian clock component PER2 in liver metab-
olism and the brain. The history of Period

General mechanism of circadian clocks In the early 1970s, Konopka and Benzer performed
a mutagenesis screen in D. melanogaster to identify
Circadian clocks arose probably independently mutations in its circadian timing system (Konopka
during evolution in different taxa such as cyano- and Benzer, 1971). Surprisingly, the identified
bacteria, plants, and animals (Rosbash, 2009). mutations provoking shortening or lengthening of
235

the free-running period length or complete was verified genetically by analyzing the corre-
arrhythmicity in constant darkness mapped all to sponding knockout mouse strains (Bae et al., 2001;
the same gene, which was baptized Period (Per) Cermakian et al., 2001; Zheng et al., 1999, 2001).
(Konopka and Benzer, 1971; Reddy et al., 1984; At about the same time the mammalian transcrip-
Zehring et al., 1984). Nearly 20 years after the iden- tional activators, CLOCK (Gekakis et al., 1998;
tification of the Per mutants, it was realized that Per King et al., 1997) and BMAL1/MOP3 (Bunger
regulated the accumulation of its own mRNA and et al., 2000; Hogenesch et al., 1998), and a pair of
the concept of the feedback loop generating overt additional repressors, the CRYPTOCHROME
rhythms of about 24 h was born (Hardin et al., (CRY) 1 and 2 proteins were also found (Griffin
1990). However, it was not until about 10 years later et al., 1999; Kume et al., 1999; van der Horst et al.,
that the mammalian Per genes were identified and 1999). These proteins cooperate and make up the
cloned (Albrecht et al., 1997; Shearman et al., core loop of the molecular oscillator (Fig. 1). As
1997; Sun et al., 1997; Tei et al., 1997), and that mentioned above, all of the mammalian genes have
the function of these Per genes for the mammalian their counterparts in Drosophila, maybe with the
circadian oscillator (at least of Per1 and Per2) exceptions that the mammalian CRY proteins

Bmal1 Rev-Erbα

Clock

Per2

Fig. 1. The network of the mammalian molecular oscillator. The core loop (red shaded circle), responsible for generating rhythms
of about 24 h, is composed of activation of the Per genes by BMAL1 (red) and CLOCK (blue) and the increasing repression of Per
gene expression by the accumulation of its own gene product. The feedback inhibition by the PER proteins is delayed by
posttranslational modifications and interaction with the CRY proteins (not shown). In the stabilizing loop (blue circle), the Rev-
Erba gene is also activated by BMAL1 and CLOCK (not shown) and later on repressed by the PER proteins but immediately
REV-ERBa starts to inhibit transcription of the Bmal1 and Clock genes and of its own gene. PER2 can interact with REV-
ERBa (or PPARa, not shown) to regulate the Bmal1 gene. The overall organization of the network allows for a tight
synchronization of the core and stabilizing loops. Extended and adapted from Ko and Takahashi (2006).
236

are no (or less) important light-sensitive compo- maintain circadian oscillator function. Interestingly,
nents of the circadian oscillator (Griffin et al., the combination of any Per knockout with any Cry
1999; Okamura et al., 1999), and the function of knockout yielded a whole plethora of phenotypes
the Drosophila Timeless protein (which functions ranging from a rescue of the Per2’s arrhythmic phe-
similar as the mammalian CRY proteins) in the notype in Per2:Cry2 double knockout mice to
mammalian circadian clock is less pronounced completely arrhythmic Per2:Cry1 double knockout
(Gotter et al., 2000). mice (Oster et al., 2002, 2003a,b). These genetic
interactions suggest that the PER and CRY proteins
form multiple different complexes with each other
CRYs or PERs? to fulfill their function(s). Interestingly, all of
the mammalian clock components are redundant
Ever since the discovery of the main repressor (i.e., there exist Bmal1 and Bmal2, Clock and
molecules of the mammalian circadian oscillator, Npas2, Per1 and Per2, Cry1 and Cry2) (Ripperger
the question was posed, which of both kinds of et al., 2011). A reason for this is unknown yet. Are
molecules were the more important ones. The there two oscillators running in parallel in the same
genetics were not really helpful to unravel the cell or do the possible combinations mediate, for
precise functions of these supposed repressor example, tissue-specific gene regulation? At least
molecules. Mutations that affect the free-running for CLOCK and Npas2, it was established that
period of the circadian oscillator are thought to CLOCK may be more important in the periphery,
be integral parts of the clock mechanism or are and Npas2 may be more important in the brain
affecting its integral parts. Single knockout mice (DeBruyne et al., 2007a,b).
deficient for Per1 (Bae et al., 2001; Cermakian
et al., 2001; Zheng et al., 2001), Per2 (Bae et al.,
2001; Zheng et al., 1999), or Cry1 (van der Horst Function of PER proteins on the core loop of
et al., 1999; Vitaterna et al., 1999) had shorter the circadian oscillator
free-running circadian oscillators and mice defi-
cient for Cry2 (van der Horst et al., 1999; Considering the role of Per within the Drosophila
Vitaterna et al., 1999) had a longer free-running circadian oscillator (Hardin et al., 1990), it was
circadian oscillator, while the knockout for Per3 tempting to speculate that the mammalian PER
did not affect circadian rhythmicity (Bae et al., proteins fulfilled a similar function to create a tran-
2001). Interestingly, depending on the strain, scriptional feedback loop. However, cotransfection
Per2 single knockout mice had a shorter free- experiments in vitro revealed that, by contrast to
running period length but became arrhythmic the CRY proteins (Kume et al., 1999), the impact
after 1–2 weeks in constant darkness, suggesting of the PER proteins on BMAL1 and CLOCK-
that PER2 has a more prominent function mediated transcription was quite subtle (Kume
for the circadian clock than PER1. Hence, most et al., 1999). This was surprising because the bio-
of these individual components have an impact chemical analysis clearly indicated that PER and
on the clock mechanism. The combination of CRY proteins can be coimmunoprecipitated and
knockout mice for Per1 and Per2 (Zheng et al., consequently they are in the same (repressor?)
2001) and the combination of knockout mice for complex (Brown et al., 2005; Duong et al., 2011;
Cry1 and Cry2 (van der Horst et al., 1999) had Lee et al., 2001). Indeed, the regions within both
nearly the same phenotype, yielding arrhythmic classes of proteins mediating the interaction were
behavior immediately after placing the mice in biochemically mapped (Fig. 2) (Langmesser et al.,
constant conditions. Consequently, both classes 2008). Nevertheless, nonrhythmic overexpression
of proteins appear to be equally important to of PER2 but not CRY1 strongly interfered with
237

PAS A
PAS B
PAC Pro-rich Coiled-
coil
HLH

H3N COOH
1 1257
CLD NLS
NES1 NES 3 NES 2
CoRNR
LXXLL LXXLL

GSK3β β−TrCP CKI Cry, E4BP4

Fig. 2. Protein–protein interaction regions of PER2. The protein belongs to the class of PAS domain-containing factors. The PAS A
and B domains and the associated PAC domain mediate the interaction with other PER proteins and the BMAL1 and CLOCK
heterodimer. The region is flanked by binding sites for the regulatory kinase GSK3b on the left and the ubiquitin ligase b-TrCP
on the right. Both enzymatic activities affect the half-life of PER2. The basic helix-turn-helix (HLH) motif resembles a DNA-
binding motif but is probably not functional. The protein has three nuclear-export sites (NES), one nuclear-localization site
(NLS) and a cytoplasm-localization domain (CLD), which influence the nuclear localization of the protein. Interaction with the
CRY proteins (and E4BP4) is mediated by the C-terminus. The protein contains a sequence of casein kinase 1 (CKI)
phosphorylation sites organized as a relay, from which the priming site is affected in FASPS. PER2 also harbors some potential
motifs, which resemble binding sites for nuclear receptors (LXXLL and CoRNR), but only for the most N-terminal LXXLL
motif an interaction with nuclear receptors was identified. Finally, the protein contains potential interaction motifs resembling
proline-rich (Pro-rich) motifs and coiled-coil interaction motifs. Interactions with further kinds of proteins have been described,
such as with WDR5 (an adaptor for the histone methyl transferase Set1), NONO and SFPQ1, and histone demethylases of the
Jarid1 family; however, these were mainly identified in complexes containing also PER1 and a direct interaction has not been
demonstrated yet. Adapted from Albrecht et al. (2007).

normal circadian oscillator function, which is a et al., 2011), but also proteins that remove methyl
strong indicator for a prominent function of the groups from (DiTacchio et al., 2011) or add methyl
PER2 protein for the circadian oscillator (Chen groups to 50 -termini of histones (Brown et al.,
et al., 2009). The more detailed analysis suggested 2005; Etchegaray et al., 2006). However, it is
that the PER proteins form a kind of scaffold for necessary to decipher the precise biochemical
the CRY proteins to bind to their target sequences function(s) and their impact on the molecular
in the genome (Chen et al., 2009). Interestingly, oscillator of both classes of repressor molecules in
both kinds of proteins can interact with other more detail. Taken together, the available litera-
proteins that repress transcription via the modifica- ture on the function of PER proteins in vivo and
tion of the chromatin structure. For CRYs, these in vitro suggests that it is not the protein itself that
include histone deacetlyases (HDAC) (Naruse acts as a regulator within the circadian oscillator
et al., 2004), which attenuate local transcription mechanism, but it is the interaction of PER
by removing specific acetyl groups from 50 -termini proteins with other proteins, which mediates the
of histones, and for PER deacetylases (Duong effect(s).
238

Role of PER2 in the stabilizing loop to a specific amino acid motif in PER2, which res-
embles a typical nuclear receptor/coregulator
Initially, the characterization of the available cir- binding site (Fig. 2). However, the functional
cadian oscillator mutant and knockout mice extension of PER2 due to its interaction with
uncovered a surprise (Shearman et al., 2000). nuclear receptors appears not to be restricted to
The PER2 protein, but not PER1, was suggested the core circadian oscillator alone (see below).
as activator of circadian Bmal1 transcription. This
puzzle resolved itself about 2 years later, when an
additional feedback loop was discovered based on The rhythmic metabolism of the liver
the transcriptional repressor REV-ERBa (Pre-
itner et al., 2002; Ueda et al., 2002). Hence, the As revealed by DNA-microanalysis experiments,
model of the circadian transcriptional feedback many metabolic processes in the liver occur
loop was extended by the stabilizing loop (Fig. 1). rhythmically (Miller et al., 2007; Panda et al.,
In the core loop, BMAL1 and CLOCK activate 2002; Storch et al., 2002; Vollmers et al., 2009).
transcription of the Per genes until the PER At the center of this regulation is glucose homeo-
proteins (together with the CRY proteins) feed stasis to keep the blood glucose concentration at a
back onto their own synthesis. To achieve this, constant level (Lamia et al., 2008). Hence, over
there is a typical delay in the action of the repres- the day, the liver has to switch steadily from glu-
sor molecules as a prerequisite for the near 24 h cose uptake from the blood and the storage of
oscillations. In parallel, BMAL1 and CLOCK glucose as glycogen to the mobilization of glucose
also activate transcription of Rev-Erba. The from its glycogen stores and its secretion into
accumulating REV-ERBa subsequently represses the blood stream. About 15% of the transcripts
transcription of the Bmal1 and also the Clock found in the liver accumulate in rhythmic fashion
genes by replacing the transcriptional activator (Vollmers et al., 2009); however, only about 10%
RORa (Akashi and Takumi, 2005; Sato et al., of those rhythmic transcripts can be assigned to
2004) or PPARa (Canaple et al., 2006), depen- the circadian oscillator. This means that the bulk
dent on the cell type, and also by recruiting of rhythmic transcripts is induced solely by the
HDAC3 activity to the target genes (Feng et al., presence of their substrates. What is the advantage
2011). Because the action of REV-ERBa does of either way of regulation by induction through
not involve a delay mechanism typical for the substrates or by the circadian clock? Induction
action of the PER proteins, the phases of tran- through substrates necessitates the presence of
scription of the Per and Rev-Erba genes on substrates, but the response is normally fairly pro-
one hand and of the Bmal1 and Clock genes on portional to the substrate concentration. Hence, it
the other hand are separated by about 12 h. As is a quite economical way of gene regulation
conclusion, loss of repression of Bmal1 transcrip- according to the needs. On the other hand, the cir-
tion as observed in Per2 knockout mice was prob- cadian clock allows the expression of genes before
ably due to the reduction of REV-ERBa and a the substrates become available, obviously with
concomitant increase in Bmal1 expression. The the drawback that the gene is expressed even
story became even more complicated about if the substrates never come. However, the antici-
2 years ago. It was found that PER2 rather than pation of daily recurring events may be the main
PER1 directly interacted with both REV-ERBa biological advantage of possessing a circadian
and PPARa, contributing to about 10% of the clock as evidenced by the increase of fitness of cya-
repression by REV-ERBa and 20–30% of activa- nobacteria that are in better resonance with the
tion by PPARa of the Bmal1 gene in the liver environment. The circadian oscillator provides sta-
(Schmutz et al., 2010). This interaction was due ble rhythms of about 24 h, but over the course of a
239

year the photoperiod (i.e., the day/night ratio) superfamily (Yang et al., 2006). There are 49
changes with an impact on the time when the ani- members of this family, of which about one half
mal gets active and starts eating (Schultz and Kay, is regulated in a circadian fashion. In essence,
2003). Consequently, there exist mechanisms in rhythmic expression of the transcriptional regu-
the liver that uncouple anticipation mechanisms lators of the nuclear receptor family suffices to reg-
from the circadian oscillator to keep the anticipa- ulate their target genes in a rhythmic fashion. In
tion potential intact (Stratmann et al., 2010). As addition, many of these nuclear receptors require
example, some detoxification enzymes of the specific ligands for their activity. Some of these lig-
P450 family are rhythmically regulated by the cir- ands are produced and secreted in a rhythmic fash-
cadian transcription factor CAR, which itself is ion. Under normal conditions, this system could
regulated by the circadian transcription factors of stably regulate rhythmic processes but appears to
the PAR-Zip family (Gachon et al., 2006). The be circuitous and inflexible to respond rapidly to
reason may be that those detoxification enzymes, changes of the circadian clock. The extended
in the absence of their substrates, produce radical capacity of PER2 to directly interact with these
oxygen species, which may cause damage to the nuclear receptors provides an elegant way to trans-
cell. Hence, basal expression of these enzymes mit circadian clock information to the nuclear
is restricted to a phase when the occurrence of receptor regulated target genes (Fig. 3) (Schmutz
substrates is probable to provide some basic pro- et al., 2012). Nuclear receptors involved in the reg-
tection. In addition, the genes of these enzymes ulation of metabolic pathways and able to interact
can be super-induced by the nuclear receptor with PER2 in vitro include REV-ERBa, PPARa
CAR according to the presence of their substrates. and PPARg, and HNF4a and affect the lipid
Consequently, the risk of damage to the cell is and glucose metabolism, respectively. Interest-
optimally reduced. The expression of CAR is ingly, PER2 acts as nuclear receptor coregulator,
adjusted to the photoperiod by the PAR-Zip context-dependent as coactivator or corepressor.
factors, demonstrating that the precise temporal However, further experiments are required to
action of these regulators and their target genes fully understand these differences in its action.
relative to the activity phase of the animals is Nevertheless, PER2 transmits circadian oscillator
important (Stratmann et al., 2010). In conclusion, information directly to nuclear receptor-mediated
temporal organization of metabolic processes in metabolic pathways.
the liver optimizes not only energy expenditure
but also protects the organ from excessive damage.
The circadian clock represents an elegant means Function of PER2 in the brain
to fulfill such a coordinating function. Interest-
ingly, the analysis of Per knockout mice revealed Per2 expression was initially observed in the
a function of PER proteins in the concerted suprachiasmatic nuclei in the ventral part of the
action of DNA damage repair enzymes and hypothalamus and various other brain regions as
proto-oncogenes, rendering the Per knockout well as in peripheral tissues (Albrecht et al.,
animals more prone to tumor formation under 1997). Because of its inducibility by light, food
certain circumstances (Fu et al., 2002). restriction, and temperature pulses, Per2 was
recognized soon as a link between signals from
the environment (the input) and the clock mecha-
Role of PER2 in the liver metabolism nism. Per2 acts as a responder to environmental
signals and is a component of the core clock mech-
Many of the metabolic processes in the liver are anism, but it can also affect physiological pathways
regulated by members of the nuclear receptor downstream of the clock (the output, Fig. 3).
240

Input Clock Output

NR homo/hetero Target gene


dimers promoters

Nurr1 Dopaminergic system,


Light inflammation
PER2
Hnf4a Glucose metabolism
Food

TRa Temperature

PPARa Fatty acid metabolism

Rev-Erba Circadian timing

? Others

Fig. 3. Transmission of circadian clock information to nuclear receptor-target genes by PER2. PER2 mediates primary output from
the molecular oscillator. By direct interaction with the nuclear receptor (NR) homo- or heterodimers, it can affect the corresponding
target gene promoters and metabolic or physiological processes. The activity of PER2 may be modulated by the input (e.g., light or
food) to the circadian oscillator. Adapted from Schmutz et al. (2010).

Hence, it appeared that Per2 is an important activity, when animals were kept under a timed
player in determining the transcription status of feeding schedule (Feillet et al., 2006). Taken
the genome in a specific environmental context. together, it appears that the Per2 gene is at the
Therefore, it was not astonishing to find that crossroads of the neurobiological circuitry that is
Per2 plays a role in many brain related functions. common to feeding signals and drugs of abuse.
The reward system regulating addiction appeared In support of this view is the fact that Per2 is
to be at least partially affected when the Per2 expressed in regions of the midbrain such as the
gene was mutated (Abarca et al., 2002; Hampp arcuate nucleus and the ventral tegmental area,
et al., 2008; Spanagel et al., 2005). In particular, which regulate food uptake and reward pro-
it appeared that the rate-limiting step in dopa- cessing, respectively. Many of the metabolic pro-
mine degradation involving the monoamine cesses in the brain involve also nuclear receptors
oxidase A was under the control of the circadian such as Nurr1, which can physically interact with
clock mechanism including the Per2 gene PER2 (Ripperger et al., 2010; Schmutz et al.,
(Hampp et al., 2008). In line with this view is 2010), and hence, brain function is at least par-
the observation that PER2 variation in humans tially regulated by this avenue.
is associated with depression vulnerability
(Lavebratt et al., 2010; Partonen et al., 2007).
Recent findings indicate that the reward system Function of PER2 in sleep
and addiction processes share neurobiological
mechanisms with overeating and obesity (Simerly, Per2 appears to play a role in various parameters
2006). Interestingly, a mutation in the Per2 gene of sleep. Loss of Per2 affected proper sleep timing
of mice resulted in the loss of food anticipatory (Kopp et al., 2002), which is also evidenced in
241

advanced sleep-phase syndrome (Jones et al., phosphorylation of the PER2 protein reflects
1999). Per2 may affect parameters of sleep directly oscillator time and each phosphorylation
homeostasis as well (Franken et al., 2007), espe- site may represent a docking site for a specific
cially sleep deprivation and recovery sleep appear protein. Specific phosphorylation of a particular
to involve PER2. However, whether this function site is achieved by equilibrium of the action of
of PER2 is only attributable to its role in the kinases on one side and of phosphatases on the
brain is not known, because metabolic demands other side. For PER2, about 21 phosphorylation
in the liver involving this protein may also impact sites have been identified in fibroblasts although
on sleep mechanisms (Albrecht, 2011). Sleep the kinetics of phosphorylation of the individual
appears to be regulated by remodeling of neuro- sites remains to be established (Maier et al., 2009).
nal connectivity either strengthening or weaken- Some sites when phosphorylated by casein kinase
ing synaptic connections. The finding that PER2 1 (CK1) d or e serve as binding sites for b-TrCP1/2,
is involved in gating the light/dark information which add ubiquitination marks to the PER2 pro-
to vesicular glutamate transporter 1 (vGLUT1) tein, rendering it unstable and finally provoke deg-
content on synaptic vesicles (Yelamanchili et al., radation by the proteasome (Reischl et al., 2007;
2006) was of great interest, because the number Shirogane et al., 2005). An interesting case is the
of vGLUT1 molecules on the synaptic vesicle familial advanced sleep-phase syndrome (FASPS)
correlates with the glutamate filling state of this (Jones et al., 1999). Individuals with specific
vesicle and hence affects glutamate release poten- mutations in either CK1d (Xu et al., 2005) or the
tial, which may impinge on synaptic strength. CK1d binding site on PER2 display sleep rhythms
Hence, light may impact via PER2 on brain advanced by several hours (Toh et al., 2001) (a sim-
function and therefore, alterations in lighting ilar mutation in CK1e is provoking the familial
schedule as experienced in jet-lag and shift work delayed sleep-phase syndrome, (Takano et al.,
may affect behavior. 2004)). Detailed analysis revealed that the ratio of
kinase to substrate was important for the free-
running period length of the molecular oscillator
Posttranslational modifications of PER2 (Xu et al., 2007). On the phosphatase side, regu-
latory subunits of the protein phosphatase 1 were
How is the interaction potential of PER2 with shown to regulate the nuclear-localization PER2
other proteins regulated? One possibility is the and the free-running period length (Lee et al.,
accumulation of the PER2 protein over the circa- 2011; Schmutz et al., 2011). Are the interactions of
dian cycle (Lee et al., 2001). The peaks of tran- nuclear receptors affected by reversible phosphory-
scription and translation of the Per genes are lation? At the moment, there is no direct evidence
characteristically separated by a couple of hours. in favor of this hypothesis. The interaction motif
In addition, the accumulation of the PER proteins appears not to be adjacent to a known phosphoryla-
in the nucleus occurs with high amplitude every tion side. However, it could be that phosphor-
day and consequently, the interaction of PER ylation at a distant side affects the accessibility of
proteins with other proteins would follow their the interaction motif. As speculation, the region
affinity of interaction (i.e., first the high affinity between the two PAS domains is flexible to
binders and then the low affinity binders). dynamically shift from a conformation facilitating
Exchange of interaction partners would occur by interaction with BMAL1 and CLOCK to another,
simple competition mechanisms. This simple affin- which confers interaction with nuclear receptors.
ity model may be extended by posttranslational Phosphorylation mediating interaction with PER2
modifications (Vanselow and Kramer, 2010). In could also be present on the nuclear receptors, pos-
the extended model, the dynamic combination of sibly regulated by the presence of specific ligands.
242

Conclusions Albrecht, U. (2012). Circadian rhythms and sleep-the meta-


bolic connection. Pflügers Archiv European Journal of
Physiology, 463, 23–30.
Synchronization of the individual branches of the Albrecht, U., Bordon, A., Schmutz, I., & Ripperger, J. (2007).
circadian oscillator and the different routes of The multiple facets of Per2. Cold Spring Harbor Symposia
input and output might necessitate more effort on Quantitative Biology, 72, 95–104.
than simple transcriptional feedback loops can Albrecht, U., Sun, Z. S., Eichele, G., & Lee, C. C. (1997).
provide. Temporally controlled protein–protein A differential response of two putative mammalian circadian
regulators, mper1 and mper2, to light. Cell, 91, 1055–1064.
interactions enable the coupling of the high- Bae, K., Jin, X., Maywood, E. S., Hastings, M. H.,
precision timer of the circadian clock to, for Reppert, S. M., & Weaver, D. R. (2001). Differential
example, nuclear receptor-target genes. The high functions of mPer1, mPer2, and mPer3 in the SCN circadian
flexibility of these interactions allows PER2 to clock. Neuron, 30, 525–536.
act as coactivator or corepressor depending on Brown, S. A., Ripperger, J., Kadener, S., Fleury-Olela, F.,
Vilbois, F., Rosbash, M., et al. (2005). PERIOD1-associated
the regulatory context and hence to fine-tune proteins modulate the negative limb of the mammalian
the activity of these nuclear receptors. To fully circadian oscillator. Science, 308, 693–696.
understand the interaction potential of PER2 and Bunger, M. K., Wilsbacher, L. D., Moran, S. M.,
its regulation by posttranslational modifications, Clendenin, C., Radcliffe, L. A., Hogenesch, J. B., et al.
however, requires further experiments. These (2000). Mop3 is an essential component of the master circa-
dian pacemaker in mammals. Cell, 103, 1009–1017.
interactions may form the base of the many addi- Canaple, L., Rambaud, J., Dkhissi-Benyahya, O., Rayet, B.,
tional functions of PER2 in the metabolism and Tan, N. S., Michalik, L., et al. (2006). Reciprocal regulation
the brain. Other members of the core circadian of brain and muscle Arnt-like protein 1 and peroxisome
oscillator can interact with a variety of other proliferator-activated receptor alpha defines a novel positive
proteins and regulators, and it is tempting to spec- feedback loop in the rodent liver circadian clock. Molecular
Endocrinology, 20, 1715–1727.
ulate that the overall function(s) of these Cermakian, N., Monaco, L., Pando, M. P., Dierich, A., &
interactions are similar to those of PER2. How- Sassone-Corsi, P. (2001). Altered behavioral rhythms and
ever, it is also tempting to speculate that PER2 clock gene expression in mice with a targeted mutation in
can directly interact with many more classes of the Period1 gene. The EMBO Journal, 20, 3967–3974.
transcriptional regulators. Chen, R., Schirmer, A., Lee, Y., Lee, H., Kumar, V.,
Yoo, S. H., et al. (2009). Rhythmic PER abundance defines
a critical nodal point for negative feedback within the circa-
Acknowledgments dian clock mechanism. Molecular Cell, 36, 417–430.
Daan, S., Spoelstra, K., Albrecht, U., Schmutz, I., Daan, M.,
We thank for the support of our laboratory by the Daan, B., et al. (2011). Lab mice in the field: Unorthodox
daily activity and effects of a dysfunctional circadian clock
State of Fribourg and the Swiss National Science allele. Journal of Biological Rhythms, 26, 118–129.
Foundation. DeBruyne, J. P., Weaver, D. R., & Reppert, S. M. (2007a).
CLOCK and NPAS2 have overlapping roles in the
suprachiasmatic circadian clock. Nature Neuroscience, 10,
References 543–545.
DeBruyne, J. P., Weaver, D. R., & Reppert, S. M. (2007b).
Abarca, C., Albrecht, U., & Spanagel, R. (2002). Cocaine sen- Peripheral circadian oscillators require CLOCK. Current
sitization and reward are under the influence of circadian Biology, 17, R538–R539.
genes and rhythm. Proceedings of the National Academy of Dibner, C., Schibler, U., & Albrecht, U. (2010). The mamma-
Sciences of the United States of America, 99, 9026–9030. lian circadian timing system: Organization and coordination
Akashi, M., & Takumi, T. (2005). The orphan nuclear receptor of central and peripheral clocks. Annual Review of Physiol-
RORalpha regulates circadian transcription of the mamma- ogy, 72, 517–549.
lian core-clock Bmal1. Nature Structural & Molecular Biol- DiTacchio, L., Le, H. D., Vollmers, C., Hatori, M.,
ogy, 12, 441–448. Witcher, M., Secombe, J., et al. (2011). Histone lysine
243

demethylase JARID1a activates CLOCK-BMAL1 and hypoxia factors. Proceedings of the National Academy of
influences the circadian clock. Science, 333, 1881–1885. Sciences of the United States of America, 95, 5474–5479.
Dong, G., Kim, Y. I., & Golden, S. S. (2010). Simplicity and Hut, R. A., & Beersma, D. G. (2011). Evolution of time-
complexity in the cyanobacterial circadian clock mechanism. keeping mechanisms: Early emergence and adaptation to
Current Opinion in Genetics & Development, 20, 619–625. photoperiod. Philosophical Transactions of the Royal Soci-
Dunlap, J. C. (1999). Molecular bases for circadian clocks. ety of London: Series B, Biological Sciences, 366, 2141–2154.
Cell, 96, 271–290. Johnson, C. H., Stewart, P. L., & Egli, M. (2011). The
Duong, H. A., Robles, M. S., Knutti, D., & Weitz, C. J. (2011). cyanobacterial circadian system: From biophysics to
A molecular mechanism for circadian clock negative feed- bioevolution. Annual Review of Biophysics, 40, 143–167.
back. Science, 332, 1436–1439. Jones, C. R., Campbell, S. S., Zone, S. E., Cooper, F.,
Etchegaray, J. P., Yang, X., DeBruyne, J. P., Peters, A. H., DeSano, A., Murphy, P. J., et al. (1999). Familial advanced
Weaver, D. R., Jenuwein, T., et al. (2006). The polycomb group sleep-phase syndrome: A short-period circadian rhythm var-
protein EZH2 is required for mammalian circadian clock func- iant in humans. Nature Medicine, 5, 1062–1065.
tion. Journal of Biological Chemistry, 281, 21209–21215. King, D. P., Zhao, Y., Sangoram, A. M., Wilsbacher, L. D.,
Feillet, C. A., Ripperger, J. A., Magnone, M. C., Dulloo, A., Tanaka, M., Antoch, M. P., et al. (1997). Positional cloning
Albrecht, U., & Challet, E. (2006). Lack of food anticipation of the mouse circadian clock gene. Cell, 89, 641–653.
in Per2 mutant mice. Current Biology, 16, 2016–2022. Ko, C. H., & Takahashi, J. S. (2006). Molecular components of
Feng, D., Liu, T., Sun, Z., Bugge, A., Mullican, S. E., the mammalian circadian clock. Human Molecular Genetics,
Alenghat, T., et al. (2011). A circadian rhythm orchestrated 15(Spec No 2), R271–R277.
by histone deacetylase 3 controls hepatic lipid metabolism. Konopka, R. J., & Benzer, S. (1971). Clock mutants of Dro-
Science, 331, 1315–1319. sophila melanogaster. Proceedings of the National Academy
Franken, P., Thomason, R., Heller, H. C., & O’Hara, B. F. of Sciences of the United States of America, 68, 2112–2116.
(2007). A non-circadian role for clock-genes in sleep homeo- Kopp, C., Albrecht, U., Zheng, B., & Tobler, I. (2002). Homeo-
stasis: A strain comparison. BMC Neuroscience, 8, 87. static sleep regulation is preserved in mPer1 and mPer2 mutant
Fu, L., Pelicano, H., Liu, J., Huang, P., & Lee, C. (2002). The cir- mice. European Journal of Neuroscience, 16, 1099–1106.
cadian gene Period2 plays an important role in tumor sup- Kume, K., Zylka, M. J., Sriram, S., Shearman, L. P.,
pression and DNA damage response in vivo. Cell, 111, 41–50. Weaver, D. R., Jin, X., et al. (1999). mCRY1 and mCRY2
Gachon, F., Olela, F. F., Schaad, O., Descombes, P., & are essential components of the negative limb of the circa-
Schibler, U. (2006). The circadian PAR-domain basic leu- dian clock feedback loop. Cell, 98, 193–205.
cine zipper transcription factors DBP, TEF, and HLF mod- Lamia, K. A., Storch, K. F., & Weitz, C. J. (2008). Physiologi-
ulate basal and inducible xenobiotic detoxification. Cell cal significance of a peripheral tissue circadian clock.
Metabolism, 4, 25–36. Proceedings of the National Academy of Sciences of the
Gekakis, N., Staknis, D., Nguyen, H. B., Davis, F. C., United States of America, 105, 15172–15177.
Wilsbacher, L. D., King, D. P., et al. (1998). Role of the Langmesser, S., Tallone, T., Bordon, A., Rusconi, S., &
CLOCK protein in the mammalian circadian mechanism. Albrecht, U. (2008). Interaction of circadian clock proteins
Science, 280, 1564–1569. PER2 and CRY with BMAL1 and CLOCK. BMC Molecu-
Gotter, A. L., Manganaro, T., Weaver, D. R., lar Biology, 9, 41.
Kolakowski, L. F., Jr., Possidente, B., Sriram, S., et al. Lavebratt, C., Sjoholm, L. K., Partonen, T., Schalling, M., &
(2000). A time-less function for mouse timeless. Nature Neu- Forsell, Y. (2010). PER2 variantion is associated with
roscience, 3, 755–756. depression vulnerability. American Journal of Medical
Griffin, E. A., Jr., Staknis, D., & Weitz, C. J. (1999). Light- Genetics. Part B, Neuropsychiatric Genetics, 153B, 570–581.
independent role of CRY1 and CRY2 in the mammalian cir- Lee, H. M., Chen, R., Kim, H., Etchegaray, J. P.,
cadian clock. Science, 286, 768–771. Weaver, D. R., & Lee, C. (2011). The period of the circa-
Hampp, G., Ripperger, J. A., Houben, T., Schmutz, I., Blex, C., dian oscillator is primarily determined by the balance
Perreau-Lenz, S., et al. (2008). Regulation of monoamine oxi- between casein kinase 1 and protein phosphatase 1. Pro-
dase A by circadian-clock components implies clock influence ceedings of the National Academy of Sciences of the United
on mood. Current Biology, 18, 678–683. States of America, 108, 16451–16456.
Hardin, P. E., Hall, J. C., & Rosbash, M. (1990). Feedback of Lee, C., Etchegaray, J. P., Cagampang, F. R., Loudon, A. S., &
the Drosophila period gene product on circadian cycling of Reppert, S. M. (2001). Posttranslational mechanisms regu-
its messenger RNA levels. Nature, 343, 536–540. late the mammalian circadian clock. Cell, 107, 855–867.
Hogenesch, J. B., Gu, Y. Z., Jain, S., & Bradfield, C. A. Liu, A. C., Lewis, W. G., & Kay, S. A. (2007). Mammalian cir-
(1998). The basic-helix-loop-helix-PAS orphan MOP3 forms cadian signaling networks and therapeutic targets. Nature
transcriptionally active complexes with circadian and Chemical Biology, 3, 630–639.
244

Maier, B., Wendt, S., Vanselow, J. T., Wallach, T., Reischl, S., Reddy, P., Zehring, W. A., Wheeler, D. A., Pirrotta, V.,
Oehmke, S., et al. (2009). A large-scale functional RNAi Hadfield, C., Hall, J. C., et al. (1984). Molecular analysis
screen reveals a role for CK2 in the mammalian circadian of the period locus in Drosophila melanogaster and identifi-
clock. Genes & Development, 23, 708–718. cation of a transcript involved in biological rhythms. Cell, 38,
Miller, B. H., McDearmon, E. L., Panda, S., Hayes, K. R., 701–710.
Zhang, J., Andrews, J. L., et al. (2007). Circadian and Reischl, S., Vanselow, K., Westermark, P. O., Thierfelder, N.,
CLOCK-controlled regulation of the mouse transcriptome Maier, B., Herzel, H., et al. (2007). Beta-TrCP1-mediated
and cell proliferation. Proceedings of the National Academy degradation of PERIOD2 is essential for circadian dynamics.
of Sciences of the United States of America, 104, 3342–3347. Journal of Biological Rhythms, 22, 375–386.
Naruse, Y., Oh-hashi, K., Iijima, N., Naruse, M., Yoshioka, H., Ripperger, J. A., & Brown, S. A. (2010). Transcriptional regu-
& Tanaka, M. (2004). Circadian and light-induced transcrip- lation of circadian clocks. In U. Albrecht (Ed.), The circa-
tion of clock gene Per1 depends on histone acetylation and dian clock (pp. 37–78). New York: Springer.
deacetylation. Molecular and Cellular Biology, 24, 6278–6287. Ripperger, J. A., Jud, C., & Albrecht, U. (2011). The daily
Okamura, H., Miyake, S., Sumi, Y., Yamaguchi, S., Yasui, A., rhythm of mice. FEBS Letters, 585, 1384–1392.
Muijtjens, M., et al. (1999). Photic induction of mPer1 Ripperger, J. A., Schmutz, I., & Albrecht, U. (2010). PER-
and mPer2 in cry-deficient mice lacking a biological clock. suading nuclear receptors to dance the circadian rhythm.
Science, 286, 2531–2534. Cell Cycle, 9, 2515–2521.
O’Neill, J. S., & Reddy, A. B. (2011). Circadian clocks in Rosbash, M. (2009). The implications of multiple circadian
human red blood cells. Nature, 469, 498–503. clock origins. PLoS Biology, 7, e62.
O’Neill, J. S., van Ooijen, G., Dixon, L. E., Troein, C., Sato, T. K., Panda, S., Miraglia, L. J., Reyes, T. M.,
Corellou, F., Bouget, F. Y., et al. (2011). Circadian rhythms Rudic, R. D., McNamara, P., et al. (2004). A functional
persist without transcription in a eukaryote. Nature, 469, genomics strategy reveals Rora as a component of the mam-
554–558. malian circadian clock. Neuron, 43, 527–537.
Oster, H., Baeriswyl, S., Van Der Horst, G. T., & Albrecht, U. Schmutz, I., Albrecht, U., & Ripperger, J. A. (2012). The role
(2003a). Loss of circadian rhythmicity in aging mPer1/ of clock genes and rhythmicity in the liver. Molecular and
mCry2/ mutant mice. Genes & Development, 17, Cellular Endocrinology, 349, 38–44.
1366–1379. Schmutz, I., Ripperger, J. A., Baeriswyl-Aebischer, S., &
Oster, H., van der Horst, G. T., & Albrecht, U. (2003b). Daily Albrecht, U. (2010). The mammalian clock component
variation of clock output gene activation in behaviorally PERIOD2 coordinates circadian output by interaction with
arrhythmic mPer/mCry triple mutant mice. Chronobiology nuclear receptors. Genes & Development, 24, 345–357.
International, 20, 683–695. Schmutz, I., Wendt, S., Schnell, A., Kramer, A., Mansuy, I. M.,
Oster, H., Yasui, A., van der Horst, G. T., & Albrecht, U. & Albrecht, U. (2011). Protein phosphatase 1 (PP1) is a
(2002). Disruption of mCry2 restores circadian rhythmic- post-translational regulator of the mammalian circadian
ity in mPer2 mutant mice. Genes & Development, 16, clock. PLoS One, 6, e21325.
2633–2638. Schultz, T. F., & Kay, S. A. (2003). Circadian clocks in daily
Ouyang, Y., Andersson, C. R., Kondo, T., Golden, S. S., & and seasonal control of development. Science, 301, 326–328.
Johnson, C. H. (1998). Resonating circadian clocks enhance Shearman, L. P., Sriram, S., Weaver, D. R., Maywood, E. S.,
fitness in cyanobacteria. Proceedings of the National Acad- Chaves, I., Zheng, B., et al. (2000). Interacting molecular
emy of Sciences of the United States of America, 95, loops in the mammalian circadian clock. Science, 288,
8660–8664. 1013–1019.
Panda, S., Antoch, M. P., Miller, B. H., Su, A. I., Shearman, L. P., Zylka, M. J., Weaver, D. R.,
Schook, A. B., Straume, M., et al. (2002). Coordinated tran- Kolakowski, L. F., Jr., & Reppert, S. M. (1997). Two period
scription of key pathways in the mouse by the circadian homologs: Circadian expression and photic regulation in the
clock. Cell, 109, 307–320. suprachiasmatic nuclei. Neuron, 19, 1261–1269.
Partonen, T., Treutlein, J., Alpman, A., Frank, J., Shirogane, T., Jin, J., Ang, X. L., & Harper, J. W. (2005).
Johansson, C., Depner, M., et al. (2007). Three circadian SCFbeta-TRCP controls clock-dependent transcription via
clock genes Per2, Arntl, and Npas2 contribute to winter casein kinase 1-dependent degradation of the mammalian
depression. Annals of Medicine, 39, 229–238. period-1 (Per1) protein. Journal of Biological Chemistry,
Preitner, N., Damiola, F., Lopez-Molina, L., Zakany, J., 280, 26863–26872.
Duboule, D., Albrecht, U., et al. (2002). The orphan nuclear Simerly, R. (2006). Feeding signals and drugs meet in the mid-
receptor REV-ERBalpha controls circadian transcription brain. Nature Medicine, 12, 1244–1246.
within the positive limb of the mammalian circadian oscilla- Spanagel, R., Pendyala, G., Abarca, C., Zghoul, T., Sanchis-
tor. Cell, 110, 251–260. Segura, C., Magnone, M. C., et al. (2005). The clock gene
245

Per2 influences the glutamatergic system and modulates rhythmicity by cryptochromes 1 and 2. Proceedings of the
alcohol consumption. Nature Medicine, 11, 35–42. National Academy of Sciences of the United States of America,
Storch, K. F., Lipan, O., Leykin, I., Viswanathan, N., 96, 12114–12119.
Davis, F. C., Wong, W. H., et al. (2002). Extensive and Vollmers, C., Gill, S., DiTacchio, L., Pulivarthy, S. R.,
divergent circadian gene expression in liver and heart. Le, H. D., & Panda, S. (2009). Time of feeding and the
Nature, 417, 78–83. intrinsic circadian clock drive rhythms in hepatic gene
Stratmann, M., Stadler, F., Tamanini, F., van der Horst, G. T., & expression. Proceedings of the National Academy of Sciences
Ripperger, J. A. (2010). Flexible phase adjustment of circadian of the United States of America, 106, 21453–21458.
albumin D site-binding protein (DBP) gene expression by Xu, Y., Padiath, Q. S., Shapiro, R. E., Jones, C. R., Wu, S. C.,
CRYPTOCHROME1. Genes & Development, 24, 1317–1328. Saigoh, N., et al. (2005). Functional consequences of a
Sun, Z. S., Albrecht, U., Zhuchenko, O., Bailey, J., CKIdelta mutation causing familial advanced sleep phase
Eichele, G., & Lee, C. C. (1997). RIGUI, a putative mam- syndrome. Nature, 434, 640–644.
malian ortholog of the Drosophila period gene. Cell, 90, Xu, Y., Toh, K. L., Jones, C. R., Shin, J. Y., Fu, Y. H., &
1003–1011. Ptacek, L. J. (2007). Modeling of a human circadian muta-
Takano, A., Uchiyama, M., Kajimura, N., Mishima, K., Inoue, Y., tion yields insights into clock regulation by PER2. Cell,
Kamei, Y., et al. (2004). A missense variation in human casein 128, 59–70.
kinase I epsilon gene that induces functional alteration Yang, X., Downes, M., Yu, R. T., Bookout, A. L., He, W.,
and shows an inverse association with circadian rhythm sleep Straume, M., et al. (2006). Nuclear receptor expression links
disorders. Neuropsychopharmacology, 29, 1901–1909. the circadian clock to metabolism. Cell, 126, 801–810.
Tei, H., Okamura, H., Shigeyoshi, Y., Fukuhara, C., Yelamanchili, S. V., Pendyala, G., Brunk, I., Darna, M.,
Ozawa, R., Hirose, M., et al. (1997). Circadian oscillation Albrecht, U., & Ahnert-Hilger, G. (2006). Differential
of a mammalian homologue of the Drosophila period gene. sorting of the vesicular glutamate transporter 1 into a
Nature, 389, 512–516. defined vesicular pool is regulated by light signaling involv-
Toh, K. L., Jones, C. R., He, Y., Eide, E. J., Hinz, W. A., ing the clock gene Period2. Journal of Biological Chemistry,
Virshup, D. M., et al. (2001). An hPer2 phosphorylation 281, 15671–15679.
site mutation in familial advanced sleep phase syndrome. Yerushalmi, S., & Green, R. M. (2009). Evidence for the adap-
Science, 291, 1040–1043. tive significance of circadian rhythms. Ecology Letters, 12,
Ueda, H. R., Chen, W., Adachi, A., Wakamatsu, H., 970–981.
Hayashi, S., Takasugi, T., et al. (2002). A transcription fac- Zehring, W. A., Wheeler, D. A., Reddy, P., Konopka, R. J.,
tor response element for gene expression during circadian Kyriacou, C. P., Rosbash, M., et al. (1984). P-element trans-
night. Nature, 418, 534–539. formation with period locus DNA restores rhythmicity to
van der Horst, G. T., Muijtjens, M., Kobayashi, K., mutant, arrhythmic Drosophila melanogaster. Cell, 39,
Takano, R., Kanno, S., Takao, M., et al. (1999). Mammalian 369–376.
Cry1 and Cry2 are essential for maintenance of circadian Zheng, B., Albrecht, U., Kaasik, K., Sage, M., Lu, W.,
rhythms. Nature, 398, 627–630. Vaishnav, S., et al. (2001). Nonredundant roles of the mPer1
Vanselow, J. T., & Kramer, A. (2010). Posttranslational regu- and mPer2 genes in the mammalian circadian clock. Cell,
lation of circadian clocks. In U. Albrecht (Ed.), The circa- 105, 683–694.
dian clock (pp. 79–104). New York: Springer. Zheng, B., Larkin, D. W., Albrecht, U., Sun, Z. S., Sage, M.,
Vitaterna, M. H., Selby, C. P., Todo, T., Niwa, H., Eichele, G., et al. (1999). The mPer2 gene encodes a func-
Thompson, C., Fruechte, E. M., et al. (1999). Differential tional component of the mammalian circadian clock. Nature,
regulation of mammalian period genes and circadian 400, 169–173.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 15

Generation of mouse mutants as tools in


dissecting the molecular clock

Sneha N. Anand, Jessica K. Edwards and Patrick M. Nolan*

Neurobehavioural Genetics, MRC Harwell, Harwell Science and Innovation Campus, Oxfordshire,
United Kingdom

Abstract: Elucidation of the molecular basis of mammalian circadian rhythms has progressed
dramatically in recent years through the characterization of mouse mutants. With the implementation
of numerous mouse genetics programs, comprehensive sets of mutations in genes affecting circadian
output measures have been generated. Although incomplete, existing arrays of mutants have been
instrumental in our understanding of how the internal SCN clock interacts with the environment and
how it conveys its rhythm to remote oscillators. The use of ENU mutagenesis has proven to be a
significant contributor, generating mutations leading to subtle and distinct alterations in circadian
protein function. In parallel, progress with mouse gene targeting allows one to study gene function in
depth by ablating it entirely, in specific tissues at specific times, or by targeting specific functional
domains. This has culminated in worldwide efforts to target every gene in the mouse genome allowing
researchers to study multiple gene targeting effects systematically.

Keywords: ENU; mutagenesis; mouse; transgenics; gene-driven screen.

Introduction might play a role in the progression of complex


diseases and behaviors have now been identified.
A major challenge for scientific research in the Questions such as how these genes act and
twenty-first century will be to assign functions what pathways they are involved in can be
to ca. 25,000 genes of the mammalian genome. answered by determining the consequences of
With innovations made in high-throughput geno- gene disruption using mutants (Picciotto and
mics techniques, a large number of genes that Wickman, 1998). The systematic generation of
mouse mutants through the use of two comple-
mentary approaches, forward and reverse genet-
*Corresponding author.
Tel.: þ44-01235-841091; Fax: þ44-01235-842000 ics, has been key in advancing our knowledge
E-mail: p.nolan@har.mrc.ac.uk of how genes function in biological systems.

http://dx.doi.org/10.1016/B978-0-444-59427-3.00015-0 247
248

Forward genetics begins with the investigation of been established from the analysis of mutant
a given phenotype and makes use of genetics phenotypes in mice and other rodents. This strat-
and genomics tools to identify the causative gene. egy continues to provide functional insight into
Conversely, reverse genetics uses our knowledge the molecular basis of circadian rhythm genera-
of a candidate gene or gene sequence, its expres- tion and maintenance.
sion patterns and presumed functions, and then
introduces mutations in the gene using available
The mouse as a model organism
technologies to study phenotype.
These two complementary approaches have been
The house mouse, Mus musculus, has long been
equally invaluable in contributing to our under-
used as a vital tool for molecular genetics. There
standing of how the molecular clock ticks. In this
are currently more than 450 inbred mouse strains
chapter, we provide a brief overview of the
available (Beck et al., 2000), the accessibility of
approaches and technologies currently used to gen-
which has allowed for finely detailed genetic
erate mouse clock mutants and review the progress
maps to be generated. The mouse is an ideal
in behavioral and molecular analyses using these
model, given its high homology with the human
tools. Finally, we consider new developments and
genome. The structure of the mouse genome is
directions in this area that promise to provide signif-
similar to humans, with approximately the same
icant mouse mutant resources for the future.
number of genes that display extensive synteny
(Boguski, 2002). The mouse also displays measur-
able endophenotypes of emotional states, such as
Forward genetics
anxiety, depression, and fear, which allow for
complex behavioral conditions and traits to be
The benefit of forward genetics approaches as a
analyzed. The genetic composition and behav-
complement to gene targeting strategies is that it
ioral characteristics of the mouse have ensured
is unbiased, requiring no a priori assumptions
that it is at the frontier of mammalian molecular
about gene function. Importantly, it is hypothesis
genetics research across a range of disciplines.
generating, meaning that novel genes may be
attributed to known pathologies and behaviors.
Most behaviors are complex traits, relying upon Random mutagenesis in mice
genetic and environmental interactions, with most
genetic components currently unknown. This is Prior to the era of gene targeting, our knowledge of
true for the circadian system. Insights into the mouse genes and genomes was established through
molecular basis of circadian function have been the analysis of spontaneous or induced mutations.
possible primarily through the use of forward Spontaneous mutations occur at a low rate of
genetics screens and the analysis of mutant approximately 5  10 6 per locus (Silver, 1995).
phenotypes in model organisms such as Drosoph- Nevertheless, many spontaneous mutants with vis-
ila (Konopka and Benzer, 1971) and Arabidopsis ible or easily detectable phenotypes were used to
(Ahmad and Cashmore, 1993). Mammalian circa- establish genetic maps in the mouse and these, in
dian function has depended on the analysis of turn, were used as a backbone to establish the
mutations in orthologous genes in Drosophila order of numerous molecular markers on each
and Arabidopsis, the period and cryptochrome chromosome, progress that eventually led to the
genes (Bae et al., 2001; Cermakian et al., 2001; sequencing of the entire mouse genome. Given
Shearman et al., 2000; van der Horst et al., 1999; the rarity of their occurrence, it is unusual for spon-
Zheng et al., 1999, 2001). Nevertheless, several taneous mutants to be identified solely on the basis
important insights into circadian function have of circadian phenotypes. Exceptions to this can, of
249

course, be found in other mammalian species. (Cordes, 2005). The resultant mutations are
Cloning of the spontaneous hamster mutant, tau diverse, including missense or nonsense, loss or
(Ralph and Menaker, 1988), has been crucial for gain of STOP codons, or those which affect splic-
the investigation of posttranslational mechanisms ing or transcriptional activity. In this way, ENU
in clock function (Lowrey et al., 2000). Spontane- can create an allelic series to finely dissect gene
ous mutants affecting other components of circa- function. ENU may therefore be considered a
dian regulatory pathway have been identified in complementary approach to transgenic strategies,
mice through phenotypes associated with other since issues such as redundancy or pleiotropy,
gene functions, for example, the staggerer mutant may be avoided if the induced mutation still
carries a mutation in the orphan nuclear receptor enables gene function.
Rora (Sidman et al., 1962). Despite the historical
role of spontaneous mutations, their utility in the
future of functional genetic analysis is limited due ENU mutagenesis screening strategies
to the large colonies required.
Several mutagenic agents have been used in ENU screens are generally used to identify domi-
mouse genetics to increase the rate of identifying nant or recessive traits. Screens for dominant
functional mutations, including those that affect traits are easiest to carry out as progeny of
behavior. X-ray mutagenesis was one of the first ENU-mutagenized males (G0) crossed to wild-
mutagens used, increasing mutation frequency by type females are directly phenotyped (Fig. 1).
20–100 times (Cordes, 2005). The disadvantage Each product of this cross (G1 founder) can
of X-ray mutagenesis is that it commonly induces potentially carry mutations at up to 100 different
chromosomal rearrangements which inevitably loci. Each G1 individual can be phenotyped for
affect multiple genes or may result in positional dominant mutations either using a high-through-
effects. Chemical mutagenesis emerged as an effi- put strategy or for specific phenotypes using a
cient alternative although many chemicals, includ- specialized phenotyping pipeline (Nolan et al.,
ing chlorambucil, induced similarly confounding 2000a). G1 founders carrying abnormal traits are
contiguous gene effects (Rinchik et al., 1990). Of then bred to assess inheritance of the trait. This
all the chemical mutagens available for mamma- approach is quick, since only one generation of
lian studies, ENU (N-ethyl-N-nitrosourea) is breeding is required before phenodeviants are
unique in that it almost exclusively induces point identified. Dominant mutations, however, are
mutations at an extremely high efficiency. It acts recovered on average 10 times less frequently
through direct alkylation of nucleic acids, showing than recessive mutations. A further disadvantage
a slight bias toward alkylating A-T bases, at a rate to this screening approach lies in the fact that only
of 1 mutation every 1–1.5 Mb (Quwailid et al., mutations of moderate to large effect size are lia-
2004). ENU has a high affinity for spermatogonial ble to be detected. This can be especially con-
stem cells allowing for mutations to be inherited founding if a particular trait is subject to high
through the germline. Crucially, it does not cause variability. An alternative method, though much
complete sterility in males; rather, after adminis- less high throughput, is to screen for recessive
tration, it results in a period of temporary sterility mutations. The recessive screen requires two
as spermatogonia are depleted from, and then more generations of breeding to generate a
repopulate, the testis with individual mature screening population. Here, the G1 progeny are
sperm carrying unique catalogs of mutations in again mated to a wild-type animal to produce
their genome (Nolan et al., 2000b). G2 offspring. The ENU mutations will still be in
The most common point mutations are trans- a heterozygous state, so these offspring are inter-
versions (A-T! T-A) or transitions (A-T ! G-C) crossed to produce a G3 pedigree, with 25% now
250

ENU

X
G0
+/+ m/+

G3
Backcross recessive
screen

G1
X
Dominant screen +/+ m/+
+/+

X
m/+ m/m
X
G2

m/+ +/+ +/+ m/+

G3
Intercross
recessive screen m/m m/+ +/+

Fig. 1. ENU mutagenesis dominant and recessive breeding scheme. Dominant mutations may be observed in G1 progeny, whereas
recessive mutations may be observed in G3 progeny using either a backcross or intercross strategy.

homozygous for any of the ENU-induced activity that are subject to high variability.
mutations. Alternatively, G2 female progeny Another approach is to target a specific molecular
may be backcrossed to the male founder (G1) pathway by screening for mutations that alter
and again, G3 pedigrees containing homozygous reporter gene expression on a specific sensitized
animals will be produced (Balling, 2001). Gener- background (Stottmann et al., 2011).
ally, 20–40 animals per pedigree are screened
ensuring that 5–10 animals will carry the homozy-
gous mutant phenotype. Screening ENU phenodeviants for circadian
Several modifications of either approach can be phenotypes
used to target specific phenotypic domains or
molecular pathways. One modification is to screen Circadian phenodeviants can be identified in a
for dominant mutations in a two-generation screen based upon wheel-running activity under
screen. Using this strategy, mutations with smaller an automated lighting regime (Bacon et al.,
effect size can be readily screened (Kumar et al., 2004). Each animal, from either G1 or G3 litters,
2011), for example, in screens such as open field is singly housed in a cage fitted with a running
251

wheel. Each revolution of the wheel is recorded mapping. In the case of Afh, 13 of the most severe
and plotted as a measure of the animal’s activity phenodeviants for period length in constant dark-
throughout the entire screen to produce a ness were selected. This region of linkage is
double-plotted actogram. A typical primary known as a “candidate region” and sequencing
screen consists of an initial 1-week period under of the region is required to identify the functional
a 12:12 light–dark schedule to entrain the mutation. Technological advances now make it
animals. The animals are then exposed to a possible for the whole genome to be sequenced
period of constant darkness for up to 5 weeks. following a genome-wide scan, such that all
The screen may then continue with a period of sequences within a critical region can be aligned
constant light, under which a wild-type animal and the functional mutation identified. This is
will normally lengthen its period. Other lighting becoming the most cost-effective and time-efficient
schedules may include light pulses to assess mas- method by which to identify mutations.
king or phase shifting, or returning to a 12:12 Once a mutation has been identified, it is
light–dark cycle to assess reentrainment after important to prove that it is causative of the phe-
exposure to constant light conditions. As well notype, through techniques such as bacterial arti-
as period length, levels of activity, phase angle, ficial chromosome (BAC) rescue (Antoch et al.,
and responses to light intensity may also be ana- 1997; Justice et al., 1999). Further phenotypic
lyzed to further dissect the circadian phenotype analyses may be required to fully characterize
on a behavioral level. Typical examples of the the mutant, and often further crosses to other
diversity of mutants identified in wheel-running mutants or inbred lines may be useful in deter-
screens are shown in Fig. 2. mining whether there are genetic interactions
with other known and characterized genes. For
circadian research, forward genetic screens and
Mapping and positional cloning of the utilization of techniques such as ENU muta-
ENU mutations genesis has so far been essential in expanding
our knowledge of the molecular basis of the circa-
Following identification of phenodeviants, dian oscillator and the vital role of the circadian
mutations are tested for inheritance and pene- clock on gross behavior and physiology.
trance. Phenodeviants are mated to wild-type
animals and the progeny are tested for the
observed phenotype. Those mutations which Classes of mutants identified in mouse
are inherited and maintain high levels of pene- ENU screens
trance are then selected for mapping and posi-
tional cloning of the mutation. The principles of Mutations that uncover new clock elements
genetic mapping and positional cloning have been
followed in identifying the after-hours (Afh) circa- The use of similar forward genetics screens in Dro-
dian mutant (Godinho et al., 2007). This mutant sophila and Arabidopsis has informed our knowl-
was identified in an ENU mutagenesis dominant edge of how the mammalian circadian oscillator
screen and was subsequently intercrossed to pro- functions, uncovering core elements such as
duce homozygous animals displaying a period period, timeless, and cryptochrome that play
length of approximately 27 h in constant darkness. important roles in multiple species. Although this
The first step in identifying a causative muta- approach is certainly more laborious in mammals,
tion is to generate a high resolution map of the it has led to key discoveries that have, in turn,
mutant locus and a region of linkage identified. informed our knowledge of circadian function in
Typically, 10–20 phenodeviants are selected for other species. The identification of the Clock gene
252

(a) (b)

(c) (d)

(e) (f)

(g) (h)
253

in a mouse ENU mutagenesis program was the cat- timed degradation of core mammalian circadian
alyst for the elucidation of the molecular basis of proteins. Both mutations are missense mutations
mammalian circadian function and the establish- resulting in amino acid substitutions at the
ment of functional conservation across multiple C-terminal end of the protein. Although neither
species (Antoch et al., 1997; King et al., 1997b; mutation affects Fbxl3 stability or the ability of
Vitaterna et al., 1994). The identification of this Fbxl3 to associate with the ubiquitin ligase com-
core element through the forward genetics plex, both affect the ability of Fbxl3 to associate
approach was particularly relevant as the mutation with mammalian cryptochromes, reducing their
was shown to be antimorphic (King et al., 1997a). efficiency in targeting them for degradation. The
The Clock mutation, an A ! T transversion, was ability of any cell to rapidly clear protein levels
identified in the 50 splice donor site of intron 19 over circadian time or even in response to changes
resulting in skipping of exon 19 and a 51-amino in environmental condition appears to be highly
acid deletion in the mutant protein. Subsequently, functionally conserved across a number of species
generation of a Clock/ mutant produced and would suggest that multiple additional com-
surprising results (Debruyne et al., 2006). The pre- ponents in this regulatory pathway remain to be
diction was that the circadian oscillator in Clock/ identified and characterized.
mice would be severely compromised and circa-
dian rhythmicity would be lost. Unexpectedly,
circadian rhythmicity was maintained in the homo- Additional mutations in known clock genes
zygous null mutants. Only a mild phenotype was
observed—a shortening of the period length in The idea that one mutant is sufficient to study gene
constant darkness by approximately 20 min and function is redundant. Even in Drosophila, where
an advanced onset of activity of 2 h. In retrospect, there is a single period gene, multiple allelic
it would appear that the significance of Clock gene mutations with varying effects on phenotype have
function might not have been identified if one were been established resulting in period lengthening,
to use reverse genetics alone, and this early exam- period shortening, or complete arrhythmicity.
ple highlights the complementarity of both Flybase (http://flybase.org/) lists over 10 Per
approaches in understanding gene function. It alleles identified in chemical mutagenesis screens
would now appear that the paralog Npas2 may that result in defective circadian rhythms. Each
compensate for the loss of Clock gene function. of these mutants can add to our knowledge of
More recently, the discovery of multiple muta- how a particular protein functions by mapping
tions within the Fbxl3 locus in mice (Godinho functional domains onto each protein sequence
et al., 2007; Siepka et al., 2007) has highlighted (protein interaction, DNA binding, enzyme re-
the importance of the proteasomal system in the cognition, etc.). The same strategy applies to

Fig. 2. Selected mutants identified in an ENU mutagenesis screen. Representative images illustrating wheel-running activity rhythm
records for mutant lines identified in dominant and recessive ENU screens. Each image is a double-plotted actogram of wheel-
running activity under light/dark conditions (LD), followed by constant dark conditions (DD) and, in some cases, constant light
conditions (LL). Occasionally, animals also received light pulses during the LD phase of the screen. Shaded blocks represent
times at which the lights are on and short vertical bars represent periods of wheel-running activity. Records of littermate activity
patterns for (a) a wild-type mouse; (b) a mutant line with a long period in DD and arrhythmicity in LL; (c) a mutant line free-
running in LD with an extremely long period in DD and activity suppression in LL; (d) a mutant line with reduced entrainment,
low amplitude short period rhythms in DD, and low amplitude rhythms in LL; (e) a mutant line with reduced period lengthening
in LL; (f) a mutant line with extreme period lengthening in LL; (g) a wild-type mouse; and (h) mutant line with extreme short
period in DD.
254

mammalian ENU mutagenesis screens although, rhythms in response to changing environmental


with exceptions, the identification of multiple alle- cues. Although this leads to an efficient biological
lic mutations is far more infrequent. In one process, it has consequences when trying to inves-
instance, a hypomorphic mutation in Cry1, named tigate the function of any single gene by knocking
part-time (prtm), has recently been described it out. Such effects on output rhythms are evident
(Siepka et al., 2007) indicating that ENU screens when comparing the Csnk1etau/tau mutant in mice
in mice have the potential to identify other func- to the Csnk1e/ mutant; whereas the null allele
tionally relevant mutations in clock genes apart shows only a slight deficit in period, introduction
from nulls. of the tau point mutation results in a large and sig-
nificant shortening of period (Meng et al., 2008).
Due to the complexity of the circadian system,
Mutations in genes that may affect clock circuitry the presence of multimolecular interactions for
proper functioning of the clock is essential. It is
By measuring locomotor output rhythms, ENU evident that analyzing a specific locus indepen-
mutagenesis screens also have the capacity to dently does not provide enough knowledge to dis-
uncover mutations that affect clock circuitry. sect the function and interaction of particular
The earlybird mutation was identified in a modest genes in such a complex system. Hence, it is neces-
screen of 500 animals (Kapfhamer et al., 2002). sary to consider the epistatic interactions that
Although the dominant screen identified a mild contribute to the variability of the circadian system.
short period phenotype, homozygotes display a For example, Maywood et al. showed that
circadian period as short as 22 h as well as sleep pat- Csnk1etau and Fbxl3Afh showed additive and inde-
tern disturbances. Sequence analysis identified a pendent effects on regulation of the circadian feed-
mutation in the synaptic vesicle trafficking protein, back loop (Maywood et al., 2011). As illustrated,
Rab3a. The nature of the mutation would suggest the use of ENU mutagenesis has proved and will
that a deficit in synaptic transmission could be continue to prove its power to identify new alleles
responsible for the short period phenotype. In which will allow us to better understand the
another dominant screen, a seizure susceptibility mechanisms of the circadian clock.
mutant was identified, where heterozygotes showed
additional secondary phenotypes including length-
ened free-running period with associated sleep dis- Reverse genetics
turbances (Kirshenbaum et al., 2011). The mouse
carries a missense mutation in the Naþ,Kþ-ATPase Gene knockout through gene targeting
a3 isoform, resulting in a I810N substitution.
The most common way to study gene function
in vivo is to ablate its function (gene knockout).
Investigating redundancy and epistasis in This approach makes use of homologous recombi-
ENU mutants nation in embryonic stem (ES) cells to introduce a
nonfunctional gene into the host genome. Initially,
The robustness of the circadian system is evident mammalian genes were targeted on the basis of
in that many of the components involved in the phenotypes identified in other organisms such as
regulation of the system exist as paralog pairs, Drosophila. However, as more genomic data
often serving partially overlapping functions. became available, genes were knocked out on the
One reason for this could be to ensure that basis of their oscillating expression patterns, gene
deficits in any single gene need not dramatically or protein interactions, or in vitro gene expression
affect the ability of an organism to drive or adjust or knockdown data. Undoubtedly, this approach
255

has led to many important advances in our under- phenotype in Cry2/ mice, their functions could
standing of the molecular basis of rhythm genera- be interpreted as being nonredundant. However,
tion and maintenance. Nevertheless, limitations of Cry1/; Cry2/ double knockout mice are
this approach in the characterization of circadian arrhythmic in constant darkness suggesting that
mutants can often be attributed to functional the two paralogs may have partial overlapping
redundancy as discussed earlier. The implication functions (van der Horst et al., 1999). The double
here is that no one gene is necessary to maintain knockout mutant phenotype of paralogous positive
circadian rhythms in mammals. In fact, the only regulators Clock and Npas2 was also shown to
single gene knockout to result in arrhythmicity, be arrhythmic in constant darkness, while each sin-
Bmal1 (Bunger et al., 2000), does so by virtue of gle mutant had negligible effects on clock function
the fact that Bmal1 regulates the expression of its (DeBruyne et al., 2007). Although the basic helix-
paralog, Bmal2 (Shi et al., 2010). loop-helix genes Dec1 and Dec2 (Bhlhe40 and
Bhlhe41) were proposed as negative regulators
of the clock, neither null mutant had significant
Interactions between paralogous genes effects on clock function. Again, double mutants
exhibited a significant lengthening of the circadian
This approach identifies where functional red- period (Rossner et al., 2008). Additional molecular
undancies in paralogs or gene families might be analysis of these mutants provided further evi-
present. A systematic investigation would be dence that may well act as transcriptional
highly informative, as it has been in species coactivators as well as repressors depending on
such as Drosophila melanogaster, Caenorhabditis how they interact with the Clock/Bmal complex.
elegans, and Arabidopsis thaliana; nevertheless, Mutant interaction studies may also help in
the time, cost, and space involved in carrying out defining roles for paralogous genes outside their
these studies in mice means that only limited circadian function. All three PAR bZIP transcrip-
focused work has been carried out. As all three tion factors (D site albumin promoter binding
mammalian homologues of the Drosophila Per protein, DBP; hepatic leukemia factor, HLF;
gene were cloned and knocked out, an initial focus and thyrotroph embryonic factor, TEF) oscillate
was to determine whether they had distinct or in brain and liver. Although Dbp/ animals were
overlapping functions in period regulation. All shown to have a small but significant shortening of
three individual Per mutants display short free- free-running period (Lopez-Molina et al., 1997),
running period phenotypes, Per2/ exhibiting no further circadian phenotype was observed when
the strongest phenotype and Per1/ and Per3/ all three genes were knocked out (Gachon et al.,
progressively weaker (Cermakian et al., 2001; 2004). However, their role in other aspects of
Shearman et al., 2000; Zheng et al., 1999, 2001). CNS function was identified as triple mutants
The generation of double mutants led to the con- showed an increased susceptibility to spontaneous
clusion that Per1 and Per2 had overlapping roles and audiogenic seizures.
as the mice became completely arrhythmic when
released into constant darkness. Per3, on the other
hand, seemed to have a distinct and less vital role Interactions between nonparalogous genes
in rhythm generation as double mutants with it
and either Per1 or Per2 showed no additional phe- Investigation of compound mutant phenotypes can
notype (Bae et al., 2001). Similar conclusions were confirm how two unrelated genes may be inter-
drawn following the generation of double Cry1/; acting in a circadian pathway. These investigations
Cry2/ mutants. With a shorter circadian period can produce surprising results that question our
observed in Cry1/ mice and a longer circadian prior understanding of how molecular rhythms
256

are orchestrated. Oster et al. refined the roles of is driven by a promoter that will express the
period and cryptochrome by comparing double recombinase in a specific spatiotemporal pattern.
mutant phenotypes (Oster et al., 2002, 2003). As technologies are advancing, conditional circa-
Unexpectedly, they found that Per1/; Cry1/ dian mutants have become available for use, includ-
mice had free-running periods similar to wild-type, ing those for Bmal1 (Marcheva et al., 2010), Clock
while single gene mutants had short free-running (Debruyne et al., 2006), Csnk1d and Csnk1e
periods. A similar paradoxical result showed that (Etchegaray et al., 2009), and Sirt1 (Nakahata
Per1/ further lengthened Cry2/ period. Con- et al., 2008). As we discuss later, the numbers and
versely, Per2/ mice appeared to maintain their availability of conditional or “conditional-ready”
phenotype in a Cry/ background. Per2/; resources and mutants should start to increase and
Cry1/ double mutants lose rhythmicity immedi- become readily available through worldwide efforts
ately upon release into free-running conditions, such as EUCOMM (the European conditional
whereas Per2/; Cry2/ mutants maintain a mouse mutagenesis program) and KOMP (the
rhythm in constant darkness that is slightly shorter knockout mouse project).
than wild-type. This study would support the belief The power of the conditional mutagenesis
that Per2 and Cry1 are the strongest negative reg- approach will develop with the emergence of
ulators of clock function. Evidence for molecular new Cre driver lines. Reports in the literature to
interactions between Per2 and Rev-erba has also date have predominantly been on conditional
been identified. Support for this was elucidated ablation of gene expression in liver or pancreas
through generation of double mutants, where the for two reasons. First, the consequences of clock
double mutant phenotype is significantly shorter gene dysfunction in these organs can help in dis-
than either single mutant (Schmutz et al., 2010). secting the role of these genes in metabolic
As well as assigning a genetic interaction for these homeostasis. The Sirt1/ and Csnk1d/ mutants
two proteins, the authors provide evidence that are both homozygous lethal, whereas conditional
these two proteins can act as transcriptional ablation in the liver shows that they are both
corepressors. effective in disrupting molecular oscillations. On
the other hand, Bmal1/ is not lethal but condi-
tional ablation in pancreatic islets results in defec-
Inducible and tissue-specific knockouts tive beta-cell function. The second reason for
these lines of study is that mouse lines that
It is becoming increasingly evident that many express Cre in these tissues or cells are more read-
components involved in regulation of clock func- ily available. Up to now no mouse lines driving
tion can serve multiple purposes from develop- Cre expression exclusively in the suprachiasmatic
ment and differentiation through to metabolic nucleus (SCN) have been available. One of the lat-
homeostasis and cell cycle regulation in adults. est mouse lines that may be useful has been
Although gene knockouts have provided essential reported recently (Husse et al., 2011). Although
insights into clock gene function, they can also not exclusively expressed in SCN, this
result in confounding pleiotropic phenotypes. To Synaptotagmin-10 Cre line is highly expressed in
overcome these effects, the spatiotemporal this region with limited expression in other brain
expression of mutant alleles can be controlled regions and no expression in peripheral organs
through the generation of inducible and tissue- apart from testis. Using this line, conditional dele-
specific knockouts. By this method, the use of tion of Bmal1 results in arrhythmicity.
two engineered mutants is required, one where New Cre driver lines currently being generated
loxP sites flank a critical target exon in the gene may be archived in public resources (Table 1).
of interest and a second where Cre recombinase One such resource is the GENSAT transgenic
257

Table 1. Information on and access to useful mouse resources

Resource Acronym Web details Description

International Mouse Phenotyping IMPC http://www. Information on and portal to worldwide mouse
Consortium mousephenotype. genetics programs and infrastructures
org/
European Mouse Mutant EMMA http://www.emmanet. Searchable database of mouse mutant strains;
Archive org/ available as cryopreserved material or live mice
Mutant Mouse Regional MMRRC http://www.mmrrc. Database of mouse mutant strains, KOMP lines,
Resource Centers Archive org/ Cre driver lines, ES cell lines, etc.
Europhenome Mouse Europhenome http://www. Searchable database of mouse strains and mutant
Phenotyping Resource europhenome.org/ lines phenotyped in primary phenotyping pipelines;
databrowser/viewer. searchable by gene or phenotype
jsp
International Knockout Mouse IKMC http://www. Searchable database of mouse mutant strains;
Consortium knockoutmouse.org/ status on mutant generation for any gene provided
European Mouse Phenotyping EMPReSS http://empress.har. Database of standard operating procedures for
Resource of Standardized mrc.ac.uk/ screens in primary phenotyping pipelines
Screens
Gene Expression Nervous GENSAT http://www.gensat. Database of neuroscience-relevant mouse
System Atlas org/index.html transgenic BAC-EGFP and BAC-Cre recombinase
driver lines
Coordination of Resources for CREATE http://creline.org/ Unified portal to access worldwide Cre line
Conditional Expression of home resources
Mutated Mouse Alleles
RIKEN ENU-Based Gene- RGDMS http://www.brc.riken. Archive of ENU mutants identified in reverse
Driven Mutagenesis System jp/lab/mutants/ genetics screens; details of how to request
genedriven.htm additional screens
MouseBook MouseBook http://www. Archive of mouse mutants archived at Harwell
mousebook.org/ including ENU lines; searchable by phenotype or
index.php gene

BAC-Cre recombinase driver line resource Transgenics


(Gong et al., 2007), providing access to and in-
formation on over 100 Cre driver lines expressed Expression of transgenes is quite commonly used in
in region-specific patterns in mouse brain (see circadian research to help identify specific functions
http://www.gensat.org/cre.jsp). This includes of genes. One way in which to do this is by pronu-
many driver lines of interest to the circadian com- clear injection of a BAC expressing the gene of
munity including Avp, Vip, Vipr2, Prokr2, and interest and containing all or most of the elements
Dbp lines. Other resources and searchable required for the expression of the native protein.
databases include CREATE, a coordination of This approach was used to generate a Bmal1 trans-
resources for conditional expression of mutated genic as one of a series of mouse genetic tools to
mouse alleles (http://creline.org/home). This site distinguish the effects of Bmal1 dysfunction in brain
acts as a portal to access Cre line resources and and muscle (McDearmon et al., 2006). Transgenic
searchable databases worldwide. The diverse constructs can also be generated and genetically
expression patterns of these lines will no doubt modified to constitutively overexpress a gene and/
enable biologists to dissect circadian function in or to introduce specific mutations or tags. These
a systematic way. modifications have been used to complement
258

mouse knockout research and can often identify (McDearmon et al., 2006). In this study, Bmal1
functionally relevant domains or uncover novel was constitutively expressed in either brain using
roles for clock proteins. In two cases, transgenic the secretogranin-2 promoter or in muscle using
mice have been generated to model the human the human a-actin-1 promoter in Bmal1/ mice.
inherited condition, familial advanced sleep phase While the brain-rescued mice displayed circadian
syndrome (FASPS). Transgenic mice expressing rhythmicity, the muscle-rescued mice displayed no
the Ser662Gly mutation in Per2 had an advanced circadian rescue but activity and body weight
phase of activity in a light–dark schedule and a phenotypes were alleviated. Using the same tTA
short free-running period of activity (Xu et al., mouse line, transgenic overexpression of the wild-
2007). A second missense mutation in Csnk1d is type Clock gene in brain shortened circadian
also associated with FASPS (Xu et al., 2005), the period, whereas it was lengthened by over-
mutation resulting in reduced enzymatic activity expressing the dominant negative transgene. Both
in vitro. Again, the phenotype in a mouse trans- effects could be reversed by treating animals with
genic line mimicked that of the human subjects. doxycycline (Hong et al., 2007). In a similar study,
Consequences of Cry1 overexpression were Kornmann et al. used a hepatocyte-specific pro-
investigated using a Cry1 cDNA construct driven moter driving tTA to constitutively express a Rev-
by the chick b-actin promoter and a CMV erba transgene in the liver, thus effectively turning
enhancer (Okano et al., 2009). Although a 5- to off the liver clock while retaining a functional
10-fold ubiquitous overexpression of Cry1 had SCN clock (Kornmann et al., 2007). They con-
no evident circadian phenotype, the introduction cluded that some molecular oscillations in the liver
of a Cys414Ala substitution resulted in a circa- could still be driven by systemic factors coming
dian period of ca. 28 h, indicating that this domain from the SCN clock.
of Cry1 has a crucial rhythm-related function. In Targeted gene knock-ins can be useful in deter-
defining the role of the Bmal1 paralog, Bmal2, mining whether mutations or polymorphisms in
in biological rhythms, a Bmal2 transgenic con- humans or in other species are functional and
struct constitutively driven by the human EF-1a can specifically be used to investigate the
promoter was generated (Shi et al., 2010). By functions of particular protein domains or amino
crossing these mice to Bmal1/ mice, behavioral acids. Rather than using pronuclear injection to
wheel-running and metabolic rhythmicity could achieve random integration of the DNA of inter-
be restored. This finding was crucial in defining est, this technique uses an approach similar to the
an otherwise ambiguous role for Bmal2 in circa- knockout approach where DNA is integrated spe-
dian function and suggested that Bmal1 and cifically at the target locus via homologous recom-
Bmal2 may function as a paralogous pair. bination. In this way, a mutant allele can replace
Implementing temporal control over the expres- the wild-type allele. This approach circumvents
sion of transgenes can further dissect gene function positional effects (site of integration effects) and
(Picciotto and Wickman, 1998). Tetracycline- gene dosage effects. The technique has been used
controlled transcriptional activation is widely used rarely in the circadian field but has been used to
for this purpose and makes use of a specific mouse introduce the hamster tau mutation (Arg178Cys)
mutant line where the tetracycline transactivator into mice (Meng et al., 2008).
(tTA) is controlled by a promoter directing tissue-
specific expression and the expression of a target Reverse genetics using ENU
gene is under the control of the tetracycline
operator. This system has been used in a number Using ENU mutagenesis in a reverse genetics
of circadian studies, for example, examining strategy is a powerful tool, complementary to for-
the phenotypes associated with tissue-specific ward genetic and transgenic approaches. The high
transgenic rescue of the Bmal1/ phenotype efficiency mutation induction rate of ENU enables
259

the potential detection of multiple mutations phylogenetic analysis can be carried out to
within a single gene (Quwailid et al., 2004). As determine whether the mutation lies in an evolu-
ENU-induced mutations are predominantly mis- tionarily conserved motif. Moreover, the ability
sense mutations, diverse mutations can be uncov- to perform functional in vitro assays prior to
ered at a single locus that might differentially generating mutants is desirable. This type of
affect protein–protein interactions, protein–DNA in vitro prescreen is ideally suited to investigating
interactions, and amino acids that are likely to be mutations that might affect the circadian system.
posttranslationally modified (e.g., phosphorylation, In vitro assays in well-established cell lines are
acetylation, and ubiquitiation). widely used to determine phenotypes and pri-
A number of research centers using ENU mary effects of mutations in circadian biology.
mutagenesis have been developing archives of Cell lines such as U2OS, NIH3T3 (Akashi and
frozen material from progeny of mutagenized Nishida, 2000), and Rat-1 fibroblasts (Balsalobre
animals or from mutagenized ES cells. In the et al., 1998; Izumo et al., 2003) are known to
Harwell ENU archive (http://www.har.mrc.ac.uk/ express clock genes. The protein products of
services/enu-dna-archive), parallel DNA and these genes are also accumulated in a circadian
sperm samples have been archived from over fashion and model events in vivo.
10,000 animals, each animal potentially carrying Real-time monitoring assays, making use of
one nonsynonymous or nonsense mutation clock gene promoters such as Per2, Bmal1, Rev-
per  2 Mb of DNA. Mutation detection within erba, driving the expression of luciferase
the DNA archive can be carried out rapidly using reporters have been of great value in determining
a number of methods including high-throughput the effects of a gene on circadian function (Baggs
sequencing (Fig. 3). Many technologies have been et al., 2009). This approach can be modified to
developed for high-throughput detection of focus on particular target molecules. For example,
mutations including heteroduplex analysis and Maier et al. carried out a large-scale RNAi screen
denaturing high-performance liquid chromatogra- by knocking down all the known and predicted
phy (Quwailid et al., 2004). To date, users of the kinases, phosphatases, and F-box proteins indi-
Harwell ENU archive have screened coding vidually and studied the effect on molecular
regions of interest and more than 250 mutations oscillations (Maier et al., 2009). This same
have been found in over 80 genes, with a proven approach can be used to determine whether
multiple hit rate in all genes tested. Mouse lines overexpression of mutant proteins identified in
can subsequently be resurrected from the parallel ENU screens can affect oscillator function.
frozen sperm archive and mutant phenotypes The utility of the ENU gene-driven approach is
characterized. Many of the mutations identified evident in the mouse behavioral field as more
in the archive have been rederived in this way, novel mutant phenotypes are being identified,
resulting in  70 new mouse mutant lines. To including those where mouse knockout phenotypic
date, specific genes have been screened using analysis has failed to reveal convincing data.
this approach. With the advent of next-generation Mutants in Disrupted-in-Schizophrenia-1 (Disc1)
sequencing, it is becoming increasingly realistic to and Serine racemase (Srr) are recently identified
carry out whole exome sequencing on the entire examples known to be associated with behavioral
ENU archive uncovering thousands of new and and psychiatric disorders in humans. Q31L and
unique mouse mutations (Gondo et al., 2010). L100P are two missense mutations identified in
Identification of ENU mutations in this Disc1, the human homolog of which is a genetic
way posits the challenge of determining whether risk factor for schizophrenia (Clapcote et al.,
the mutation identified will be functional. Conse- 2007). Q31L displayed a recessive major depres-
quently, it is advisable to conduct in silico analysis sion phenotype, while the L100P proved to be a
before deriving a particular line. For example, dominant schizophrenia mouse model. Srr is also
260

Candidate gene selection

Exon specific amplification


by PCR

Wild-type female ENU treated male


(G0) (G0)

Heteroduplex analysis for


mutation detection

Wild-type Mutant Heteroduplexes Homoduplexes

Male G1 founder heat


cool
+ A C G T
A T G C A T G C

Sperm DNA
archive archive Mutation confirmation by DNA
sequencing

ENU mutant rederivation from


archived frozen sperm

Fig. 3. ENU gene-driven screen. Parallel DNA and sperm samples are archived from G1 founder males. Any gene can be targeted
using this approach. The DNA archive is amplified using primers flanking exons of candidate genes. The light scanner is one of a
number of high-throughput apparatus used to identify heteroduplexes in pools of amplified DNA. The presence of a mutation is
confirmed in individual DNA samples and by sequencing. After in silico analysis to confirm that the mutation is likely to be
functional, the mutant line is rederived from the parallel frozen sperm sample.

considered as a genetic risk factor for schizophre- are proving extremely fruitful in terms of identifi-
nia. ENU archive screening lead to the identifica- cation and annotation of novel genes, much work
tion of a nonsense mutation in Srr. Mice carrying remains in annotating the whole genome. It is clear
this mutation were found to display a knockout- that manipulation of the genome, through muta-
like phenotype with altered Srr mRNA levels genesis, gene targeting, transgenics, or gene silenc-
(50% reduction) and no SRR protein (Labrie ing in vitro, has become an effective strategy for
et al., 2009). An independent study that generated determining gene function, but now efforts are
Srr/ mice using the gene targeting approach also focused on scaling these approaches up through
reported similar phenotypes (Basu et al., 2009). sharing resources and working systematically on
functional characterization of the genome, gene
Large-scale mouse mutant production and by gene. Over the past few years, collaborative
phenotyping pipelines efforts have been initiated that aim to broadly
annotate each gene in the mouse genome, through
Although forward and reverse genetic screens car- generating and phenotyping mouse gene
ried out by individual institutes and researchers knockouts, and share this information with the
261

wider research community. This is being achieved home cage analysis. Data for each of the several
through projects such as EUCOMM, KOMP, hundred lines produced under this program is
EUMODIC, the European Mouse Disease Clinic; now collated in the Europhenome database
and IMPC, the International Mouse Phenotyping (http://www.europhenome.org/databrowser/
Consortium (Table 1). viewer.jsp) (Mallon et al., 2008; Morgan et al.,
2010) and is searchable by phenotype or by gene
name. The database currently includes information
EUCOMM, KOMP, and IKMC on circadian-relevant genes such as Nr1d1 (REV-
ERBa), Nr1d2 (REV-ERBb), and Hdac3. Nr1d2
These programs were set up to generate large mice, for example, show significant deficits in
numbers of mutations by gene targeting in mouse respiratory, blood biochemistry, and immune
ES cells (Friedel et al., 2007; Ringwald et al., function parameters.
2011; Skarnes et al., 2011). Currently, many of
these mutations are being generated such that
they can be rendered into conditional alleles, IMPC
allowing Cre recombinase-mediated disruption
of gene function in a time- and tissue-specific As an extension of EUMODIC (in terms of scale
manner. The cell lines and mutants are available and number of partners involved), the IMPC was
as “knockout first” and “conditional-ready” mice. formed with the aim of characterizing mouse
The programs are now organized in such a way mutants produced by the International Mouse
that ES and mutant line generation is been shared Knock-Out Consortium, IMKC. The consortium
among the worldwide mouse genetics community. consists of large mouse genetics centers, including
institutes based in Europe, Australia, Japan,
USA, and Canada, which have the capacity
EUMODIC for large-scale, high-throughput phenotyping as
well the ability to generate “conditional-ready”
EUMODIC was originally established to com- knockout mutants from targeted ES cells. This
plement the EUCOMM program. The purpose approach will allow researchers to complement
of the program was to generate mutant lines their own reverse or forward genetic research by
from ES cells and phenotype  500 of these. In having access to general multisystem phenotypic
phenotyping, the program uses a selection of care- data for individual gene knockouts. Mouse
fully selected high-throughput tests (EMPReSS, mutants themselves will be available through a
http://empress.har.mrc.ac.uk/) that act as a primer number of searchable mutant archives including
for further characterization by interested inves- the European Mouse Mutant Archive (EMMA,
tigators. Age-matched cohorts of 10 males and 10 http://www.emmanet.org/) and the Mutant Mouse
females are subjected to phenotyping pipelines Regional Resource Centers Archive (MMRRC,
that consist of morphological, metabolic, cardio- http://www.mmrrc.org/).
vascular, and bone assessment as well as neuro- IMPC, like forward genetics, uses an unbiased
behavioral, sensory, hematological, and immune approach to understanding gene function where
system characterization. Phenodeviants may be no assumptions or hypotheses are made
selected by individual researchers for more specific concerning phenotypes expected. In this way, it
phenotypic assessment. For example, animals dis- is likely that many new associations between gene
playing particular neurobehavioral phenotypes and biological function will be identified. As pre-
may then be selected for circadian and sleep viously discussed, relying on knockout studies
phenotyping through wheel-running screens and may not always reveal all true functions of a gene
262

because of confounds such as redundancy and Acknowledgments


homozygous lethality. This approach will, how-
ever, provide researchers with complementary This work has been supported by the MRC and
information regarding candidate genes and may by the 6th Framework Project EUCLOCK
also highlight critical function of novel genes in (No. 018741).
a particular biological domain. These programs
will also allow the researcher to consider spatio- References
temporal aspects of gene function by conditional
ablation. Among important considerations for the Ahmad, M., & Cashmore, A. R. (1993). HY4 gene of A.
future in the identification and characterization of thaliana encodes a protein with characteristics of a blue-light
mutants will be the continued development of high photoreceptor. Nature, 366, 162–166.
Akashi, M., & Nishida, E. (2000). Involvement of the MAP
throughput, automated data-capture technologies
kinase cascade in resetting of the mammalian circadian
to phenotype even complex behaviors, such as cir- clock. Genes & Development, 14, 645–649.
cadian rhythm disturbances. Antoch, M. P., Song, E. J., Chang, A. M., Vitaterna, M. H.,
The aim and subsequent success of the past Zhao, Y., Wilsbacher, L. D., et al. (1997). Functional identi-
decade in genetic research has been the comple- fication of the mouse circadian Clock gene by transgenic
BAC rescue. Cell, 89, 655–667.
tion of genome sequencing of a number of differ-
Bacon, Y., Ooi, A., Kerr, S., Shaw-Andrews, L., Winchester, L.,
ent model organisms and the development of Breeds, S., et al. (2004). Screening for novel ENU-induced
efficient methods for generating mutants. Future rhythm, entrainment and activity mutants. Genes, Brain,
aims lie in the domain of functional genomic and Behavior, 3, 196–205.
research, to annotate and assign function to each Bae, K., Jin, X., Maywood, E. S., Hastings, M. H.,
Reppert, S. M., & Weaver, D. R. (2001). Differential
gene through the phenotypic analysis of mouse
functions of mPer1, mPer2, and mPer3 in the SCN circadian
mutants. Large-scale projects such as IMPC will clock. Neuron, 30, 525–536.
propel this research forward, with the aim that Baggs, J. E., Price, T. S., Ditacchio, L., Panda, S.,
this should be completed within the next decade. Fitzgerald, G. A., & Hogenesch, J. B. (2009). Network features
Future challenges will lie in developing an under- of the mammalian circadian clock. PLoS Biology, 7, e52.
Balling, R. (2001). ENU mutagenesis: Analyzing gene function
standing of molecular pathways and interactions
in mice. Annual Review of Genomics and Human Genetics,
of individual components within systems. It will 2, 463–492.
be important to consider strategies to model Balsalobre, A., Damiola, F., & Schibler, U. (1998). A serum
multigenic traits, complex conditions, epistasis, shock induces circadian gene expression in mammalian tis-
quantitative trait loci, and gene–environment sue culture cells. Cell, 93, 929–937.
Basu, A. C., Tsai, G. E., Ma, C. L., Ehmsen, J. T.,
interactions. This remains a challenge for devel-
Mustafa, A. K., Han, L., et al. (2009). Targeted disruption
oping our understanding of circadian systems as of serine racemase affects glutamatergic neurotransmission
circadian phenotypes are frequently associated and behavior. Molecular Psychiatry, 14, 719–727.
with other disorders and systems such as metabo- Beck, J. A., Lloyd, S., Hafezparast, M., Lennon-Pierce, M.,
lism and diabetes (Froy, 2011), the cell cycle and Eppig, J. T., Festing, M. F., et al. (2000). Genealogies of
cancers (Rosbash and Takahashi, 2002), and mouse inbred strains. Nature Genetics, 24, 23–25.
Boguski, M. S. (2002). Comparative genomics: The mouse that
behavior and psychiatric disease (Wulff et al., roared. Nature, 420, 515–516.
2010). In order to rise to this challenge, it will be Bunger, M. K., Wilsbacher, L. D., Moran, S. M.,
necessary to combine forward and reverse genetic Clendenin, C., Radcliffe, L. A., Hogenesch, J. B., et al.
approaches together with in vitro approaches. (2000). Mop3 is an essential component of the master circa-
Our ability to target multiple components of a par- dian pacemaker in mammals. Cell, 103, 1009–1017.
Cermakian, N., Monaco, L., Pando, M. P., Dierich, A., &
ticular system will pave the way to understanding Sassone-Corsi, P. (2001). Altered behavioral rhythms and
how genes interact to direct mammalian biological clock gene expression in mice with a targeted mutation in
processes. the Period1 gene. The EMBO Journal, 20, 3967–3974.
263

Clapcote, S. J., Lipina, T. V., Millar, J. K., Mackie, S., real-time luminescence reporting: Temperature compensa-
Christie, S., Ogawa, F., et al. (2007). Behavioral phenotypes tion and damping. Proceedings of the National Academy of
of Disc1 missense mutations in mice. Neuron, 54, 387–402. Sciences of the United States of America, 100, 16089–16094.
Cordes, S. P. (2005). N-ethyl-N-nitrosourea mutagenesis: Justice, M. J., Noveroske, J. K., Weber, J. S., Zheng, B., &
Boarding the mouse mutant express. Microbiology and Bradley, A. (1999). Mouse ENU mutagenesis. Human
Molecular Biology Reviews, 69, 426–439. Molecular Genetics, 8, 1955–1963.
Debruyne, J. P., Noton, E., Lambert, C. M., Maywood, E. S., Kapfhamer, D., Valladares, O., Sun, Y., Nolan, P. M.,
Weaver, D. R., & Reppert, S. M. (2006). A clock shock: Rux, J. J., Arnold, S. E., et al. (2002). Mutations in Rab3a
Mouse CLOCK is not required for circadian oscillator func- alter circadian period and homeostatic response to sleep loss
tion. Neuron, 50, 465–477. in the mouse. Nature Genetics, 32, 290–295.
Debruyne, J. P., Weaver, D. R., & Reppert, S. M. (2007). King, D. P., Vitaterna, M. H., Chang, A. M., Dove, W. F.,
CLOCK and NPAS2 have overlapping roles in the Pinto, L. H., Turek, F. W., et al. (1997). The mouse Clock
suprachiasmatic circadian clock. Nature Neuroscience, 10, mutation behaves as an antimorph and maps within the
543–545. W19H deletion, distal of Kit. Genetics, 146, 1049–1060.
Etchegaray, J. P., Machida, K. K., Noton, E., King, D. P., Zhao, Y., Sangoram, A. M., Wilsbacher, L. D.,
Constance, C. M., Dallmann, R., Di Napoli, M. N., et al. Tanaka, M., Antoch, M. P., et al. (1997). Positional cloning
(2009). Casein kinase 1 delta regulates the pace of the mam- of the mouse circadian clock gene. Cell, 89, 641–653.
malian circadian clock. Molecular and Cellular Biology, 29, Kirshenbaum, G. S., Clapcote, S. J., Duffy, S., Burgess, C. R.,
3853–3866. Petersen, J., Jarowek, K. J., et al. (2011). Mania-like behav-
Friedel, R. H., Seisenberger, C., Kaloff, C., & Wurst, W. ior induced by genetic dysfunction of the neuron-specific
(2007). EUCOMM–the European conditional mouse muta- Naþ, KþATPase {alpha}3 sodium pump. Proceedings of
genesis program. Briefings in Functional Genomics & Prote- the National Academy of Sciences of the United States of
omics, 6, 180–185. America, 108, 18144–18149.
Froy, O. (2011). The circadian clock and metabolism. Clinical Konopka, R. J., & Benzer, S. (1971). Clock mutants of Dro-
Science (London, England), 120, 65–72. sophila melanogaster. Proceedings of the National Academy
Gachon, F., Fonjallaz, P., Damiola, F., Gos, P., Kodama, T., of Sciences of the United States of America, 68, 2112–2116.
Zakany, J., et al. (2004). The loss of circadian PAR bZip Kornmann, B., Schaad, O., Bujard, H., Takahashi, J. S., &
transcription factors results in epilepsy. Genes & Develop- Schibler, U. (2007). System-driven and oscillator-dependent
ment, 18, 1397–1412. circadian transcription in mice with a conditionally active
Godinho, S. I., Maywood, E. S., Shaw, L., Tucci, V., liver clock. PLoS Biology, 5, e34.
Barnard, A. R., Busino, L., et al. (2007). The after-hours Kumar, V., Kim, K., Joseph, C., Thomas, L. C., Hong, H., &
mutant reveals a role for Fbxl3 in determining mammalian Takahashi, J. S. (2011). Second-generation high-throughput
circadian period. Science, 316, 897–900. forward genetic screen in mice to isolate subtle behavioral
Gondo, Y., Fukumura, R., Murata, T., & Makino, S. (2010). mutants. Proceedings of the National Academy of Sciences
ENU-based gene-driven mutagenesis in the mouse: A next- of the United States of America, 108(Suppl. 3), 15557–15564.
generation gene-targeting system. Experimental Animals, 59, Labrie, V., Fukumura, R., Rastogi, A., Fick, L. J., Wang, W.,
537–548. Boutros, P. C., et al. (2009). Serine racemase is associated
Gong, S., Doughty, M., Harbaugh, C. R., Cummins, A., with schizophrenia susceptibility in humans and in a mouse
Hatten, M. E., Heintz, N., et al. (2007). Targeting Cre model. Human Molecular Genetics, 18, 3227–3243.
recombinase to specific neuron populations with bacterial Lopez-Molina, L., Conquet, F., Dubois-Dauphin, M., &
artificial chromosome constructs. The Journal of Neurosci- Schibler, U. (1997). The DBP gene is expressed according
ence, 27, 9817–9823. to a circadian rhythm in the suprachiasmatic nucleus and
Hong, H. K., Chong, J. L., Song, W., Song, E. J., influences circadian behavior. The EMBO Journal, 16,
Jyawook, A. A., Schook, A. C., et al. (2007). Inducible 6762–6771.
and reversible Clock gene expression in brain using the Lowrey, P. L., Shimomura, K., Antoch, M. P., Yamazaki, S.,
tTA system for the study of circadian behavior. PLoS Zemenides, P. D., Ralph, M. R., et al. (2000). Posi-
Genetics, 3, e33. tional syntenic cloning and functional characterization
Husse, J., Zhou, X., Shostak, A., Oster, H., & Eichele, G. of the mammalian circadian mutation tau. Science, 288,
(2011). Synaptotagmin10-Cre, a driver to disrupt clock 483–492.
genes in the SCN. Journal of Biological Rhythms, 26, Maier, B., Wendt, S., Vanselow, J. T., Wallach, T., Reischl, S.,
379–389. Oehmke, S., et al. (2009). A large-scale functional RNAi
Izumo, M., Johnson, C. H., & Yamazaki, S. (2003). Circadian screen reveals a role for CK2 in the mammalian circadian
gene expression in mammalian fibroblasts revealed by clock. Genes & Development, 23, 708–718.
264

Mallon, A. M., Blake, A., & Hancock, J. M. (2008). Quwailid, M. M., Hugill, A., Dear, N., Vizor, L., Wells, S.,
EuroPhenome and EMPReSS: Online mouse phenotyping Horner, E., et al. (2004). A gene-driven ENU-based
resource. Nucleic Acids Research, 36, D715–D718. approach to generating an allelic series in any gene. Mam-
Marcheva, B., Ramsey, K. M., Buhr, E. D., Kobayashi, Y., malian Genome, 15, 585–591.
Su, H., Ko, C. H., et al. (2010). Disruption of the clock com- Ralph, M. R., & Menaker, M. (1988). A mutation of the
ponents CLOCK and BMAL1 leads to hypoinsulinaemia circadian system in golden hamsters. Science, 241, 1225–1227.
and diabetes. Nature, 466, 627–631. Rinchik, E. M., Bangham, J. W., Hunsicker, P. R.,
Maywood, E. S., Chesham, J. E., Meng, Q. J., Nolan, P. M., Cacheiro, N. L., Kwon, B. S., Jackson, I. J., et al. (1990).
Loudon, A. S., & Hastings, M. H. (2011). Tuning the period Genetic and molecular analysis of chlorambucil-induced
of the mammalian circadian clock: Additive and indepen- germ-line mutations in the mouse. Proceedings of the
dent effects of CK1epsilonTau and Fbxl3Afh mutations National Academy of Sciences of the United States of
on mouse circadian behavior and molecular pacemaking. America, 87, 1416–1420.
The Journal of Neuroscience, 31, 1539–1544. Ringwald, M., Iyer, V., Mason, J. C., Stone, K. R.,
Mcdearmon, E. L., Patel, K. N., Ko, C. H., Walisser, J. A., Tadepally, H. D., Kadin, J. A., et al. (2011). The IKMC
Schook, A. C., Chong, J. L., et al. (2006). Dissecting web portal: A central point of entry to data and resources
the functions of the mammalian clock protein BMAL1 by from the International Knockout Mouse Consortium.
tissue-specific rescue in mice. Science, 314, 1304–1308. Nucleic Acids Research, 39, D849–D855.
Meng, Q. J., Logunova, L., Maywood, E. S., Gallego, M., Rosbash, M., & Takahashi, J. S. (2002). Circadian rhythms:
Lebiecki, J., Brown, T. M., et al. (2008). Setting clock speed The cancer connection. Nature, 420, 373–374.
in mammals: The CK1 epsilon tau mutation in mice Rossner, M. J., Oster, H., Wichert, S. P., Reinecke, L.,
accelerates circadian pacemakers by selectively destabilizing Wehr, M. C., Reinecke, J., et al. (2008). Disturbed clock-
PERIOD proteins. Neuron, 58, 78–88. work resetting in Sharp-1 and Sharp-2 single and double
Morgan, H., Beck, T., Blake, A., Gates, H., Adams, N., mutant mice. PLoS One, 3, e2762.
Debouzy, G., et al. (2010). EuroPhenome: A repository for Schmutz, I., Ripperger, J. A., Baeriswyl-Aebischer, S., &
high-throughput mouse phenotyping data. Nucleic Acids Albrecht, U. (2010). The mammalian clock component
Research, 38, D577–D585. PERIOD2 coordinates circadian output by interaction
Nakahata, Y., Kaluzova, M., Grimaldi, B., Sahar, S., with nuclear receptors. Genes & Development, 24,
Hirayama, J., Chen, D., et al. (2008). The NADþdependent 345–357.
deacetylase SIRT1 modulates CLOCK-mediated chromatin Shearman, L. P., Jin, X., Lee, C., Reppert, S. M., &
remodeling and circadian control. Cell, 134, 329–340. Weaver, D. R. (2000). Targeted disruption of the mPer3
Nolan, P. M., Peters, J., Strivens, M., Rogers, D., Hagan, J., gene: Subtle effects on circadian clock function. Molecular
Spurr, N., et al. (2000). A systematic, genome-wide, and Cellular Biology, 20, 6269–6275.
phenotype-driven mutagenesis programme for gene func- Shi, S., Hida, A., Mcguinness, O. P., Wasserman, D. H.,
tion studies in the mouse. Nature Genetics, 25, 440–443. Yamazaki, S., & Johnson, C. H. (2010). Circadian clock
Nolan, P. M., Peters, J., Vizor, L., Strivens, M., gene Bmal1 is not essential; functional replacement with its
Washbourne, R., Hough, T., et al. (2000). Implementation paralog, Bmal2. Current Biology, 20, 316–321.
of a large-scale ENU mutagenesis program: Towards Sidman, R. L., Lane, P. W., & Dickie, M. M. (1962). Staggerer,
increasing the mouse mutant resource. Mammalian a new mutation in the mouse affecting the cerebellum.
Genome, 11, 500–506. Science, 137, 610–612.
Okano, S., Akashi, M., Hayasaka, K., & Nakajima, O. (2009). Siepka, S. M., Yoo, S. H., Park, J., Lee, C., & Takahashi, J. S.
Unusual circadian locomotor activity and pathophysiology (2007). Genetics and neurobiology of circadian clocks in
in mutant CRY1 transgenic mice. Neuroscience Letters, mammals. Cold Spring Harbor Symposia on Quantitative
451, 246–251. Biology, 72, 251–259.
Oster, H., Baeriswyl, S., Van Der Horst, G. T., & Albrecht, U. Silver, L. M. (1995). Mouse genetics. Concepts and
(2003). Loss of circadian rhythmicity in aging mPer1/ applications. Oxford University Press: New York, USA.

mCry2/ mutant mice. Genes & Development, 17, Skarnes, W. C., Rosen, B., West, A. P., Koutsourakis, M.,
1366–1379. Bushell, W., Iyer, V., et al. (2011). A conditional knockout
Oster, H., Yasui, A., Van Der Horst, G. T., & Albrecht, U. resource for the genome-wide study of mouse gene function.
(2002). Disruption of mCry2 restores circadian rhythmicity Nature, 474, 337–342.
in mPer2 mutant mice. Genes & Development, 16, 2633–2638. Stottmann, R. W., Moran, J. L., Turbe-Doan, A., Driver, E.,
Picciotto, M. R., & Wickman, K. (1998). Using knockout and Kelley, M., & Beier, D. R. (2011). Focusing forward genet-
transgenic mice to study neurophysiology and behavior. ics: A tripartite ENU screen for neurodevelopmental
Physiological Reviews, 78, 1131–1163. mutations in the mouse. Genetics, 188, 615–624.
265

Van Der Horst, G. T., Muijtjens, M., Kobayashi, K., CKIdelta mutation causing familial advanced sleep phase
Takano, R., Kanno, S., Takao, M., et al. (1999). Mammalian syndrome. Nature, 434, 640–644.
Cry1 and Cry2 are essential for maintenance of circadian Xu, Y., Toh, K. L., Jones, C. R., Shin, J. Y., Fu, Y. H., &
rhythms. Nature, 398, 627–630. Ptacek, L. J. (2007). Modeling of a human circadian muta-
Vitaterna, M. H., King, D. P., Chang, A. M., tion yields insights into clock regulation by PER2. Cell,
Kornhauser, J. M., Lowrey, P. L., Mcdonald, J. D., et al. 128, 59–70.
(1994). Mutagenesis and mapping of a mouse gene, Clock, Zheng, B., Albrecht, U., Kaasik, K., Sage, M., Lu, W.,
essential for circadian behavior. Science, 264, 719–725. Vaishnav, S., et al. (2001). Nonredundant roles of the mPer1
Wulff, K., Gatti, S., Wettstein, J. G., & Foster, R. G. (2010). and mPer2 genes in the mammalian circadian clock. Cell,
Sleep and circadian rhythm disruption in psychiatric and 105, 683–694.
neurodegenerative disease. Nature Reviews. Neuroscience, Zheng, B., Larkin, D. W., Albrecht, U., Sun, Z. S., Sage, M.,
11, 589–599. Eichele, G., et al. (1999). The mPer2 gene encodes a func-
Xu, Y., Padiath, Q. S., Shapiro, R. E., Jones, C. R., Wu, S. C., tional component of the mammalian circadian clock. Nature,
Saigoh, N., et al. (2005). Functional consequences of a 400, 169–173.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 16

In search of a temporal niche: Social interactions

Alexandra Castillo-Ruiz{,*, Matthew J. Paul{ and William J. Schwartz{

{
Department of Neurology, University of Massachusetts Medical School, Worcester, MA, USA
{
Department of Psychology, University of Massachusetts Amherst, Amherst, MA, USA

Abstract: Circadian rhythms can be entrained to periodic cues in the environment including the solar
day, food resources, and temperature. Work on a variety of organisms has suggested that social
interactions within and between species may also influence circadian rhythmicity, but conceptual and
technical difficulties relating to animal models, housing environments, rhythm assays, and experimental
design have complicated mechanistic investigations in the laboratory. We review these issues and
introduce the gregarious Nile grass rat, Arvicanthis niloticus, as a suitable model for research on this
problem. Understanding social influences on temporal organization at this supra-organismal,
community level is of considerable translational value, as its implications range from conservation
biology to human health.

Keywords: circadian clock; circadian rhythm; entrainment; masking; behavioral ecology.

Introduction environment. These temporal programs are not


rigid, however, and respond to changes in photo-
Circadian clocks are pervasive across all period, food resources, temperature, as well as
kingdoms of life, from unicellular to multicellular to other organisms. The influence of abiotic
organisms. Their ubiquity has been taken as evi- factors on circadian rhythmicity is reviewed by
dence for their adaptive value. It is believed that Hut and collaborators in this issue (Hut et al.
circadian rhythms evolved not only to provide 2012), while here we focus on inter-organismal
an internal temporal organization for physiologi- effects on circadian clock function, a topic which
cal and behavioral processes but also to help we term socio-chronobiology.
organisms anticipate challenges and opportunities Unlike the typically uneventful life in the labo-
that occur predictably and periodically in the ratory setting, life in the natural world is a con-
stant struggle to survive, procure resources, and
*Corresponding author. reproduce. Survival hinges on finding favorable
Tel.: þ1-508-856-1128; Fax: þ1-508-856-6778 spatial and temporal niches, and conspecific or
E-mail: alexandra.castillo-ruiz@umassmed.edu heterospecific individuals can impact the niche

http://dx.doi.org/10.1016/B978-0-444-59427-3.00016-2 267
268

through cooperation, competition, or predation, of symbiont population size. Study of gene expres-
among other interactions. Previous critical reviews sion patterns revealed corresponding rhythms of
have discussed the effects of social cues on circadian transcription that time the changes in squid epithe-
rhythmicity (Davidson and Menaker, 2003; Favreau lial (light organ) anatomy with switches in bacterial
et al., 2009; Mistlberger and Skene, 2004; Regal and metabolic pathways. Although the cues that couple
Connolly, 1980), but they have focused mainly on such interlocked rhythms are not known, or
whether social cues can entrain organisms, sp., on whether they are circadian clock-driven, these
their ability to act as zeitgebers to synchronize the studies of mutualism suggest that temporal coordi-
behavioral rhythms of conspecifics. Here, we pro- nation may be key for the evolution and conserva-
vide some examples in different taxa of the wide tion of interspecific alliances.
range of social interactions that might be considered
to be socio-chronobiological, outline the search for
underlying mechanisms, and highlight the concep- Antagonistic interactions
tual and technical difficulties that have complicated
research progress. Among species, sharing the same time preference
for activity can have detrimental consequences
for one of the parties, such as infestation by a par-
Interspecific relationships and circadian rhythms asite, death from starvation, or being eaten by a
predator.
Mutualism

Cooperation between individuals of different Parasitism


species is known as mutualism. Theoretically,
the coordination of timing of behavioral and Parasites depend on their hosts for survival and
physiological events between allies in a beneficial reproduction. Presumably, they have been sub-
relationship ought to be extremely important for jected to a strong selective pressure to evolve
conservation of the partnership. We are aware traits that increase their chances of interactions
of only a handful of studies that have investigated with hosts, and one of these traits likely involves
circadian rhythms in mutualistic species. As an temporal coordination of behavior. For instance,
example, Flint et al. (2005) found that bacterial schistosomiasis is a parasitic disease caused by
symbionts (Epulopiscium spp.) living in the intes- trematodes of the genus Schistosoma. The worms
tinal tract of the surgeonfish, Naso lituratus, show have complex life cycles, in which eggs released
a prominent daily cycle, remaining dormant dur- into fresh water hatch into miracidia, infect an
ing the night (when the fish is mostly inactive) intermediate (snail) host to develop as sporocysts,
and proliferating during the day. The authors and then emerge as cercariae to infect and produce
argue that this temporal pattern may be driven by eggs in their definitive mammalian host. Cercarial
the lower availability of host nutrients during the emergence is rhythmic, appears to be genetically
night. Wier et al. (2010) analyzed the symbiotic determined (Theron and Combes, 1988), and
relationship between the Hawaiian squid, may exhibit a phase that matches the activity
Euprymna scolopes, and its symbiont, the lumi- rhythm of the intended host (for Schistosoma
nous bacteria Vibrio fischerii. Rhythmic bacterial japonicum, late afternoon for a nocturnal rodent
bioluminescence, highest during the night, is used reservoir, early morning for a diurnal cattle reser-
by the host for its nocturnal behaviors; and rhyth- voir; Lu et al., 2009). The possible significance of
mic bacterial expulsion from the squid’s light- such coordination is exemplified by a recent study
emitting organ, highest at dawn, insures stability (O’Donnell et al., 2011) demonstrating that fitness
269

in malaria parasites (Plasmodium chabaudi) is activity patterns. Two species of closely related
increased when both the parasites and their rodent spiny mice (Acomys cahirinus and A. russatus)
hosts are synchronized to the same light–dark cohabit the same rocky terrain of the Dead Sea
(LD) cycle as compared to when they are under Valley, with overlap in microhabitat use, home
a mismatch condition. Such data hint that circadian ranges, food habits, and reproductive period (for
rhythms may be a force driving evolution. review, see Kronfeld-Schor and Dayan, 2008).
As with mutualism, the mechanisms underlying Notably, A. cahirinus is nocturnal and A. russatus
such interspecific coordination are not known. is diurnal. Shkolnik (1971) attributed this pattern
Toledo et al. (1999) examined the rhythmic pat- to interspecific competition; upon removal of
tern of larval emergence of the intestinal parasite A. cahirinus from the field site, A. rusattus expands
Euparyphium albuferensis; this rhythm does not its activity profile to the night, suggesting that the
appear to be under the control of an endogenous presence of A. cahirinus prevents A. russatus from
clock because it cannot be identified under con- being active at its preferred phase of the daily
stant conditions. Future studies on the chronobi- cycle. When A. cahirinus is released back into the
ology of parasitism will need to assess the field site, both species reestablish their temporal
rhythmicity of hosts as well as of parasites, given segregation. The nature of the limiting resource
that parasites may affect the rhythms of their (s) is unknown. More recently, laboratory studies
hosts (Lima-Camara et al., 2011). of Acomys have been informative and are
reviewed in a later section of this review.
A second example relates to the American
Competition mink (Neovison vison), which was introduced to
Great Britain in the early 1900s and became
Competition in the face of limited resources can established in the absence of its natural com-
be mitigated by the separation of species in space petitors, the Eurasian otter (Lutra lutra) and the
and by diet but also in time (Schoener, 1986). For European polecat (Mustela putoris). Under these
example, a redistribution of nighttime activity conditions, Harrington et al. (2009) observed in
patterns has been demonstrated in Gerbillus the 1990s that the mink was predominantly active
allenbyi and G. pyramidum, two species of sand during the night. Ten years later, however, after
dune gerbils, studied separately or together in the native competitor populations had recovered,
100 m2 seminatural grids. When separated, the mink appeared to have switched to a day-
both species show nocturnal activity mostly con- active pattern.
fined to the first part of the night; but when To our knowledge, only one laboratory study
together, the activity of the subordinate species, has explored whether time-dependent exposure
G. allenbyi, is displaced to the second half of the to an interspecific competitor can induce acute
night (Ziv et al., 1993). Temporal partitioning phase shifts of activity rhythms. Scheibler and
also occurs between the cotton rat (Sigmodon Wollnik (2009) introduced a Mongolian gerbil
hispidus) and prairie vole (Microtus ochrogaster) (Meriones unguiculatus) housed in a small cage
studied in 4 m2 arenas, with voles becoming most into the home cage of a desert hamster (Pho-
active at phases when rats are least active. Spatial dopus roborovskii) for 2 h at eight time points
partitioning arises additionally only in the pres- spanning the 24-h circadian cycle in constant
ence of reproductive rats, which show higher darkness; in this pairing, gerbils are believed to
levels of interspecific aggression than nonrepro- be dominant and hamsters subordinate. The
ductive rats (Glass and Slade, 1980). authors reported that hamsters exhibited phase
Even more dramatic are reports of interspecific advances of their circadian wheel-running rhythm
interactions that lead to an inversion of daily during the subjective day and mid-subjective
270

night. Shortening of hamster free-running period caused by changes in the strength and/or sign of
with increased wheel running was also observed coupling mechanisms that link the clock to subor-
after gerbils were introduced during the subjective dinate downstream oscillators responsible for
day, perhaps related to increased arousal. rest-activity rhythms? Or do social interactions
modulate or modify rhythm phase and waveform
directly, by circumventing or overriding the clock
Predator–prey (a mechanism referred to as masking), rather than
by resetting (entraining) the clock’s oscillation?
Throughout the day/night cycle there are times at
which activity outside the nesting site or burrow
can be extremely dangerous due to the presence Conspecific relationships and circadian rhythms
of predators. Thus, most predator avoidance
strategies involve the avoidance of increased pre- Eusocial insects
dation risk times, and it has been postulated that
these strategies have enabled the coexistence of Of insects that live in colonies with highly devel-
predator and prey (for review, see Kronfeld- oped social structures—including cooperative
Schor and Dayan, 2003). The sudden arrival of a care of the young and individual division of
novel predator to a species niche could poten- labor—the most extensively studied from a socio-
tially induce a redistribution of activity to less chronobiological view is the honeybee (Apis
dangerous times, and examples of such behavioral mellifera; for reviews, see Bloch, 2010; Moore,
plasticity have been reported. Naturally, noctur- 2001). Worker honeybees include younger nurses,
nal wild rats (Rattus norvegicus) become diurnal which care for the brood continuously and are active
when a predator, the red fox (Vulpes vulpes), is during both the day and night; and older foragers,
in the vicinity. When exposure to the fox is exper- which periodically leave the hive and express robust
imentally prevented, the rats revert to their noc- circadian activity rhythms. Social mediation of this
turnal behavior (Fenn and Macdonald, 1995). temporal organization becomes apparent when
A similar phenomenon was observed for Santa the colony experiences a shortage of nurses. Some
Cruz foxes (Urocyon littoralis; Swarts et al., foragers return to nursing and adopt an arrhythmic,
2009). In the 1990s, before predatory golden eagles persistently active state (Bloch and Robinson,
(Aquila chrysaetos) colonized Santa Cruz Island, 2001). Nursing arrhythmicity represents a context-
foxes were active at all times of day and night. dependent behavioral plasticity, since robust circa-
After the eagles arrived, foxes showed a prefer- dian rest-activity rhythms can be expressed by
ence for nocturnal activity. The authors argued nurse-age bees if they are restricted to broodless
that this change was related to predator-induced combs or transferred from the cyclically illuminated
selection for a nocturnal trait but a more recent hive to constantly dark laboratory conditions
report demonstrates that eagle removal efforts (Shemesh et al., 2010). These authors have also
have been associated with a rapid vulpine rever- provided indirect evidence to suggest that
sion to diurnality (Hudgens and Garcelon, 2011). arrhythmicity induced by brood contact is not due
The neurobiological mechanisms responsible for to an arrest of the central circadian pacemaker; it
adaptive temporal switches in response to interspe- seems instead that at least some clock cells must be
cific antagonistic interactions are unknown. Specif- keeping track of the passage of time. The lack of
ically, might they represent altered phase angles of overt behavioral rhythmicity could be the result
entrainment to the day–night cycle secondary to of desynchronized or low amplitude cellular
changes in the circadian clock’s free-running oscillators, and/or masking or uncoupling of oscilla-
period, perhaps related to arousal? Or are they tor outputs from downstream effectors.
271

Maternal entrainment birth and postnatal maternal influences on pup


rhythmicity appear to be minimal (Weaver and
In mammals, the circadian clock in the fetus Reppert, 1987). In a marsupial species, the gray
appears to be oscillating in utero and entrained by short-tailed opossum (Monodelphis domestica),
the mother’s rhythm; in altricial rodent species the offspring are born extremely immature but
(sp., hamsters, mice, and rats), there is also evi- are directly exposed to the environmental LD
dence that mothers entrain the rhythms of their cycle while attached to their pouchless mothers;
immature offspring during the postnatal period, in this species also, the dam wields no influence
before the pups become responsive to the external on the timing of pup rhythmicity (Rivkees and
LD cycle (for review, see Davis and Reppert, Reppert, 1990). At the other extreme is the Euro-
2001). The influence of maternal care has been pean rabbit (Oryctolagus cuniculus): rabbit
examined by cross-fostering and maternal depriva- mothers are absent from the nest except to nurse
tion studies; in the former, prenatal natural their pups for a 3- to 4- min interval once daily
mothers are replaced by postnatal foster mothers, (for review, see Jilge and Hudson, 2001). An
and in the latter, nursing mothers are cyclically endogenous anticipatory rhythm in the pups is
present or absent from their pups. Maternal rhyth- entrained to this maternal rhythm, including a
micity is required for pup synchronization (Davis rhythm of “clock” gene expression in their SCN
and Gorski, 1988; Jud and Albrecht, 2006; Reppert (Caldelas et al., 2009).
and Schwartz, 1986), and in its absence pups within
a litter are unable to entrain each other’s rhythms
(Hiroshige et al., 1982; Reppert et al., 1984). Ohta Social synchronization
et al. (2003) studied the entrainment of two groups
of neonatal rats (Rattus norvegicus) blinded by Social synchronization of the temporal behaviors
bilateral orbital enucleation at birth, and subse- of adult conspecifics should allow for the realiza-
quently, nursed by foster mothers entrained to a tion of common goals, for example, defending
LD cycle; in one of the two groups, mothers were against predators, hunting in packs, or huddling
removed during the light phase (when they nor- as a means of energy conservation. While there
mally would be nursing) and returned during the have been tantalizing hints of social entrainment
dark phase each day for 6 days, after which they in some (but not all) studies, it has been difficult
were present continually until the pups were to find robust, reliable, and experimentally tracta-
weaned at postnatal day 20. This maternal depriva- ble animal models for further analysis.
tion schedule resulted in the phase reversal of pup
rhythms of “clock” gene expression in the
suprachiasmatic nucleus (SCN) on postnatal day 6 Field and laboratory
and of locomotor activity rhythms after weaning.
While the data are consistent with maternal In the field, temporal synchronization among
entrainment of the molecular circadian clock in animals has been inferred from studies of Canadian
the neonatal SCN, the authors also noted the possi- beavers (Castor canadensis) living together as a
bility that entrainment may not have been via family unit, as they exhibit a single coherent free-
maternal time cues but instead due to the stress running rhythm in daily noise while they overwinter
that was associated with her periodic absence. in constant darkness under thick ice (Bovet and
The strength and persistence of postnatal Oertli, 1974). In the laboratory, several studies have
maternal entrainment differs among species. In a demonstrated that co- or group-housed animals—
precocious rodent species, the common spiny including killifish (Fundulus heteroclitus; Kavaliers,
mouse (A. cahirinus), the eyes open shortly after 1980b), deer mice (Peromyscus maniculatus;
272

Crowley and Bovet, 1980), palm squirrels other was blinded and free ran with a period
(Funambulus pennanti; Rajaratnam and Redman, > 24 h. There was no synchronization or relative
1999), honey bees (Apis mellifera; Frisch and coordination of the activity of the blind hamster
Koeniger, 1994), and fruit flies (Drosophila to that of the entrained one. The conclusion from
melanogaster; Levine et al., 2002)—may display all these studies has been that social factors can
coherent or converging activity rhythms in constant provide only weak inputs to the circadian clock.
lighting conditions but their reliance on group activ-
ity data has generally prohibited much further
mechanistic analyses. The expression of a single Mixed sex interactions
free-running rhythm in group-housed individuals
could be due to social entrainment (mutual or unidi- Strategies similar to those described above have
rectional), but it could also reflect similar phases and been employed to study the possible interactions
periods before cohabitation or a masking effect of of paired male and female hamsters or degus.
one, possibly socially dominant, individual. Male and female hamsters either closely or co-
Researchers have used different strategies to dis- housed without physical contact, either continu-
tinguish individual activity under social conditions. ally for 6 weeks (Davis et al., 1987) or for 3 h
Some have housed animals in close proximity to, per day for intervals ranging from about 2 to 6
but not in direct contact with, each other, or weeks (Aschoff and von Goetz, 1988), show no
together but separated by a barrier (wire mesh, change in the free-running periods of their activ-
clear perforated plastic, or a restraining cage); ity rhythms, although the males show masking of
others have restricted the interval of social contact their rhythms by the females. Honrado and
to a few hours per day. Under such housing Mrosovsky (1989) exposed males for 3 h to
conditions, social effects on the circadian system estrous females in restraining cages—or for 1 h
have been inconsistent (e.g., in golden hamsters, to actually mate—and reported more rapid
Mesocricetus auratus; Davis et al., 1987; Gattermann reentrainment to a phase-advanced LD cycle for
and Weinandy, 1997; Mistlberger et al., 2003; the exposed but not for the mated males. When
Mrosovsky, 1988; Refinetti et al., 1992) or relatively these authors entrained males to an LD cycle
modest, with observed alterations in activity levels, and then exposed them to females in restraining
reentrainment rates to a photically induced phase cages for 1 h daily at a phase 4 h before lights-
shift, or free-running periods in constant darkness off, the males exhibited advances in locomotor
(e.g., in degus, Octodon degus; Goel and Lee, activity onsets that habituated after repeated
1995, 1996, 1997). One study has demonstrated syn- exposure (Honrado and Mrosovsky, 1991). In
chronous activity due to social cues in individuals degus housed together (separated by a wire mesh
separated by a barrier. Schneider’s roundleaf bats barrier), females reentrain faster to a phase-
(Hipposideros speoris) individually caged in a cave advanced LD cycle when co-housed with another
without access to light synchronize their activity female but not with a male (Goel and Lee, 1995).
rhythms to local time, in phase with their free-living
conspecifics that leave the nest daily around sunset;
a control bat caged in a cave devoid of other Challenges in bringing social interactions into
bats instead exhibited a free-running rhythm the laboratory
(Marimuthu et al., 1981).
Refinetti et al. (1992) co-housed two hamsters While experimentation in the laboratory has obvi-
in direct physical contact with each other but with ous advantages, especially for elucidation of pos-
markedly different free-running periods: one was sible molecular and neurobiological mechanisms,
entrained to a short 23.3 h LD cycle while the it is an unnatural setting. Researchers intent on
273

analyzing social interactions under these con- field mice (Apodemus sylvaticus), in which three
ditions need to be aware of special considerations subordinate mice were observed to avoid the time
with regard to their choice of animal subjects, of peak out-of-burrow activity of a dominant
caged environment, selected assays, and experi- mouse (Bovet, 1972). In rats, a single episode of
mental design. social defeat influences the amplitude of locomo-
tor, heart rate, and temperature rhythms for sev-
eral cycles, apparently by a masking mechanism
Animals (Meerlo et al., 1999). And Hansen and Closs
(2005), investigating how food depletion can alter
One difficulty with domesticated animals is that the daily activity pattern in the giant kokopu
their physiology and behavior may be quite dif- (Galaxias argenteus), demonstrated that domi-
ferent than that of their counterparts in the nant fish are the ones that switch from nocturnal
wild (for review, see Calisi and Bentley, 2009). to diurnal activity when food is limited.
It is now well known that some prototypically
“nocturnal” laboratory rodents are not necessar-
ily nocturnal in their native habitat (e.g., emer- Environment
gence from the burrow in the golden hamster;
Gattermann et al., 2008); laboratory-bred and In the field, species must make complex decisions,
wild-derived hamsters also differ in other physio- “eat or be eaten,” so the relative safety of the lab
logical and behavioral traits (Krause and undoubtedly eliminates most of the abiotic and
Schuler, 2010). Golden spiny mice (A. russatus) biotic factors that drive the phenotypic plasticity
and cururos (Spalacopus cyanus), while nocturnal of animals in their natural temporal niche. In a
in captivity (Begall et al., 2002; Cohen and wild population with underlying genetic variation,
Kronfeld-Schor, 2006), are evidently diurnal traits that are induced in response to such envi-
when free-living (Levy et al., 2007; Urrejola ronmental factors may then become genetically
et al., 2005). Obviously, these kinds of differences determined, by a process of directional selection
between laboratory and field could also impact on extreme phenotypes (Ghalambor et al.,
the effects, and even the presence, of social 2007). In this way, chronotypic plasticity that
influences on rhythms. was originally induced by a social interaction
Not often considered in socio-chronobiological might become selectively assimilated over time
experiments is whether the subjects are familiar as an evolved circadian activity pattern, that is,
with each other. The significance of this variable as a “ghost of selection past” (Ghalambor et al.,
is exemplified by work with common marmosets 2007) that outlasts the initial social stimulus. It is
(Callithrix j. jacchus), which are social primates; now appreciated that such adaptive evolution
paired animals may synchronize their free- can occur on ecological time scales (Carroll
running circadian periods if they are in full social et al., 2007).
contact and know each other prior to the experi- Consider the two species of spiny mice
ment (Erkert and Schardt, 1991). Awareness of (A. cahirinus and A. russatus) described in an ear-
sex differences in activity patterns in the wild lier section of this review; A. cahirinus is noctur-
(e.g., as dramatically expressed by some mustelid nal while A. russatus is diurnal, and observations
species; Marcelli et al., 2003; Zschille et al., 2010) in the 1970s suggested that the presence of
is also important for interpreting observed A. cahirinus competitively displaces A. russatus
patterns in the lab. Lastly, dominant/subordinate from the nocturnal niche (Shkolnik, 1971).
relationships must be considered. There is a Indeed, individually caged A. russatus maintained
report of laboratory group housing of long-tailed in the laboratory in a 12 h:12 h LD cycle primarily
274

exhibits nocturnal activity and temperature males (Liu, 2004). In black-capped chickadees
rhythms (Cohen and Kronfeld-Schor, 2006). (Poecile atricapillus) it is the phase (onset) of the
However, A. russatus exposed to A. cahirinus in dawn chorus that is modified (advanced) by the
the room, or in the same cage, does not become appearance of a simulated conspecific competitor
diurnal (Cohen, 2004); and, on the other hand, (Foote et al., 2011).
A. russatus held in 1000 m2 field enclosures in
the Judean desert, without A. cahirinus present,
is diurnal, becoming immediately nocturnal when Design
introduced into the laboratory (Levy et al., 2007).
Thus, diurnality in A. russatus does not appear to Assuming that “social coupling” is relatively weak
be fully explained by behavioral plasticity in (at least as compared to light), and that individual
response to interference or resource competition animals represent nonidentical oscillators
with A. cahirinus. Perhaps such competition in governed by their own intrinsic properties, then
the past, and/or other nonsocial factors, produced one would predict that strong interindividual
selective forces to which A. russatus became effects might be precipitated only under particu-
highly adapted. lar circumstances, sp., when oscillators are in
direct physical contact for a relatively long period
of time and their respective free-running periods
Assays are relatively close to one another (Winfree,
2001). Thus, actual prolonged cohabitation might
If the experimental question is whether a social lead to robust effects on the rhythmicity of co-
cue is capable of acting as an independent zeitge- housed individuals, effects not easily reproduced
ber for entraining the master circadian clock in by indirect or pulsatile social contacts. As
the SCN, then longitudinal assessment of a single, described previously in this review, relatively few
precise, circadian marker rhythm that exhibits studies have recorded the rhythms of individuals
unambiguous phase and period (e.g., wheel- during long-term cohabitation.
running activity in rodents) makes sense. But if Theoretically, network strength under such
social forces act on alternative substrates—on conditions should increase with the number of
rhythm amplitude or waveform, phase of entrain- coupled oscillators (Kuramoto, 1984). Indeed,
ment to a geophysical cycle, or mechanisms groups of fruit flies (Levine et al., 2002), colonies
downstream or separate from those that generate of honey bees (Frisch and Koeniger, 1994), and
the locomotor rhythm—inspection of circadian shoals of white suckers (Catostomus commersoni
wheel-running records will be an insensitive Kavaliers, 1980a) do exhibit more precise and/or
and/or irrelevant assay. In root voles (Microtus stable circadian periods than do the individual
oeconomus) under winter snow and lacking daily flies, bees, or fish. At a different level of biological
photo-entrainment, social interactions synchronize organization, similar effects have been reported
the activities of individuals at the ultradian rather for SCN neuronal oscillators (Herzog et al., 2004;
than the circadian level (Korslund, 2006). The right Liu et al., 2007; Webb et al., 2009).
assays are also required for understanding the Advances have been made in identifying candi-
effects of social cues on the timing of bird song. Male date cues that might mediate social coupling,
chipping sparrows (Spizella passerina) shorten their especially chemical (in spiny mice: Fluxman and
dawn chorus in response to the removal of neigh- Haim, 1993; Haim and Rozenfeld, 1993; in degus:
boring males but not to the removal of their female Governale and Lee, 2001; Jechura et al., 2006; in
mates; they do increase their daytime singing fruit flies: Krupp et al., 2008), and gonadal hor-
after removal of their mates but not neighboring monal (in degus: Jechura and Lee, 2004; Jechura
275

et al., 2003). But it is probable that multiple inter- show behavioral plasticity in the lab, with a noc-
acting signals play physiological roles, as is likely turnal activity pattern upon exposure to a running
the case for melatonin and food intake in mater- wheel, returning to a diurnal pattern after wheel
nal-fetal entrainment (for review, see Davis and removal (Blanchong et al., 1999). Another nota-
Reppert, 2001). Arousal may be a critical variable. ble trait of the grass rat is its social behavior. Wild
Some bird species are reported to entrain to colonies consist of several males, females, and
daily playbacks of conspecific birdsong (Gwinner, their offspring; adult males migrate to other col-
1966; Menaker and Eskin, 1966), but house onies while adult females do not disperse (Packer,
sparrows (Passer domesticus) can also entrain to 1983; Senzota, 1990). Additionally, although little
periodic noise (Reebs, 1989), highlighting a possi- is known about the male’s role in parental care in
bly permissive role for nonspecific arousal. the wild, observations in captivity suggest that he
participates in the rearing of offspring; in the lab,
the male not only nests with the female and her
For the future offspring but also grooms the pups even as
weanlings. Importantly, due to continuous efforts
We have recently begun a research program to prevent inbreeding in captivity, grass rats are
in mammals (initially in hamsters; Paul and more likely than the nocturnal rodents commonly
Schwartz, 2007) that seeks to address some of used in circadian research to resemble their wild
these socio-chronobiological challenges. counterparts.
The unstriped Nile grass rat, Arvicanthis In order to longitudinally measure rhythmicity in
niloticus (Fig. 1), native to sub-Saharan Africa, individual, co-housed grass rats, we are implanting
was introduced to circadian rhythms research in miniature temperature sensors (ibuttons; Dallas
the early 1990s (Katona and Smale, 1997) and Semiconductor DS1922L-F5, Maxim Integrated
should be an interesting animal subject for future Products, Inc., Sunnyvale, CA), which are com-
research on social interactions and rhythms. It mercially available for this purpose. Temperature
has already served as a useful model for studies can be recorded at an accuracy of 0.5  C from
on the neural substrates of diurnality, as it  10  C to þ 65  C and a resolution of 0.0625  C
exhibits robust diurnal activity patterns in the by a digital thermometer at a sampling rate
field and in the lab (Blanchong and Smale, 2000; from 1 s to 273 h. Each ibutton contains 8 kB of
McElhinny et al., 1997). Notably, some individuals datalog memory for storage of up to 4096 16-bit
temperature readings. Data collection can begin
immediately after implantation or programmed
to begin after a user-defined delay. The device is
encased in stainless steel, resulting in a height,
radius, and weight of 5.89 mm, 17.35 mm, and
2.95 g, respectively.
Of note, we believe that a crucial part of our
program will be the application of a rigorous
quantitative method—based on wavelet analysis—
that will allow us to analyze rhythm phase, ampli-
tude, and period at multiple scales. In fact, other
circadian researchers have begun to use wavelet
transforms, for their rigor, utility, and minimal
assumptions for analyzing nonstationary datasets
Fig. 1. Grass rat family in captivity (photo courtesy of A. C. R.). (Leise and Harrington, 2011). Wavelet transforms
276

do not have a single set of basis functions (like not be the most informative animals to investigate
the Fourier transform, which utilizes only sines these questions (Smale et al., 2005), in part because
and cosines) but an infinite set of possible basis countless generations have been selected for par-
functions. ticular traits (like tameness).
Socio-chronobiology has enormous transla-
tional value. Its implications range from pest con-
Coda trol (e.g., to plan strategies to control population
growth), animal welfare (e.g., to identify the best
Understanding how circadian rhythmicity is housing situation for laboratory animals), conser-
endogenously generated and entrained to the vation biology (e.g., to understand human effects
LD cycle has been the focus of considerable on animal populations, like coyotes, Canis latrans;
research—from molecular to organismal levels— Kitchen et al., 2000), and clinical medicine (e.g.,
for the past several decades. Relatively, less to develop strategies to prevent and control para-
attention has been directed to investigating the sitic infection while enhancing the activities of
generation and entrainment of rhythmicity at beneficial gut flora). Many blind people somehow
the community level. We know relatively little entrain to the 24-h day (Klerman et al., 1998),
about the importance and mechanisms that regu- presumably through nonphotic cues, and a better
late the timing of social interactions, whether syn- understanding of the role of social influences may
chronizing or segregating; or the role of circadian lead to new strategies for healthy adjustment to
timekeeping at the supra-organismal level; or how life on a rotating world.
the interaction of social and circadian systems
ensures adaptation and economy in the temporal
domain. New ideas, technologies, and paradigms Acknowledgments
are now emerging to build on the pioneering
work reviewed here. Chronobiologists have This work is supported in part by R01 GM094109
already benefited from their collaborations with (W. J. S.). The contents of this review are solely
molecular biologists, photobiologists, and applied the responsibility of the authors and do not neces-
mathematicians, and we are looking forward to sarily represent the official views of the National
expanded interactions with behavioral ecologists, Institutes of Health.
evolutionary biologists, and social neuro-
scientists. And in so doing, we need to attend
to ultimate-level (why) questions in addition to
proximate-level (how) ones. Abbreviations
Progress is also being catalyzed by new attitudes.
While past research has concentrated on entrain- LD light–dark
ment of the clock as a mechanism for social effects, SCN suprachiasmatic nucleus
masking of rhythms has been viewed—at least in
some chronobiological circles—almost pejora-
tively. For social interactions that ultimately confer References
survival and reproductive benefits to individuals,
there is beauty in elucidating diverse mechanisms Aschoff, J., & von Goetz, C. (1988). Masking of circadian
and convergent evolution in multiple species, with activity rhythms in male golden hamsters by the presence
of females. Behavioral Ecology and Sociobiology, 22,
differential roles for entrainment, masking, and 409–412.
other mechanisms not fully understood (Thornton Begall, S., Daan, S., Burda, H., & Overkamp, G. J. F. (2002).
et al., 2010). Inbred lines of model organisms may Activity patterns in a subterranean social rodent,
277

Spalacopus cyanus (Octodontidae). Journal of Mammalogy, social stimuli and a circadian pacemaker. Physiology and
83, 153–158. Behavior, 40, 583–590.
Blanchong, J. A., McElhinny, T. L., Mahoney, M. M., & Erkert, H. G., & Schardt, U. (1991). Social entrainment of cir-
Smale, L. (1999). Nocturnal and diurnal rhythms in the cadian activity rhythms in common marmosets, Callithrix j.
unstriped Nile rat, Arvicanthis niloticus. Journal of jacchus (Primates). Ethology, 87, 189–202.
Biological Rhythms, 14, 364–377. Favreau, A., Richard-Yris, M. A., Bertin, A., Houdelier, C., &
Blanchong, J. A., & Smale, L. (2000). Temporal patterns of Lumineau, S. (2009). Social influences on circadian behav-
activity of the unstriped Nile rat, Arvicanthis niloticus. Jour- ioural rhythms in vertebrates. Animal Behavior, 77,
nal of Mammalogy, 81, 595–599. 983–989.
Bloch, G. (2010). The social clock of the honeybee. Journal of Fenn, M. G. P., & Macdonald, D. W. (1995). Use of middens
Biological Rhythms, 25, 307–317. by red foxes—Risk reverses rhythms of rats. Journal of
Bloch, G., & Robinson, G. E. (2001). Chronobiology. Reversal Mammalogy, 76, 130–136.
of honeybee behavioural rhythms. Nature, 410, 1048. Flint, J. F., Drzymalski, D., Montgomery, W. L., Southam, G.,
Bovet, J. (1972). On the social behavior in a stable group of & Angert, E. R. (2005). Nocturnal production of endospores
long-tailed field mice (Apodemus sylvaticus). II. Its relations in natural populations of epulopiscium-like surgeonfish
with distribution of daily activity. Behaviour, 41, 55–67. symbionts. Journal of Bacteriology, 187, 7460–7470.
Bovet, J., & Oertli, E. F. (1974). Free-running circadian activ- Fluxman, S., & Haim, A. (1993). Daily rhythms of body tem-
ity rhythms in free-living beaver (Castor canadensis). Jour- perature in Acomys russatus: The response to chemical
nal of Comparative Physiology, 92, 1–10. signals released by Acomys cahirinus. Chronobiology Inter-
Caldelas, I., Gonzalez, B., Montufar-Chaveznava, R., & national, 10, 159–164.
Hudson, R. (2009). Endogenous clock gene expression in Foote, J. R., Fitzsimmons, L. P., MennillB, D. J., &
the suprachiasmatic nuclei of previsual newborn rabbits is Ratcliffe, L. M. (2011). Male black-capped chickadees begin
entrained by nursing. Developmental Neurobiology, 69, dawn chorusing earlier in response to simulated territorial
47–59. insertions. Animal Behavior, 81, 871–877.
Calisi, R. M., & Bentley, G. E. (2009). Lab and field Frisch, B., & Koeniger, N. (1994). Social synchronization of
experiments: Are they the same animal? Hormones and the activity rhythms of honeybees within a colony. Behav-
Behavior, 56, 1–10. ioral Ecology and Sociobiology, 35, 91–98.
Carroll, S. P., Hendry, A. P., Reznick, D. N., & Fox, C. W. Gattermann, R., Johnston, R. E., Yigit, N., Fritzsche, P.,
(2007). Evolution on ecological time-scales. Functional Ecol- Larimer, S., Ozkurt, S., et al. (2008). Golden hamsters are
ogy, 21, 387–393. nocturnal in captivity but diurnal in nature. Biology Letters,
Cohen, R. (2004). Circadian rhythms in Acomys russatus. 4, 253–255.
M.Sc. dissertation. Tel Aviv University. Gattermann, R., & Weinandy, R. (1997). Lack of social
Cohen, R., & Kronfeld-Schor, N. (2006). Individual variability entrainment of circadian activity rhythms in the solitary
and photic entrainment of circadian rhythms in golden spiny golden hamster and in the highly social Mongolian gerbil.
mice. Physiology and Behavior, 87, 563–574. Biological Rhythm Research, 28, 85–93.
Crowley, M., & Bovet, J. (1980). Social synchronization of cir- Ghalambor, C. K., McKay, J. K., Carroll, S. P., &
cadian rhythms in deer mice (Peromyscus maniculatus). Reznick, D. N. (2007). Adaptative versus non-adaptive phe-
Behavioral Ecology and Sociobiology, 7, 99–105. notypic plasticity and the potential for contemporary adap-
Davidson, A. J., & Menaker, M. (2003). Birds of a feather tation in new environments. Functional Ecology, 21,
clock together—Sometimes: Social synchronization of circa- 394–407.
dian rhythms. Current Opinion in Neurobiology, 13, Glass, G. E., & Slade, N. A. (1980). The effect of Sigmodon
765–769. hispidus on spatial and temporal activity of Microtus
Davis, F. C., & Gorski, R. A. (1988). Development of hamster ochrogaster: Evidence for competition. Ecology, 61,
circadian rhythms: Role of the maternal suprachiasmatic 358–370.
nucleus. Journal of Comparative Physiology. A, 162, Goel, N., & Lee, T. M. (1995). Sex differences and effects of
601–610. social cues on daily rhythms following phase advances in
Davis, F. C., & Reppert, S. M. (2001). Development of mam- Octodon degus. Physiology and Behavior, 58, 205–213.
malian circadian rhythms. In J. S. Takahashi, F. W. Turek Goel, N., & Lee, T. M. (1996). Relationship of Circadian activ-
& R. Y. Moore (Eds.), Circadian clocks. Handbook of ity and social behaviors to reentrainment rates in diurnal
behavioral neurobiology (Vol. 12), (pp. 247–290). New Octodon degus (Rodentia). Physiology and Behavior, 59,
York: Kluwer Academic/Plenum Publishers. 817–826.
Davis, F. C., Stice, S., & Menaker, M. (1987). Activity and Goel, N., & Lee, T. M. (1997). Social cues modulate free-
reproductive state in the hamster: Independent control by running circadian activity rhythms in the diurnal rodent,
278

Octodon degus. American Journal Physiology. Regulatory, diurnal rodent Octodon degus. Journal of Biological
Integrative and Comparative Physiology, 42, R797–R804. Rhythms, 18, 43–50.
Governale, M. M., & Lee, T. M. (2001). Olfactory cues Jilge, B., & Hudson, R. (2001). Diversity and development of
accelerate reentrainment following phase shifts and entrain circadian rhythms in the European rabbit. Chronobiology
free-running rhythms in female Octodon degus (Rodentia). International, 18, 1–26.
Journal of Biological Rhythms, 16, 489–501. Jud, C., & Albrecht, U. (2006). Circadian rhythms in murine
Gwinner, E. (1966). Periodicity of a circadian rhythm in birds pups develop in absence of a functional maternal circadian
by species-specific song cycles (Aves Fringillidae: Carduelis clock. Journal of Biological Rhythms, 21, 149–154.
spinus, Serinus serinus). Experientia, 22, 765–766. Katona, C., & Smale, L. (1997). Wheel-running rhythms in
Haim, A., & Rozenfeld, F. M. (1993). Temporal segregation in Arvicanthis niloticus. Physiology and Behavior, 61, 365–372.
coexisting Acomys species: The role of odour. Physiology Kavaliers, M. (1980a). Circadian activity of the white sucker,
and Behavior, 54, 1159–1161. Catostomus commersoni: Comparison of individual and
Hansen, E. A., & Closs, G. P. (2005). Diel activity and home shoaling fish. Canadian Journal of Zoology, 58, 1399–1403.
range size in relation to food supply in a drift-feeding stream Kavaliers, M. (1980b). Social groupings and circadian activity
fish. Behavioral Ecology, 16, 640–648. of the killifish, Fundulus heteroclitus. Biology Bulletin, 158,
Harrington, L. A., Harrington, A. L., Yamaguchi, N., 69–76.
Thom, M. D., Ferreras, P., Windham, T. R., et al. (2009). Kitchen, A. M., Gese, E. M., & Schauster, E. R. (2000).
The impact of native competitors on an alien invasive: Tem- Changes in coyote activity patterns due to reduced exposure
poral niche shifts to avoid interspecific aggression? Ecology, to human persecution. Canadian Journal of Zoology, 78,
90, 1207–1216. 853–857.
Herzog, E. D., Aton, S. J., Numano, R., Sakaki, Y., & Tei, H. Klerman, E. B., Rimmer, D. W., Dijk, D. J., Kronauer, R. E.,
(2004). Temporal precision in the mammalian circadian sys- Rizzo, J. F., 3rd, & Czeisler, C. A. (1998). Nonphotic
tem: A reliable clock from less reliable neurons. Journal of entrainment of the human circadian pacemaker. American
Biological Rhythms, 19, 35–46. Journal of Physiology, 274, R991–R996.
Hiroshige, T., Honma, K., & Watanabe, K. (1982). Possible Korslund, L. (2006). Activity of root voles (Microtus
zeitgebers for external entrainment of the circadian rhythm oeconomus) under snow: Social encounters synchronize
of plasma corticosterone in blind infantile rats. The Journal individual activity rhythms. Behavioral Ecology and Sociobi-
of Physiology, 325, 507–519. ology, 61, 255–263.
Honrado, G. I., & Mrosovsky, N. (1989). Arousal by sexual Krause, S., & Schuler, L. (2010). Behavioural and endocrino-
stimuli accelerates the re-entrainment of hamsters to phase logical changes in Syrian hamsters (Mesocricetus auratus)
advanced light-dark cycles. Behavioral Ecology and Sociobi- under domestication. Journal of Animal Breeding and
ology, 25, 57–63. Genetics, 127, 452–461.
Honrado, G. I., & Mrosovsky, N. (1991). Interaction between Kronfeld-Schor, N., & Dayan, T. (2003). Partitioning of time
periodic socio-sexual cues and light-dark cycles in controlling as an ecological resource. Annual Review Ecology Evolution
the phasing of activity rhythms in golden hamsters. Ethology and Systematics, 34, 153–181.
Ecology and Evolution, 3, 221–231. Kronfeld-Schor, N., & Dayan, T. (2008). Activity patterns of
Hudgens, B. R., & Garcelon, D. K. (2011). Induced changes in rodents: The physiological ecology of biological rhythms.
island fox (Urocyon littoralis) activity do not mitigate the Biological Rhythm Research, 39, 193–211.
extinction threat posed by a novel predator. Oecologia, Krupp, J. J., Kent, C., Billeter, J. C., Azanchi, R., So, A. K.,
165, 699–705. Schonfeld, J. A., et al. (2008). Social experience modifies
Hut, R. A., Kronfeld-Schor, N., van der Vinne, V., & De la pheromone expression and mating behavior in male
Iglesia, H. O. (2012). In search of a temporal niche: Environ- Drosophila melanogaster. Current Biology, 18, 1373–1383.
mental factors. Progress in Brain Research, 199, 281–304. Kuramoto, Y. (1984). Chemical oscillations, waves, and turbu-
Jechura, T., & Lee, T. M. (2004). Ovarian hormones influence lence. Mineola, New York: Dover Publications.
olfactory cue effects on reentrainment in the diurnal rodent, Leise, T. L., & Harrington, M. E. (2011). Wavelet-based time
Octodon degus. Hormones and Behavior, 46, 349–355. series analysis of circadian rhythms. Journal of Biological
Jechura, T. J., Mahoney, M. M., Stimpson, C. D., & Lee, T. M. Rhythms, 26, 454–463.
(2006). Odor-specific effects on reentrainment following Levine, J. D., Funes, P., Dowse, H. B., & Hall, J. C. (2002).
phase advances in the diurnal rodent, Octodon degus. Amer- Resetting the circadian clock by social experience in
ican Journal of Physiology. Regulatory, Integrative and Drosophila melanogaster. Science, 298, 2010–2012.
Comparative Physiology, 291, R1808–R1816. Levy, O., Dayan, T., & Kronfeld-Schor, N. (2007). The rela-
Jechura, T. J., Walsh, J. M., & Lee, T. M. (2003). Testosterone tionship between the golden spiny mouse circadian system
suppresses circadian responsiveness to social cues in the and its diurnal activity: An experimental field enclosures
279

and laboratory study. Chronobiology International, 24, rhythms in malaria parasites. Proceedings of the Royal Soci-
599–613. ety B: Biological Sciences, 278, 2429–2436.
Lima-Camara, T. N., Bruno, R. V., Luz, P. M., Castro, M. G., Ohta, H., Honma, S., Abe, H., & Honma, K. (2003). Periodic
Lourenco-de-Oliveira, R., Sorgine, M. H., et al. (2011). absence of nursing mothers phase-shifts circadian rhythms
Dengue infection increases the locomotor activity of Aedes of clock genes in the suprachiasmatic nucleus of rat pups.
aegypti females. PLoS One, 6, e17690. European Journal of Neuroscience, 17, 1628–1634.
Liu, W.-C. (2004). The effect of neighbours and females on Packer, C. (1983). Demographic changes in a colony of Nile
dawn and daytime singing behaviours by male chipping grass rats (Arvicanthis niloticus) in Tanzania. Journal of
sparrows. Animal Behavior, 68, 39–44. Mammalogy, 64, 159–161.
Liu, A. C., Welsh, D. K., Ko, C. H., Tran, H. G., Zhang, E. E., Paul, M. J., & Schwartz, W. J. (2007). On the chronobiology of
Priest, A. A., et al. (2007). Intercellular coupling confers cohabitation. Cold Spring Harbor Symposia on Quantitative
robustness against mutations in the SCN circadian clock net- Biology, 72, 615–621.
work. Cell, 129, 605–616. Rajaratnam, S. M., & Redman, J. R. (1999). Social contact
Lu, D. B., Wang, T. P., Rudge, J. W., Donnelly, C. A., synchronizes free-running activity rhythms of diurnal palm
Fang, G. R., & Webster, J. P. (2009). Evolution in a multi- squirrels. Physiology and Behavior, 66, 21–26.
host parasite: Chronobiological circadian rhythm and popu- Reebs, S. G. (1989). Acoustical entrainment of circadian activ-
lation genetics of Schistosoma japonicum cercariae indicates ity rhythms in house sparrows: Constant light is not neces-
contrasting definitive host reservoirs by habitat. Interna- sary. Ethology, 80, 172–181.
tional Journal of Parasitology, 39, 1581–1588. Refinetti, R., Nelson, D. E., & Menaker, M. (1992). Social
Marcelli, M., Fusillo, R., & Boitani, L. (2003). Sexual segrega- stimuli fail to act as entraining agents of circadian rhythms
tion in the activity patterns of European polecats (Mustela in the golden hamster. Journal of Comparative Physiology.
putorius). Journal of Zoology London, 261, 249–255. A, 170, 181–187.
Marimuthu, G., Rajan, S., & Chandrashekaran, M. K. (1981). Regal, P. J., & Connolly, M. S. (1980). Social influences on
Social entrainment of the circadian rhythm in the flight biological rhythms. Behaviour, 72, 171–199.
activity of the microchiropteran bat Hipposideros speoris. Reppert, S. M., Coleman, R. J., Heath, H. W., &
Behavioral Ecology and Sociobiology, 8, 147–150. Swedlow, J. R. (1984). Pineal N-acetyltransferase activity
McElhinny, T. L., Smale, L., & Holekamp, K. E. (1997). in 10-day-old rats: A paradigm for studying the developing
Patterns of body temperature, activity, and reproductive circadian system. Endocrinology, 115, 918–925.
behavior in a tropical murid rodent, Arvicanthis niloticus. Reppert, S. M., & Schwartz, W. J. (1986). Maternal
Physiology and Behavior, 62, 91–96. suprachiasmatic nuclei are necessary for maternal coordina-
Meerlo, P., Sgoifo, A., De Boer, S. F., & Koolhaas, J. M. tion of the developing circadian system. Journal of Neurosci-
(1999). Long-lasting consequences of a social conflict in rats: ence, 6, 2724–2729.
Behavior during the interaction predicts subsequent changes Rivkees, S. A., & Reppert, S. M. (1990). Entrainment of circa-
in daily rhythms of heart rate, temperature, and activity. dian phase in developing gray short-tailed opossums:
Behavioral Neuroscience, 113, 1283–1290. Mother vs. environment. American Journal of Physiology,
Menaker, M., & Eskin, A. (1966). Entrainment of circadian 259, E384–E388.
rhythms by sound in Passer domesticus. Science, 154, Scheibler, E., & Wollnik, F. (2009). Interspecific contact
1579–1581. affects phase response and activity in Desert hamsters.
Mistlberger, R. E., Antle, M. C., Webb, I. C., Jones, M., Physiology and Behavior, 98, 288–295.
Weinberg, J., & Pollock, M. S. (2003). Circadian clock Schoener, T. W. (1986). Resource partitioning. In J. Kikkawa
resetting by arousal in Syrian hamsters: The role of stress & D. J. Anderson (Eds.), Community ecology: Pattern and
and activity. American Journal of Physiology. Regulatory, process (pp. 91–126). Blackwell Scientific Publications:
Integrative and Comparative Physiology, 285, R917–R925. Melbourne.
Mistlberger, R. E., & Skene, D. J. (2004). Social influences on Senzota, R. (1990). Activity patterns and social behaviour of
mammalian circadian rhythms: Animal and human studies. the Grass rats [Arvicanthis niloticus (Desmarest)] in the
Biological Reviews, 79, 533–556. Serengeti National Park, Tanzania. Tropical Ecology, 31,
Moore, D. (2001). Honey bee circadian clocks: Behavioral 35–40.
control from individual workers to whole-colony rhythms. Shemesh, Y., Eban-Rothschild, A., Cohen, M., & Bloch, G.
Journal of Insect Physiology, 47, 843–857. (2010). Molecular dynamics and social regulation of
Mrosovsky, N. (1988). Phase response curves for social entrain- context-dependent plasticity in the circadian clockwork
ment. Journal of Comparative Physiology. A, 162, 35–46. of the honey bee. Journal of Neuroscience, 30, 12517–12525.
O’Donnell, A. J., Schneider, P., McWatters, H. G., & Shkolnik, A. (1971). Diurnal activity in a small desert rodent.
Reece, S. E. (2011). Fitness costs of disrupting circadian International Journal of Biometeorology, 15, 115–120.
280

Smale, L., Heideman, P. D., & French, J. A. (2005). Behav- Weaver, D. R., & Reppert, S. M. (1987). Maternal-fetal commu-
ioral neuroendocrinology in nontraditional species of nication of circadian phase in a precocious rodent, the spiny
mammals: Things the ‘knockout’ mouse CAN’T tell us. Hor- mouse. American Journal of Physiology, 253, E401–E409.
mones and Behavior, 48, 474–483. Webb, A. B., Angelo, N., Huettner, J. E., & Herzog, E. D.
Swarts, H. M., Crooks, K. R., Willits, N., & Woodroffe, R. (2009). Intrinsic, nondeterministic circadian rhythm genera-
(2009). Possible contemporary evolution in an endangered tion in identified mammalian neurons. Proceedings of
species, the Santa Cruz Island fox. Animal Conservation, the National Academy of Sciences of the United States
12, 120–127. of America, 106, 16493–16498.
Theron, A., & Combes, C. (1988). Genetic analysis of cercarial Wier, A. M., Nyholm, S. V., Mandel, M. J., Massengo-
emergence rhythms of Schistosoma mansoni. Behavior Tiasse, R. P., Schaefer, A. L., Koroleva, I., et al. (2010).
Genetics, 18, 201–209. Transcriptional patterns in both host and bacterium underlie
Thornton, A., Samson, J., & Clutton-Brock, T. (2010). Multi- a daily rhythm of anatomical and metabolic change in a ben-
generational persistence of traditions in neighbouring meer- eficial symbiosis. Proceedings of the National Academy of
kat groups. Proceedings of the Royal Society B: Biological Sciences of the United States of America, 107, 2259–2264.
Sciences, 277, 3623–3629. Winfree, A. T. (2001). The geometry of biological time. New
Toledo, R., Munoz-Antoli, C., & Esteban, J. G. (1999). Produc- York: Springer-Verlag.
tion and chronobiology of emergence of the cercariae of Ziv, Y., Abramsky, Z., Kotler, B. P., & Subach, A. (1993).
Euparyphium albuferensis (Trematoda: Echinostomatidae). Interference competition and temporal and habitat
Journal of Parasitology, 85, 263–267. partitioning in two gerbil species. Oikos, 66, 237–246.
Urrejola, D., Lacey, E. A., Wieczorek, J. R., & Zschille, J., Stier, N., & Roth, M. (2010). Gender differences in
Ebensperger, L. A. (2005). Daily activity patterns of free- activity patterns of American mink Neovison vison in
living cururos (Spalacopus cyanus). Journal of Mammalogy, Germany. European Journal of Wildlife Research, 56,
86, 302–308. 187–194.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 17

In search of a temporal niche: Environmental


factors

Roelof A. Hut{,*, Noga Kronfeld-Schor{, Vincent van der Vinne{ and


Horacio De la Iglesia}

{
Chronobiology unit, Center for Behaviour and Neurosciences, University of Groningen, Groningen, The Netherlands
{
Department of Zoology, Tel Aviv University, Tel Aviv, Israel
}
Department of Biology and Program of Neurobiology and Behavior, University of Washington, Seattle, USA

Abstract: Time as an ecological niche variable or “temporal niche” can be defined in the context of the
most prominent environmental cycles, including the tidal cycle, the lunar day and month, the solar day,
and the earth year. For the current review, we focus on the 24-h domain generated through the earth’s
rotation around its axis (solar day). The daily environmental cycles of light and temperature are a
dominant ecological factor generating a variety of adaptations among animals. In this review, we
describe these adaptations with a special focus on the visual system and on the adaptive plasticity of
activity patterns. Our goals are: (1) Underscore the importance of the 24-h time axis as critical
variable in the ecological niche. (2) Highlight cases of temporal niche switches at the evolutionary
timescale (phylogenetic level). (3) Review temporal niche switching within an individual’s lifespan. (4)
Evaluate possible underlying mechanisms for temporal niche switching. (5) Describe a new hypothesis
of circadian thermoenergetics which may explain several cases of temporal niche switching in
mammals. With this, we hope to inspire experiments under natural conditions or more complex
laboratory environments, aimed to reveal environmental factors and mechanisms underlying specific
temporal programs.

Keywords: temporal niche switching; circadian thermoenergetics hypothesis; nocturnality; diurnality; eye
properties; visual system; environment; evolution.

Introduction

*Corresponding author. More than half a century after G. Evelyn


Tel.: þ31503632060; Fax: þ 31503632148 Hutchinson’s crisp definition of ecological niche
E-mail: r.a.hut@rug.nl as a multidimensional space delineated by the

http://dx.doi.org/10.1016/B978-0-444-59427-3.00017-4 281
282

range of resources in which a species survives nocturnal or diurnal phenotypes. However, the
and reproduces (Hutchinson, 1957), “time” still temporal programs under natural conditions,
struggles to make it into the list of critical niche where environmental factors are more variable
orthogonal axes (Kronfeld-Schor and Dayan, and the behavioral repertoire of an animal has a
2003, 2008). Those of us studying how time wider range, are likely more complex and more
shapes physiology and behavior take the inclusion plastic. A classic example of this complexity is
of time in this multidimensional space for granted; the locomotor-activity patterns migratory birds
after all, it is hard to imagine the survival of Arctic display in the laboratory. Most birds are diurnal
birds hatching just before the winter solstice or and display behaviors such as foraging during
honeybees foraging at midnight. Yet, we feel that the daytime; under laboratory conditions, this is
studies that have addressed which environmental usually manifested by increased locomotor activ-
factors and mechanisms shape this temporal niche ity during the light phase. During the migratory
deserve a closer look. season, birds migrate during the night; under the
The daily temporal niche of a species is usually proper photoperiod, this migratory activity is
defined as the time of day at which its individuals manifested in the laboratory as nocturnal restless-
display locomotor activity. Thus, diurnality and ness (Bartell and Gwinner, 2005; Coverdill et al.,
nocturnality are defined as temporal patterns with 2008; Gwinner, 1996). This nocturnal-locomotor
activity, respectively, during daytime and night activity can be distinguished from the activity
time. Similarly, crepuscularity is characterized by birds display during the light phase by an experi-
activity around dawn and dusk and cathemerality enced observer. However, an automatic locomo-
by activity both during the daytime and nighttime. tor-activity detection system would only portray
These definitions have been biased to a great a bimodal pattern of activity but would provide
extent by the methodology to measure locomotor no information about the respective behaviors
activity. For instance, measuring running-wheel each activity about represents.
activity may portray a different locomotor-activity The temporal distribution of behavior can also
pattern than patterns emerging from infrared-beam shift under different environmental conditions.
interruptions. The presence of a running-wheel per In fact, more and more examples of “temporal niche
se can even cause a shift in temporal distribution of switches,” in which individuals of a species can
locomotor activity as it is the case in Nile rats switch from one temporal niche (e.g., nocturnality)
(Blanchong et al., 1999), degus (Kas and Edgar, to another (e.g., diurnality) are described both in
1999), Mongolian gerbils (Weinert et al., 2007), nature and under laboratory conditions (Table 1;
and golden spiny mice (Table 1; Cohen et al., Mrosovsky and Hattar, 2005). Similarly, temporal
2009). Importantly, locomotor activity displayed by niche switches occurred across phylogenetic bra-
an animal under laboratory conditions may— nches of closely related groups. Moreover, within
regardless of the measurement method—represent many taxa, these switches occur independently in
the expression of different behavioral programs different phylogenetic branches (e.g., Roll et al.,
under natural conditions (Fig. 1). In other words, 2006), indicating considerable changes in selective
wheel running may reflect foraging behavior, pressure on temporal niche phenotypes and conse-
predator avoidance, mate searching, or even escape quently considerable adaptive plasticity in this trait.
behavior. Thus, it is important to define the con-
ditions under which locomotor activity is measured
and the methodology. Temporal niche switching across phylogeny
Circadian studies under narrow laboratory
conditions have led to a dichotomous view of Evolutionary plasticity is limited, to a certain
temporal niche in which animals display either extent, by phylogenetic constraints. However,
283

Table 1. Examples of temporal niche switching in different species

Species TNS Cause Con References

House mouse Mus musculus N>D Natural conditions F Daan et al. (2011)
House mouse Mus musculus N>D Reduced food intake L Hut et al. (2011)
House mouse Mus musculus N>D Photoreceptor ablation L Mrosovsky and Hattar (2005),
Doyle et al. (2008)
House mouse Mus musculus N>D Subthreshold light L Doyle et al. (2008)
House mouse Mus domesticus N>D Metabolic demand L Perrigo (1987, 1990)
Golden spiny Acomys russatus N>D Interspecies competition F Shkolnik (1971)
mouse
Golden spiny Acomys russatus D>N Lab conditions F/L Levy et al. (2007)
mouse
Common vole Microtus arvalis N > Da Winter F Erkinaro (1969)
Common vole Microtus arvalis D > Ca Spring F Lehmann and Sommersberg
(1980)
Field vole Microtus agrestis N > Da Winter F Erkinaro (1969), Bäumler
(1975)
Field vole Microtus agrestis N>A Cage size reduction L Lehmann (1976)
Montane vole Microtus montanus N>D Short photoperiod L Rowsemitt et al. (1982)
Tundra vole Microtus N > Da Winter F Erkinaro (1969)
oeconomus
Syrian hamster Mesocricetus N > C, D Natural conditions F/L Gattermann et al. (2008)
auratus
Desert hamster Phodopus N > Nb Interspecies competition F Scheibler and Wollnik (2011)
roborovskii
Mongolian gerbil Meriones D>N Running wheel L Weinert et al. (2007)
unguiculatus
Kangaroo rat Dipodomys N>D Drought condition F Lockard (1978)
spectabilis
Ord’s kangaroo Dipodomys ordii N>D Drought condition F Boal and Giovanni (2007)
rat
Rat Rattus norvegicus N > N, Dc Lactation L Strubbe and Gorissen (1980)
Rat Rattus norvegicus N>D Natural conditions F/L Fenn and Macdonald (1995)
Rat Rattus norvegicus N>D Predation F Fenn and Macdonald (1995)
Rat Rattus norvegicus N>D Hypothalamic lesion L Richter (1978)
Rat Rattus norvegicus; A>N Development L Honma and Hiroshige (1977)
pup
N. American Castor canadensis N > FR Winter F Potvin and Bovet (1975),
beaver Bovet and Oertli (1992)
Rabbit Oryctolagus N>D Predation F Bakker et al. (2005)
cuniculus
Degu Octodon degus D, C > N, C Lab conditions, running F/L Fulk (1976), Bacigalupe et al.
wheel (2003), Garcia-Allegue et al.
(1999), Kas and Edgar (1999)
Degu Octodon degus D>N High temperature L Vivanco et al. (2010a),
Hagenauer and Lee (2008),
Refinetti (2006, 2008)
Coruro Spalacopus cyanus D>N Lab conditions F/L Ocampo-Garcès et al. (2006),
Begall et al. (2002)
Tuco tuco Ctenomys cf. D>N Lab conditions F/L Tomotani et al. (2011)
knighti

(Continued)
284

Table 1. Examples of temporal niche switching in different species (Continued)

Species TNS Cause Con References

Blind mole rat Spalax ehrenbergi D>N High temperature, F Kushnirov et al. (1998), Oster
summer et al. (2002)
Reindeer Rangifer taradus D>U Loss of LD cycle F Van Oort et al. (2007)
Roe deer Capreolus capreolus C, N > C, D Winter F Bubenik (1960) (in Aschoff,
1962; Tscherkowitsch, 1953)
Coyote Canis latrans N, D > Nd Hunting F Kitchen et al. (2000)
Bat-eared fox Otocyon megalotis D>N High temperature F Lourens and Nel (1990)
Geoffroy’s cats Leopardusgeoffroyi N>D Food shortage F Pereira (2010)
Owl monkey Aotus azarai C>N Moonlight F Fernandez-Duque et al. (2010)
Rhesus macaque Macaca mulatta A>D Development L Tscherkowicz (1953)
(in Aschoff, 1962)
Human Homo sapiens; A>D Development F Kleitman and Engelmann
infant (1953)
Whiskered bat Myotis mystacinus N>D Spring, autumn F Nyholm (1965)
Vampire bat Desmodus rotundus LN <> EN Moonlight avoidance F Turner (1975)
Leaf nosed bat Artibeus lituratus LN <> EN Moonlight avoidance F&L Erkert (1974), Häussler and
Erkert (1978)
Fat-tailed dunnart Sminthopsis N>D Winter F/L Warnecke et al. (2008),
crassicaudata Morton (1995), M.P. Gerkema
(personal communication)
Green toad Bufo viridis N>D Rainfall F R.A. Hut (personal
observation)
Cane toad, Bufo marinus D, N > De Cannibalism L Pizzatto et al. (2008)
metamorphs
Atlantic salmon Salmo salar A>N Low temperature F&L Fraser et al. (1995), Eriksson
(1973)
Atlantic salmon Salmo salar N>D Reduced food L Metcalfe et al. (1999)
Atlantic salmon Salmo salar N>D Growth demand L Metcalfe et al. (1998)
Atlantic salmon Salmo salar D>C>N Winter > summer L Eriksson (1973)
Brown trout Salmo trutta trutta D>C>N Winter > summer L Eriksson (1973)
Golden-lined Siganus lineatus D>N Shore > reef F Fox and Bellwood (2011)
rabbit fish
Mediterranean Nephrops N>D Deep water (>200 m) F Chiesa et al. (2010)
lobster norvegicus
Fruitfly Drosophila C>D Short photoperiod, low L Dubruille and Emery (2008)
melanogaster temperature
Fruitfly Drosophila C, D > C, N Moonlight L Kempinger et al. (2009)
melanogaster
Midge Chironomus N>D Low temperature F&L Kureck (1979)
thummi
Seed-harvesting Pheidole spp. D > N, C High temperature F Whitford et al. (1981)
ants

TNS, temporal niche switch; Con, condition; F, field study; L, lab study; F/L, comparison between field and lab study; F&L, effect observed in field
and lab. Switch in the dominant temporal niche indicated as N, nocturnal; D, diurnal; C, crepuscular; U, ultradian rhythms with no diurnal/nocturnal
preference; A, arrhythmic; FR, free running circadian rhythm; LN, late night active; EN, early night active.
a
Strong ultradian organization of activity always in either temporal niche preference.
b
Earlier-nocturnal activity.
c
Increased-diurnal feeding.
d
Reduced-diurnal activity.
e
Reduced-nocturnal activity.
285

Master Peripheral
oscillator oscillators
Nocturnal care
of offspring
LD
Diurnal
Temperature foraging

Crepuscular
Food mating

X Arrhythmic
predator
avoidance

Fig. 1. Complexity of temporal distribution of behaviors. A complex scenario of behavioral programs is likely to be present in
natural habitats. Multiple environmental cycles can act as Zeitgebers to entrain a master circadian pacemaker that in turn
synchronizes peripheral oscillators. Specific Zeitgebers could also directly entrain peripheral oscillators and bypass the control by
the master circadian pacemaker (dotted lines). Peripheral oscillators regulate specific physiological and behavioral outputs that
could each have a different temporal niche.

studies focusing on phylogeny as an evolutionary The main arguments in favor of the nocturnal
constraint limiting temporal niche switches are bottleneck hypothesis are threefold: (1) most cur-
scarce. Here, we focus on mammalian taxa for rent mammalian species seem to be nocturnal in
which a phylogenetic framework has been used to their behavior (Roll et al., 2006), (2) mammals
map temporal niches. have developed endothermia in order to occupy
the nocturnal niche, and (3) most mammals have
a rod dominated visual system (Table 2) and sev-
The mammalian nocturnal bottleneck eral cone types (except S and M cones) were lost
in mammals, indicating specialization toward a
The nocturnal bottleneck hypothesis (Crompton nocturnal life style.
et al., 1978; Foster and Menaker, 1993; Heesy The mass extinction of dinosaurs at the end of
and Hall, 2010; Young, 1962) proposes that all the Mesozoicum (65.5 mya; Cretaceous–Paleogene
mammals share a common nocturnal ancestor. extinction event or K/Pg boundary, also called the
Endothermic mammals evolved from ectothermic Cretaceous–Tertiary extinction event or K–T
reptiles during the Mesozoicum between 205 and boundary) opened the way for mammalian species
65.5 mya. The first mammal-like species are to diversify their temporal niches. Diurnality and
thought to have evolved endothermia in order to other forms of temporal niche exploitation
occupy the nocturnal temporal niche to avoid emerged through evolutionary radiation and all
ectothermic predators and competitors that were known temporal niche patterns in mammals were
tied to the diurnal temporal niche because of their probably present by the end of the Eocene (37
need for behavioral body temperature regulation mya) when most modern mammalian taxa had
through ambient temperature (Crompton et al., evolved.
1978). The term “bottleneck” indicates that all Several arguments against the nocturnal bottle-
mammalian taxa decent from species that faced a neck hypothesis can be raised. First, ectothermia
common selective pressure favoring nocturnality. is not exclusively linked to diurnal behavior.
286

Table 2. Dominant temporal niche and cone photoreceptor percentages

Species DTN Cone (%) References

European ground squirrel Spermophilus citellus D 97 Szél and Röhlich (1988)


Mexican ground squirrel Ictidomys mexicanus D 95–96 West and Dowling (1975)
13-lined ground squirrel Ictidomys D 95–96 West and Dowling (1975)
tridecemlineatus
Tupaia Tupaia belangeri D 95 Müller and Peichl (1989)
Black-tailed prairie dog Cynomys ludovicianus D 90 West and Dowling (1975)
California ground squirrel Otospermophilus D 85 Kryger et al. (1998)
beecheyi
Eastern gray squirrel Sciurus carolinensis D 60 West and Dowling (1975)
Nile grass rat Arvicanthis niloticus D, C 35–40 Gaillard et al. (2008)
Degu Octodon degus D, N 32 Jacobs et al. (2003)
Plains pocket gopher Geomys bursarius N, D 25 Feldman and Phillips (1984), Benedix (1994)
Talas tuco tuco Ctenomys talarum D, N 14–31 Schleich et al.– 2010
Magellanic tuco tuco Ctenomys magellanicus D, N 10–31 Schleich et al. (2010)
Agouti Dasyprocta aguti D, N 10–20 Rocha et al. (2009)
Pig Sus scrofa domestica D, N, C 10–20 Hendrickson and Hicks (2002)
Guinea pig Cavia porcellus C, A 8–17 Peichl and González-Soriano (1994)
Mongolian gerbil Meriones unguiculatus N, D 13 Govardovskii et al. (1992)
European mole Talpa europaea A, U 10–12 Glösmann et al. (2008)
Ansell’s mole rat Cryptomys anselli N, D, A 10 Peichl et al. (2003)
Giant mole rat Cryptomys mechowi N, D, A 10 Peichl et al. (2003)
Coruro Spalacopus cyanus N, D 10 Peichl et al. (2005)
Common shrew Sorex araneus A, U 6–13 Peichl et al. (2000), Crowford (1953)
Greater white-toothed Crocidura russula N 5–6 Peichl et al. (2000), Crowford (1953)
shrew
Human Homo sapiens D 4.7 Curcio et al. (1990)
Fraser’s musk shrew Crocidura poensis N 4–5 Peichl et al. (2000)
Rabbit Oryctolagus cuniculus N, D, C 4 Peichl (1997)
Mouse Mus musculus N 3 Carter-Dawson and LaVail (1979)
Lesser hedgehog, tenrec Echinops telfairi N 1.5–2.3 Peichl et al. (2000)
Rat Rattus norvegicus N 1 LaVail (1976), Szél and Röhlich (1992)
Eastern woodrat Neotoma floridana N 1 Feldman and Phillips (1984)
Mouse opossum Thylamys elegans N 0.4–1.2 Palacios et al. (2010)

Cones are functional at higher light intensities than rods and therefore thought to be indicative of diurnal vision. Consequently, cone densities, as per-
centage of the total number of photoreceptors in the whole retina, may indicate the dominant temporal niche of a species. DTN, dominant temporal
niche; D, diurnal; N, nocturnal; C, crepuscular; A, arrhythmic; U, ultradian.

Geckos, for instance, are small ectothermic reptiles turbinate structures (Nespolo et al., 2011). Even
that hunt for prey during the night. Second, it is without endothermia, large dinosaurs may have
debatable whether all dinosaurs were ectotherms. been able to retain enough body heat to support
The arguments indicating that dinosaurs during activity during the night through “inertial hom-
the Mesozoicum may have evolved forms of endo- eothermy” or “mass homeothermy.”. Third, intra-
thermic capacity include the discovery of insulating ocular and orbital bone sizes indicate that several
feather structures, fossils of breeding dinosaurs in dinosaur species—including relatively small, feath-
arctic regions suggesting heat transfer from adults ered predatory Theropoda like Velociraptor—
to offspring, raised body posture, fibrolamellar seemed to have had eye shapes that facilitate
skeletal bone structure, and the presence of nasal nocturnal vision (Schmitz and Motani, 2011).
287

Taken together, it seems probable that the extreme diurnal phenotype are flying squirrels, an
nocturnal bottleneck indeed has occurred in early example of strictly nocturnal Sciuridae (DeCoursey,
mammals. Recent evidence, however, seems to 1960, 1986).
indicate that potential predators of these early
mammals (like some Therapoda species) may have
responded to this novel invention of mammalian Primates
nocturnal activity by developing endothermic cap-
acities, enabling these dinosaurs to also exploit the The temporal niche of primates ranges from
nocturnal niche (as indicated by changes in eye diurnality to nocturnality with many species showing
morphology). This leads to the interesting conclu- “cathemeral” patterns of activity. Although the
sion that, already during the Mesozoicum (i.e., term cathemeral has been widely accepted in the
before the great extinction of the dinosaurs), the primatology literature, a careful analysis of so-called
mammalian nocturnal bottleneck triggered an evo- cathemeral patterns of activity indicates that these
lutionary arms race between predator and prey for patterns are crepuscular, with varying degrees of
temporal niche occupation. This in turn may have nocturnality and diurnality, which is in many cases
led to considerable adaptive plasticity in temporal determined by environmental factors (see Lumi-
niche usage among the earliest mammals. nance and Temporal niche switching). Strepsirhine
primates (lemurs and lorisiforms) are predomi-
nantly nocturnal. Haplorhine primates (tarsiers,
Temporal niche switching in mammals monkeys, and apes), in contrast, are predominantly
diurnal, with two genera representing the nocturnal
Rodents exception: Tarsius (tarsiers) and Aotus (owl
monkeys). Thus, owl monkeys represent a remark-
A literature survey of activity patterns of 700 able exception among anthropoid primates (Old
rodent species found that activity patterns at World monkeys, New World monkeys, apes, and
the family level are significantly different from humans), as the only nocturnal genus. Because
the expected random patterns, emphasizing the strepsirhine primates and tarsiers are typically noc-
strong relationship between intrafamily taxo- turnal and both groups show more primitive
nomic affiliation and daily-activity patterns. Large features than anthropoid primates, it is traditionally
families (Muridae and Sciuridae) analyzed by believed that the ancestral Eocene primate was
subfamilies and tribes showed a similar but stron- probably nocturnal (Martin and Ross, 2005). This
ger pattern than that of the family level. The notion, however, has been challenged by molecular
researchers concluded that phylogeny constrains evolution studies of the S-opsin, which usually
the evolution of activity patterns in rodents and shows relaxed selection in nocturnal species, of
may limit their ability to use the time niche axis several tarsiers and lemurs (Tan et al., 2005).
for ecological separation (Roll et al., 2006). Nev- The authors show great variability in the degree
ertheless, there are some striking examples of of relaxed selection among nocturnal species,
temporal niche dichotomy within relatively small suggesting that nocturnality has emerged at differ-
taxa. For instance, ground squirrels form a species ent times and multiple times among nocturnal
group which is typically exclusively diurnal. Euro- primates, and that the ancestral primate may have
pean ground squirrels (Spermophilus citellus), been diurnal or cathemeral. Regardless of the tem-
unlike diurnal birds, are not even active around poral niche of the ancestral primate, it is likely that
dawn and dusk, but appear above ground  3 h after nocturnality–diurnality switches, and diurnality–
sunrise and retreat below ground  3 h before sun- nocturnality switches have emerged indepen-
set (Hut et al., 1999). In stark contrast to this dently and multiple times throughout primate
288

evolution (Ankel-Simons and Rasmussen, 2008), nocturnal emergence from their burrows and
and this is consistent with the multiple mechanisms foraging during the night when they live in the
that could potentially lead to niche switches. shallow sea shelf (10–50 m deep), they have a cre-
puscular phenotype on the lower shelf (50–200 m
deep), and they are diurnal in the deeper slope
Temporal niche switching within species
(200–400 m deep; Table 1). Whereas fishermen
trolling for this commercially important species
Changes in the daily temporal niche have been
had long recognized this depth-dependent temporal
reported in a wide range of animals. These changes
niche switch (reviewed in Aguzzi and Sarda, 2008),
can be associated with environmental cycles such as
a recent study has shown that light–dark (LD) cycles
seasons and lunar month, with development and
that, respectively, mimic the brighter blue light of
aging, and with changing ecological factors such as
the shallow waters and the dimmer blue light
the presence of competitors, prey, or predators. The
of the deeper slope are sufficient to determine the
golden spiny mouse (Acomys russatus) seems to be
nocturnal and diurnal phenotypes (Chiesa et al.,
forced to the diurnal niche under natural conditions,
2010). Although similar experiments have not been
possibly through resource competition with the com-
done in other species with wide ocean-depth ranges,
mon spiny mouse (Acomys cahirinus). When golden
it is likely that the dramatic changes in light intensity
spiny mice are placed under laboratory conditions,
and wavelength across water depths represent a key
they immediately switch back to the nocturnal niche
determinant of temporal niche in other species that
(Fig. 2; Levy et al., 2007). Many such cases of tempo-
inhabit the continental margin.
ral niche switching are briefly summarized in Table 1.
Nocturnal species depending on vision may use
Here, we limit our discussion to environmental
moonlight to increase their foraging success.
determinants of niche switches, both in laboratory
In fact, most nocturnal mammals rely on moon-
and in field settings. Social determinants of niche
light, and nocturnal dim light stimulates their loco-
switches, such as inter- and intraspecific interactions
motor activity. In these species, nocturnal-activity
are reviewed by Castillo-Ruiz et al. (2012).
patterns can often be synchronized with moon
phases (Erkert, 2008). In the southernmost owl
Luminance and temporal niche switching monkey (Aotus azarai), this moonlight depen-
dence leads to predictable temporal niche switches
The Mediterranean lobster (Nephrops norvegicus) with striking nocturnal activity during moonlit
represents a unique case of diurnal–nocturnal nights and crepuscular activity during new moon
temporal niche switching. These animals show nights (Table 1, Fernandez-Duque et al., 2010).
Field
Field

Lab (DD)
Lab (DD)

5:00 18:30 5:00 18:30 5:00 18:30 5:00 18:30

Fig. 2. Temporal niche switching in golden spiny mice: field versus laboratory. Two examples of temporal niche switching in golden
spiny mice. When the animals were taken from the field to the laboratory under continuous dim light to assess the phase of their
activity rhythm, they show a clear switch from diurnal activity in the field to nocturnal activity in the lab.
289

In contrast to the stimulatory effects of moonlight (Otocyon megalotis) of the South African deserts
on locomotor activity in nocturnal primates and forages during the night in the summer when mid-
rodents, bats, which rely more on echolocation day soil temperatures reach 70 C and during the
than on vision, seem to avoid the moonlit part day in the winter when night air temperatures
of the night (Erkert, 1974; Häussler and Erkert, drop to 10 C (Lourens and Nel, 1990). Desert
1978; Turner, 1975). This behavior could in- seed-harvesting ants forage during the day in
crease their foraging success by timing hunting winter but avoid the heat of summer by foraging
when their prey exhibits minimal avoidance crepuscularly, nocturnally, or on cloudy days
behavior and also by reducing the risk of being (Whitford et al., 1981). Several rodent species
predated upon. switch activity patterns between seasons, appar-
ently for thermoregulatory reasons. For example,
montane voles (Microtus montanus) become
more diurnal during the winter (Rowsemitt
Ambient temperature and temporal niche et al., 1982). Similarly, the subterranean mole rat
switching (Spalax ehrenbergi) is predominantly diurnal dur-
ing winter but predominantly nocturnal during
In general, temperature differences in the terres- summer (Kushnirov et al., 1998; Oster et al.,
trial environment are much higher than in the 2002), a behavioral switch which seems to be
buffered aquatic environment. Many ectothermic regulated at the molecular light input pathway
species may need the high daytime temperatures of the core clock mechanism (Oster et al., 2002).
to augment their body temperature in order to Also, degus that are maintained at cooler
facilitate locomotor activity. In doing so, they face temperatures (e.g., 18  C) with food ad libitum,
the potential problem of increased evaporative demonstrate diurnal rhythms in temperature and
water loss. These effects of temperature can locomotor activity, whereas degus that are
explain why amphibians (with high evaporative maintained under the same conditions but at
water loss through their skin) in general choose more standard animal room temperatures
the nocturnal niche for terrestrial locomotion, (21–23  C) shift to more crepuscular or noc-
while reptiles (with reduced evaporative water turnal patterns (Hagenauer and Lee, 2008).
loss through their scaled skin) choose the diurnal Interestingly, in the field, there is no evidence of
niche and even start their active phase with thor- degus ever leaving their shelters at night. These
ough sun basking in the morning. The terrestrial data suggest that apart from thermoregulatory
green toad (Bufo viridis) escapes daytime high costs associated with ambient temperature,
temperatures on the Hungarian puszta plains by animals also integrate other variables like risk
hiding in abandoned ground squirrel burrows of predation, poor nocturnal eyesight, or other
(Hut and Scharff, 1998). Rainfall allows this variables that may shape activity patterns
animal to emerge above ground in daytime, (Jesseau et al., 2009).
sometimes in extremely high numbers (Table 1,
R.A. Hut, personal observation). A more subtle
response to temperature was described in the
midge (Chironomus thummi), which emerges Factors limiting temporal niche switches
from the water at dusk to swarm and reproduce
at night. Colder water temperatures, however, Factors limiting temporal niche switches may be
trigger dawn emergence and daytime swarming. divided into internal (the organism anatomy and
The low water temperature is apparently used by physiology, such as its sensory and thermoregula-
the insects as an indicator for night temperatures tory capabilities) and external (environmental
being too low to swarm and successfully repro- conditions, both biotic and abiotic). As in the case
duce (Table 1, Kureck, 1979). The bat-eared fox of the factors leading to temporal niche switches,
290

factors limiting temporal niche switches can be adaptations can take place at the anatomical,
associated with environmental factors, and with physiological, or molecular level to define a gen-
inter- and intraspecies interactions. The latter erally nocturnal or diurnal life style for a specific
are reviewed by Castillo-Ruiz et al. (2012). species.
Day and night differ in many environmental The adaptive value of a temporal niche will
factors other than light levels (e.g., temperature, likely present trade-offs due to selective forces
humidity), and therefore expose animals to differ- that act upon specific physiological and anatomical
ent challenges that require different anatomical, systems. For instance, in a species forced to
physiological, and behavioral adaptations. Diurnal become diurnal to escape nocturnal predators
animals usually use vision for predatory avoid- new physiological and behavioral strategies may
ance and foraging, while nocturnal animals use need to be selected to cope with activity at the
tactile probing, smell, and hearing. Communica- warmer phase of the day. The most prominent
tion is usually through sound and olfaction in and predictable environmental variable that may
nocturnal animals (and also, although rarely, exert selective pressure in association with spe-
through bioluminescence). Nocturnal animals cific temporal niches is the dramatic change in
may use camouflage (e.g., moths or night jars) luminance across the 24-h day. Several studies
and burrowing for concealment from their diurnal have focused on adaptations of the visual system
predators during the day, whereas diurnal animals as key features of evolutionary adaptation to
on the other hand use visual signals or even con- specific temporal niches.
spicuous aposematic coloration (Daan, 1981). The large majority of reef fish families are pri-
All these complex adaptations to the specific marily diurnal. Nocturnality has been observed in
temporal niche may limit temporal niche switches. at least 13 families across teleost reef fishes
Here, we focus on some properties of the visual (reviewed in Schmitz and Wainwright, 2011). An
system, the sensory system mostly studied in the analysis of morphological traits of the eyes in 265
context of adaptation to the nocturnal or diurnal species of teleost reef fish in 43 different families
temporal niche. (Schmitz and Wainwright, 2011) revealed that the
eye morphology of nocturnal reef teleosts is
characterized by better light sensitivity, resulting
Visual adaptations to specific temporal niches from large relative eye size, high optical ratio, and
large, rounded pupils. Improved dim-light image
Night- and day-adapted vision formation comes at the cost of reduced depth of
focus and reduction of potential accommodative
Because the activity pattern within an individual lens movement. Diurnal teleost reef fish have
may show considerable plasticity, it may not be much higher morphological and optical diversity
the best parameter to determine the dominant than nocturnal species, with large ranges of optical
temporal niche phenotype for longer time scale ratio, depth of focus, and lens accommodation.
evaluations. Physiological and anatomical deter- The authors suggest that the trade-off between
minants for temporal niche usage likely integrate improved dim-light vision and reduced optical
over a much longer time frames (generations) diversity may be a key factor in explaining the
than individual activity patterns do. The best can- lower trophic diversity of nocturnal reef teleosts.
didate for such a determinant is likely to be found Thus, it is likely that the trade-off associated with
in the eye, since this organ is specifically adapted nocturnal vision may represent a limiting factor
to light levels that range—in up to five orders of for evolutionary radiation in nocturnal taxa.
magnitude—between the darkness of the night Several studies on different taxonomic groups
and the bright light during the day. These visual tested the effect of switching from diurnal; vision
291

based foraging, to nocturnal foraging, showing but the reverse conclusion that lower cone per-
that low light levels can constrain visual abilities centages are always associated with nocturnality,
and reduce foraging success. For example, the cannot be drawn. Much lower retinal cone
common murre (Uria aalge), a diurnal, diving percentages can be found in other species which
predator seabird, extends its activity to the night also display a diurnal preference, and indeed, they
during the breeding season. Diving depths may approach, or even overlap with, cone
decreased with decreasing nocturnal light levels, percentages found in nocturnal species ( 10%,
probably decreasing foraging efficiency (Regular Table 2). Some of these diurnal species with rela-
et al., 2011). Juvenile salmonids are visual for- tive low cone percentages have a clear overrepre-
agers; they switch from being predominantly diur- sentation of cones in the central retina, either in
nal for most of the year to being nocturnal in an elongated region called the visual streak or
winter. Their foraging efficiency is high at light in a central area called the macula. These areas
intensities down to those equivalent to dawn or are tuned to high acuity cone vision and can be
dusk but drops markedly at lower levels of illumi- found in mammals, birds, amphibians, and reptiles
nation; even under the best night condition (full (Rodieck, 1973). Within the macula, some species
moon and clear sky), the feeding efficiency is only have a small indent or pit, packed with cone photo-
35% of their diurnal efficiency (Fraser and receptors to further increase the acuity in a small
Metcalfe, 1997). central area of the visual field. This structure
In the nocturnal house mouse, the retina is is called the fovea and can be found in primates
dominated by rods but also contains 3% cones (including humans) and birds, which may
(Table 2). This suggests that the mouse retina have two or even three foveae. High density cone
retained diurnal visual function. Indeed, at the areas like the macula and the fovea can be seen as
functional level, Cameron et al. (2008) described an adaptation to the diurnal niche. In fact,
increased cone functionality in the subjective most nocturnal haplorhines (Tarsius and Aotus),
day and increased rod functionality during the which likely share a diurnal ancestor with
subjective night. This indicates that an endoge- diurnal haplorhines, have a fovea; although in
nous circadian clock prepares the mouse retina Aotus individuals, it may be absent (Martin and
for visual function during daytime. The existence Ross, 2005).
of this daytime visual response indicates that phe- A study comparing the retinal structure and
notypic plasticity for temporal niche switching on composition of golden spiny mice (A. russatus),
the behavioral level might have generated enough which are diurnal at their natural habitat when
selective advantage to retain some form of diur- this is shared with the nocturnal common spiny
nal vision in this predominantly nocturnal species. mice (A. cahirinus), found that photoreceptor
cells of both species were uniform in shape and
tightly packed, and they conformed mainly to
Eye properties: Rod-cone densities, visual acuity, the morphology of rods (Kronfeld-Schor et al.,
and sensitivity 2001). This could indicate that no obvious
anatomical adaptation to the diurnal life style
The extreme diurnal life style in ground squirrels has been found in the retina of this species so
is associated with two properties of their eyes: far. Accordingly, golden spiny mice prefer to for-
the rod/cone ratio and the spectral filtering pro- age under boulders, where light intensities are
perties of the lens. The European ground squirrel low (22–47 lux; Kronfeld-Schor et al., 2001).
retina contains virtually only cone photoreceptors These results suggest that the response of the
(97%; Szél and Röhlich, 1988). Such extreme rod/ golden spiny mice to become diurnal is either
cone ratios are only found in diurnal mammals, due to the phenotypic plasticity of this species or
292

that the diurnal life style is a relatively recent evo- interaction of crystallines with a UV filter com-
lutionary adaptation. pound, 3-hydroxykynurenine glucoside. Hence,
The increased visual acuity conveyed by the mammalian eye lens could be viewed as a
structures like the macula and fovea would obvi- low-pass filter with a species specific cutoff value
ously be advantageous for nocturnal species as that relates to the species specific life style in the
well. However, these structures require a reduced nocturnal–diurnal gradient.
pupil diameter to restrict the incoming images from Ground squirrel eyes typically have lenses that
the paraxial regions, and a reduced pupil diameter appear deep yellow to the human eye, indicating
is not something nocturnal animals can afford as it stronger filtering properties in the blue region
would reduce the light-gathering ability of the eye of human visible light spectrum. Indeed, short-
(Heesy and Hall, 2010). The way in which noctur- wavelength blue light that is not absorbed by the
nal animals increase visual acuity is by increasing human lens (cutoff 445 nm) is shown to be
the visual field overlap between eyes (reviewed in absorbed by the ground squirrel lens (cutoff
Martin and Ross, 2005; Heesy and Hall, 2010). This 493 nm; Hut et al., 2000). In contrast, nocturnal
not only conveys higher acuity but also increases animals like the Syrian hamster have a cutoff
luminance detection because each point in the value  340 nm, allowing UV-A light penetration
binocular visual field will gather twice as many to the retina (Table 3). The lens filtering property
photons as a point in the monocular field. in the 300–500 nm spectral band seems to reflect
The presence of a tapetum lucidum is seen as the long-term dominant temporal niche usage of
an important adaptation to nocturnal life. The a species (Fig. 3). This lens filtering property is
tapetum is a membrane located behind the retina relatively easy to measure; it has a low within
that reflects back the light that has gone through species variance and is easier to interpret than
the retinal cell layers, increasing light detection complex retinal properties like photoreceptor dis-
by photoreceptor cells. A tapetum is typically tributions (Table 3; Fig. 3). As such, it can be a
present in nocturnal species, such as most strep- useful tool for the evaluation of temporal niche
sirhine primates. Interestingly, there is a wide var- switching within individuals.
iation in the anatomy and histology of the
tapetum, indicating that it is an analogous organ
that has evolved multiple times from widely dif- Eye properties: Ascorbic acid concentration
ferent cell types in the eye (Tan et al., 2005).
Another possible mechanism for protecting the
eye from UV radiation damages is high con-
Eye properties: Spectral filtering and UV centrations of ascorbic acid in the aqueous humor
protection (Ringvold, 1980). High concentrations of ascorbic
acid in the aqueous humor of the eye—20–40
Eyes of diurnal species are under permanent pho- times higher than in the plasma—was reported
tooxidative stress from UV radiation (Ringvold in many diurnal species including humans, ante-
et al., 2000a). It is thought that various levels of lopes, tree shrews, golden spiny mice, and rhesus
yellowish coloration of the mammalian eye lens monkeys. Conversely nocturnal mammals, includ-
shields the retina from the (DNA) damaging ing the slow loris, fruit bats, cats, common spiny
effects of short-wavelength light (ultraviolet and mice, and owl monkeys have a concentration sim-
blue-violet; below  450 nm). Lenses of different ilar to that in plasma (Koskela et al., 1989; Reiss
species may exhibit different amounts of colora- et al., 1986; Ringvold, 1980). Interestingly, such
tion resulting in different spectral filtering pro- a trend was absent in nocturnal and diurnal birds,
perties. The coloration occurs because of the which apparently use a different mechanism for
293

Table 3. Filtering properties of the eye lens relates to the species specific temporal niche

Lens cutoff
Species DTN (nm) References

European ground squirrel Spermophilus citellus D 493 Hut et al. (2000)


Black-tailed prairie dog Cynomys ludovicanus D 482 Yolton et al. (1974)
Mexican ground squirrel Ictidomys mexicanus D 478 Jacobs and Yolton (1971), Yolton et al.
(1974)
Golden mantled ground Callospermophilus D 471 Yolton et al. (1974)
squirrel lateralis
13-lined ground squirrel Ictidomys D 470 Chou and Cullen (1984), Yolton et al.
tridecemlineatus (1974)
California ground squirrel Spermophilus beecheyi D 464 Jacobs et al. (2003)
Agouti Dasyprocta aguti D, N 460 Peichl (1997)
Western gray squirrel Sciurus griseus D 447 Yolton et al. (1974)
Human (32years) Homo sapiens D 445 Stockman et al. (1999)
Guinea pig Cavia porcellus C, A 410 Zigman (1983)
Degu Octodon degus D, N 378 Jacobs et al. (2003)
Talas tuco tuco Ctenomys talarum D, N 370 Schleich et al. (2010)
Rat Rattus norvegicus N 343 Jacobs et al. (2003)
Syrian hamster Mesocricetus auratus N 341 Hut et al. (2000), Brainard et al. (1994)
Coruro Spalacopus cyanus N, D 330 Peichl et al. (2005)
Magellanic tuco tuco Ctenomys magellanicus D, N 325 Schleich et al. (2010)
European mole Talpa europaea A, U 310 Glösmann et al. (2008)

The eye lens absorbs short-wavelength light below a certain wavelength in the UV/blue band to protect the retina from tissue damaging properties of
high-energy short-wavelength light present during daytime. The wavelength at which transmission is reduced to 50% of the maximal transmittance is
presented as lens cutoff value (in nm). DTN, dominant temporal niche; D, diurnal; N, nocturnal; C, crepuscular; A, arrhythmic; U, ultradian.

UV protection. The end product of nitrogen output of circadian clock(s) that are entrained
metabolism in birds is uric acid, which similarly to specific environmental cycles and masking,
to ascorbic acid, is a powerful UV radiation namely, the direct inhibition or stimulation of
absorber. Mammals, and especially diurnal spe- behavior by environmental factors. For example,
cies, have high concentration of ascorbic acid a circadian clock that sustains locomotor activity
and low concentration of uric acid in their aque- during the night combined with an inhibitory effect
ous, while birds have low concentration of of light on activity will result in a robust nocturnal
ascorbic acid and high concentration of uric acid niche. Three basic mechanisms could result in
in their aqueous (Ringvold et al., 2000b). This switching of temporal niche usage, both across
suggests that uric acid in birds may have the same phylogeny or within a species or individual (Fig. 4).
UV protection function as ascorbic acid in First, changes in circadian oscillator properties
mammals, and it will be interesting to compare could lead to changes in its phase angle of entrain-
uric acid concentrations in the aqueous humor ment, and therefore in the phase of the rhythms, it
of diurnal and nocturnal bird species. regulates (see below). Second, a circadian oscilla-
tor with essentially the same properties could lead
to opposite temporal niches if a switch down-
Mechanisms underlying temporal niche switching stream from the oscillator translates the oscillator
output into divergent timing of activity. This seems
The timing of specific behaviors throughout the to be case for the switch between nocturnality
day is the result of the interaction between the and diurnality in rodents. Both, diurnal rats
294

Fig. 3. Cone percentages and lens filtering property relate to dominant temporal niche. Percentage of cones of total number of
photoreceptors in the total retina plotted against lens filtering cutoff wavelength (in nm) correlate with each other such that
diurnal animals are positioned in the top right corner (with the exception of humans) and nocturnal animals in the bottom left
corner. Data taken from Tables 2 and 3, symbol fills correspond to dominant species specific temporal niche. EGS, European
ground squirrel; PD, prairie dog; MGS, Mexican ground squirrel; TLGS, 13-lined ground squirrel; CGS, California ground
squirrel; A, agouti; GS, gray squirrel (Western and Eastern); H, human; GP, guinea pig; D, degu; TTT, Talas tuco tuco; R, rat;
C, cururo; MTT, Magellanic tuco tuco.

(Arvicanthis sp.) and nocturnal laboratory rats leading to a prominent nocturnal niche only
seem to carry a similarly phased master circadian during moonlit nights. The negative masking by
pacemaker within the suprachiasmatic nucleus darkness—or positive making by moonlight—was
(SCN) of the hypothalamus, but the rhythms of evident after measuring activity during total moon
locomotor activity and corticosteroid release that eclipses during full-moon nights, which leads
this oscillator regulates are oppositely phased to dramatic decreases in nocturnal activity
(Kalsbeek et al., 2008; Smale et al., 2003). (Fernandez-Duque et al., 2010). Moreover, in
In accord, the anatomy and the physiology of the order to switch to diurnal activity, animals have
SCN in the diurnal and nocturnal golden and com- to overcome the negative masking effect of light.
mon spiny mice show no differences (Cohen et al., A study comparing masking response to LD pulses
2010a,b), and both, as well as the degu circadian in diurnal golden spiny mice and nocturnal com-
system (Hagenauer and Lee, 2008), are largely mon spiny mice found that whereas the nocturnal
consistent with that of nocturnal species. Third, common spiny mouse display the expected
expression of a behavioral program during the responses, decreasing activity levels in response
daytime or nighttime could be solely determined to a light pulse and increasing activity level
by masking, regardless of the phase of the oscilla- in response to dark pulse, golden spiny mice
tor. Even if masking does not represent the sole increased their activity in response to a dark pulse
temporal niche determinant, it usually plays a crit- (as expected from a nocturnal species), and show
ical role in shaping behavioral outputs of biological insignificant and highly variable response to a light
timing systems. In the owl monkey A. azarai, noc- pulse, indicating that in this diurnally active species
turnal activity regulated by a circadian oscillator is the negative masking effect of light was indeed
inhibited by darkness during new moon nights, removed (Cohen et al., 2010b). Interestingly,
295

Master
oscillator
Output
Input
pathways
pathways
Nocturnal
LD cycle Photoreceptors

Diurnal

Downstream
switch
Nocturnal

Diurnal

Nocturnal

Masking
(+)/(-)

Diurnal

Photic Nonphotic
input input

Fig. 4. Mechanisms underlying temporal niche switches in behavioral outputs. Simplified scenario portraying a single behavioral
output driven by a circadian oscillator. Top: The properties of the circadian oscillator could change, so that its phase relationship
to the LD cycle is different by almost 180 , leading to rhythms of activity that are oppositely phased. Middle: The properties of
the oscillator could be unaltered, but a switch in the output pathways regulating locomotor activity could alternate between
nocturnal and diurnal patterns of activity. Bottom: The properties of the oscillator could be unaltered, but positive and/or
negative masking could determine the rhythmic pattern of activity regardless of the oscillator-phase relationship to the LD cycle.
Modified from Chiesa et al. (2010).

diurnal and nocturnal degus present a stable mas- The top panel of Fig. 4 portraits clocks with dif-
king by light, each according to its respective ferent properties leading to differently phased
chronotype. Thus, whereas diurnal degus increase rhythms. Subtle changes in the circadian period
their activity with light, in nocturnal degus light of a circadian oscillator can lead to robust changes
induce a sharp drop in wheel-running activity in the phase angle of entrainment of the oscillator
(Vivanco et al., 2010b). The resistance to inhibi- (Daan and Aschoff, 1997; Johnson et al., 2003). In
tory effects of light on activity is likely to be essen- the fruit fly Drosophila melanogaster, the tempo-
tial for a switch into a diurnal niche and possibly ral organization of locomotor activity is the result
occur very early in every evolutionary transition of changes in period and phase of their circadian
from nocturnality to diurnality. oscillator. The circadian clock that regulates
296

locomotor activity is constituted by a neuronal for PER to accumulate and for activity to take
network of approximately 150 neurons that place. Other mechanisms contribute to the effects
express clock genes. The products of these genes of photoperiod and ambient temperature on the
constitute the molecular clockwork of biological timing of expression of molecular clock compo-
clocks. D. melanogaster flies usually show crepus- nents. Finally, the fact that the E and M locomo-
cular activity. Recent studies have shown that the tor-activity components rely on different clock
generation of the morning (M) and evening (E) neurons offers further plasticity in the fly’s
peaks of locomotor activity are associated with clock to independently regulate dawn- and dusk-
the activity of anatomically identifiable subgroups associated behaviors and their timing (Dubruille
of cells within the clock neuronal network that and Emery, 2008).
oscillate with different phase and period (Grima The above scenarios only take into account a
et al., 2004; Stoleru et al., 2004, 2007; Yoshii master circadian oscillator and a single environ-
et al., 2012). Interestingly, the relative phase mental factor entraining it. The layout will obvi-
between M and E activity is rather plastic, and ously be more complex in natural environments
it is modulated by both room temperature and where multiple Zeitgebers will be present (Fig. 1).
photoperiod. Shorter photoperiods or lower Further, the presence of multiple peripheral
temperatures will bring the E and M peaks closer oscillators has been shown in all the species stud-
together and lead to more consolidated diurnal ied so far. The phase of these “slave” oscillators is
activity (reviewed in Dubruille and Emery, typically set by a master circadian oscillator, but
2008). The combined effect of photoperiod and they can escape the master control. In mammals,
ambient temperature likely provides adaptive specific peripheral oscillators can be entrained
value; it will push activity to the warmer hours by environmental cycles other than light and in
during the winter but to the cooler dawn and dusk an independent manner from the photic entrain-
during the summer. Remarkably, the shift in the ment of the master circadian oscillator within
fly’s temporal niche relies to some extent on the SCN (Stokkan et al., 2001). Thus, the physiologi-
transcriptional machinery of the molecular clock- cal and behavioral temporal niche of a species in
work. The clock gene period (per) has two nature may represent a complex array of rhythms
mRNA isoforms. Under lower temperatures or that emerge from a network of interacting circa-
short photoperiods intron 8 of the gene is more dian oscillators—within and outside the central
efficiently spliced, and this leads to an earlier nervous system—and their respective interactions
appearance of the PER protein, and of the E with several Zeitgebers with different abilities to
locomotor-activity peak (Majercak et al., 1999). entrain them.
This effect is further enhanced by the fact that
splicing of the per gene has a peak during the
day. This splicing occurs earlier under short Adaptive value of temporal niche switching:
photoperiods, but it is inhibited—and pushed to Unification through the hypothesis of circadian
later times of the day—if temperature is high thermoenergetics?
(Majercak et al., 2004). The effect of photoperiod
on per mRNA maturation and timing of locomo- Temporal niche switching examples have been
tor activity is synergized by the photodegradation described in many species and under various
of TIM, the product of the clock gene timeless circumstances (Table 1). The original descriptions
(tim). TIM dimerizes with PER and by doing so are often accompanied by explanations involving
protects PER from degradation. Under longer the environmental cause which triggers the switch
photoperiods, because of the photic degradation (proximate mechanism) or the adaptive value
of TIM, it is hypothesized that it will take longer of the switching response (ultimate mechanism).
297

The proximate mechanisms may involve light, tem- the same amount of food was delivered in these
perature, or social interactions. Such mechanisms enclosures on a daily basis, but the mice were
are discussed in this chapter, in the accompanying allowed to reproduce freely. This led to the
paper (social interactions; Castillo-Ruiz et al., hypothesis that food shortage might have trig-
2012) and in the original papers (Table 1). Studies gered the observed temporal niche switch at times
evaluating the ultimate mechanisms underlying of high population density. This hypothesis was
temporal niche switching rely more on parsimoni- tested in the lab by letting lab mice work for
ous reasoning than on solid experimental evidence. their food to simulate natural food shortage.
Here, we would like to propose a new hypothesis This “work-for-food” protocol indeed induced
of circadian thermoenergetics (CTE), which diurnality in house mice, showing that temporal
may unify the adaptive value of several cases of niche switching can be induced by reduced food
temporal niche switching. intake (Fig. 5, Hut et al., 2011). This phenomenon
The ultimate mechanism driving temporal has been reproduced in several strains of lab mice
niche switching should involve the optimization (CBA/CaJ, C57bl6, CD1; R.A. Hut, personal
process in which energy balance, survival, and communication), as well as in rats (S. Daan, per-
reproduction are weighed against each other to sonal communication). The results confirm earlier
maximize fitness. In the temperate zone, noctur- findings by Perrigo who showed that pregnant
nal activity is expected to be more costly because female house mice and white footed deer mice
ambient temperatures are high during daytime changed from nocturnal to a diurnal life style
and low during the night. At nighttime, when when these mice had to combine working for
ambient temperatures are generally below the food with the high metabolic demand of lactation
thermal neutral zone of most mammals, endother- (Perrigo, 1987, 1990). Together, these findings
mic animals will energetically profit from reduced indicate that, rather than reduced food intake
heat loss through insulation. This insulation is per se, the negative energy balance triggered the
often found during the rest phase when animals switch from nocturnal activity to diurnal activity.
retreat into burrows or nests, huddle together, The CTE hypothesis can explain these
or adopt a typically sleep-associated thermal insu- findings by assuming a preserved mechanism in
lation posture. Moreover, many endotherms mammals, which associates a diurnal-activity
neglect homeothermy and use torpor during the pattern with lower energy requirements than a
rest phase, making it even more energetically nocturnal-activity pattern. A negative energy bal-
profitable to rest during the night. From an ener- ance (through reduced food intake, lactation,
getics point of view, all endothermic animals in etc.) would induce a switch from nocturnal to
the temperate zone would profit from a diurnal diurnal-activity patterns, in order to save energy
life style. However, other selective forces such as expenditure through reduced heat loss through
daytime predation pressure or competition may increased insulation during the rest phase at night
have forced endothermic animals to become or when ambient temperatures are low. In a similar
remain nocturnal. This nocturnal phenotype can vein, the CTE hypothesis would also predict that
only be sustained when enough food is available heat loss reduction through daily torpor would
to allow for the energetically costly nocturnal life be optimal when it occurs at night time. This
style. If food becomes scarce, nocturnal animals indeed has been found in the house mice in which
may return to a diurnal life style to save energetic food intake was reduced by letting these mice
costs of thermoregulation. Indeed, Daan et al. work for their food. At the end of the protocol,
(2011) observed robust periods of diurnal activity after the mice robustly occupied the diurnal tem-
in populations of normally nocturnal house mice poral niche and placed their rest phase at (end
kept for 2 years in outdoor enclosures. Roughly, of) the night, the mice showed various degrees
298

Fig. 5. Experimentally induced temporal niche switching in mice that work for food. Two representative examples (a, b) of mice in
which food intake (black line in side graphs) was successfully reduced through decreasing food reward for running-wheel
revolutions (gray bars in side graphs). As a result, activity patterns (black bars in actogram) shifted into the light phase of the day
(white bars and white lines in actogram). Peak body temperatures (color coded in actograms) also shifted into the day, but body
temperatures also gradually lowered over the course of the experiment eventually resulting in clear occurrences of torpor during
the rest phase at (end of) the night.

of daily torpor in the dark phase (Fig. 5) (Hut et al., mouse (A. russatus, Levy et al., 2011a,b). It may
2011). At this time, environmental temperatures be that the well studied Djungarian hamster (Pho-
are expected to be the lowest and the energy dopus sungorus), with daily torpor timed at day-
savings through torpor will be maximized. This time (Ruby and Zucker, 1992), is an energetically
nocturnal timing of torpor has also been confirmed suboptimal exception to this rule.
in four species of elephant shrew [Elephantulus The fact that the work for food experiments
rozeti, E. myurus (Lovegrove et al., 2001a,b); E. (Hut et al., 2011) and the effect of ambient tem-
rupestris, E. edwardi (Boyles et al., 2012)], in the perature on circadian rhythms in the degus
fat-tailed dunnart (Sminthopsis crassicaudata, (Hagenauer and Lee, 2008) mentioned above
Warnecke et al., 2008) and in the golden spiny were carried out under a LD cycle under constant
299

temperature indicates that the underlying mecha- Protection of the retina against high energy short-
nism can use the LD cycle and not temperature wavelength light present during daytime is strongly
cycle per se to predict the energetically optimal reflected in the filtering properties of the lens
timing of the activity phase. (Table 3). For most mammalian species, the lens fil-
The observation that ad libitum food supply tering property together with cone densities (and
in the “work-for-food” model (Hut et al., 2011) possibly distribution; Fig. 3) offers a useful measure
immediately restores the nocturnal phase under for assessment of a species’ long-term adaptation to
continuous dim light indicates that the circadian the day- or nighttime niche.
pacemaker did not change its phase angle relative Nocturnal animals need low light levels at night
to the LD cycle. This leads to the interpretation to accommodate vision. The interaction with the
that the mechanism triggering behavioral temporal lunar cycle and the light properties of the envi-
niche switching is downstream from the circadian ronment can explain several cases of temporal
pacemaker (SCN). This situation is schematized niche switching. Such cases have been described
in the middle panel of Fig. 4 and is indeed identical above for nocturnal primates and lobsters,
to the temporal niche switch in diurnal mammals but similar responses may be wide spread among
compared to nocturnal mammals, where activity nocturnal animals.
patterns are in antiphase, but the SCN entrains Temperature can exhibit important and pre-
with a similar phase angle relative to the LD cycle dictable changes between night and day. When
(reviewed in Challet, 2007). temperatures are too high or too low, animals
may modify their activity patterns to reach their
optimal body temperature (in both ectotherms
Conclusions and endotherms) or to minimize the energetic
needs to regulate their body temperature (in
Although temporal niche switching is a complex endotherms only). As such, temperature patterns
phenomenon with a manifold of physiological can explain several cases of modification of activ-
and neurobiological mechanisms and evolution- ity patterns and temporal niche switching. In most
ary causalities, we may be able to draw a number regions on earth, the environmental temperature
of commonalities. This can be based upon the stays well below the lower critical temperatures
observation that the main environmental differ- of most endotherms for a large part of the day.
ences between night and day, which are of direct This means that mostly energy is required to
ecological importance for animals living in the maintain euthermic body temperatures. Energy-
wild, relate to temperature and/or light. saving strategies during the rest phase will cause
Adaptations to the daytime as well as to the nocturnal activity to be energetically more costly
nighttime niche are commonly apparent in the than diurnal-activity patterns. This feature has
eye. Mammals with robust nocturnality like led us to propose the CTE hypothesis which is a
microchiroptera bats have strongly reduced vision useful tool to understand cases of temporal niche
reflected in their tiny eye size, whereas exclusively switching induced by temperature, season, or
diurnal ground squirrels have well developed vision food abundance (Table 1).
reflected in their large eyes and cone dominated
retinas (Table 2; Fig. 3). In other species, long-term
adaptations to a specific temporal niche are also References
present at the level of the retina (e.g., rod and cone
Aguzzi, J., & Sarda, F. (2008). A history of recent
densities, Table 2), but sometimes difficult to inter- advancements on Nephrops norvegicus behavioral and phys-
pret due to variation in local retinal density (i.e., iological rhythms. Reviews in Fish Biology and Fisheries, 18,
fovea, visual streak, photoreceptor gradients). 235–248.
300

Ankel-Simons, F., & Rasmussen, D. T. (2008). Diurnality, Carter-Dawson, L. D., & LaVail, M. M. (1979). Rods and
nocturnality, and the evolution of primate visual systems. cones in the mouse retina. I. Structural analysis using light
Yearbook of Physical Anthropology, 51, 100–117. and electron microscopy. The Journal of Comparative Neu-
Aschoff, J. (1962). Spontaner lokomotorische Aktivität. rology, 188, 245–262.
In Handbuch der zoologie Bd.8, mammalia (pp. 1–76). Castillo-Ruiz, A., Paul, M. J., & Schwartz, W. J. (2012).
Berlin: Walter de Gruyter & Co. In search of a temporal niche: Social interactions. Progress
Bacigalupe, L. D., Rezende, E. L., Kenagy, G. J., & in Brain Research, 199, 267–280.
Bozinovic, F. (2003). Activity and space use by degus: Challet, E. (2007). Minireview: Entrainment of the
A trade-off between thermal conditions and food availability? suprachiasmatic clockwork in diurnal and nocturnal
Journal of Mammalogy, 84, 311–318. mammals. Endocrinology, 148, 5648–5655.
Bakker, E. S., Reiffers, R. C., Olff, H., & Gleichman, J. M. Chiesa, J. J., Aguzzi, J., García, J. A., Sardà, F., & de La
(2005). Experimental manipulation of predation risk and Iglesia, H. O. (2010). Light intensity determines temporal
food quality: Effect on grazing behaviour in a central-place niche switching of behavioral activity in deep-water
foraging herbivore. Oecologia, 146, 157–167. Nephrops norvegicus (Crustacea: Decapoda). Journal of
Bartell, P. A., & Gwinner, E. (2005). A separate circadian Biological Rhythms, 25, 277–287.
oscillator controls nocturnal migratory restlessness in the Chou, B. R., & Cullen, A. P. (1984). Spectral transmittance of
songbird Sylvia borin. Journal of Biological Rhythms, 20, the ocular media of the thirteen-lined ground squirrel
538–549. (Spermophilus tridecemlineatus). Canadian Journal of Zool-
Bäumler, W. (1975). Activity of some small mammals in the ogy, 62, 825–830.
field. Acta Theriologica, 20, 365–377. Cohen, R., Kronfeld-Schor, N., Ramanathan, C., Baumgras, A.,
Begall, S., Daan, S., Burda, H., & Overkamp, G. J. F. (2002). & Smale, L. (2010). The substructure of the suprachiasmatic
Activity patterns in a subterranean social rodent, nucleus: Similarities between nocturnal and diurnal spiny
Spalacopus cyanus (Octodontidae). Journal of Mammalogy, mice. Brain, Behavior and Evolution, 75, 9–22.
83, 153–158. Cohen, R., Smale, L., & Kronfeld-Schor, N. (2009). Plasticity of
Benedix, J. H. (1994). A predictable pattern of daily activity circadian activity and body temperature rhythms in golden
by the pocket gopher Geomys bursarius. Animal Behavior, spiny mice. Chronobiology International, 26, 430–446.
48, 501–509. Cohen, R., Smale, L., & Kronfeld-Schor, N. (2010). Masking
Blanchong, J. A., McElhinny, T. L., Mahoney, M. M., & and temporal niche switches in spiny mice. Journal of
Smale, L. (1999). Nocturnal and diurnal rhythms in the Biological Rhythms, 25, 47–52.
unstriped Nile rat, Arvicanthis niloticus. Journal of Biological Coverdill, A. J., Bentley, G. E., & Ramenofsky, M. (2008).
Rhythms, 14, 364–377. Circadian and masking control of migratory restlessness in
Boal, C. W., & Giovanni, M. D. (2007). Raptor predation on Gambel’s white-crowned sparrow (Zonotrichia leucophrys
Ord’s kangaroo rats: Evidence for diurnal activity by a gambelii). Journal of Biological Rhythms, 23, 59–68.
nocturnal rodent. The Southwestern Naturalist, 52, 291–295. Crompton, A. W., Taylor, C. R., & Jagger, J. A. (1978). Evo-
Bovet, J., & Oertli, E. F. (1992). Free-running activity rhythms lution of homeothermy in mammals. Nature, 272, 333–336.
in free-living beaver (Castor canadensis). Journal of Com- Crowford, P. (1953). The daily cycle of activity in British
parative Physiology, 92, 1–10. shrews. Proceedings of the Zoological Society of London,
Boyles, J. G., Smit, B., Sole, C. L., & McKechnie, A. E. (2012). 123, 715–729.
Body temperature patterns in two syntopic elephant shrew Curcio, C. A., Sloan, K. R., Kalina, R. E., &
species during winter. Comparative Biochemistry and Physi- Hendrickson, A. E. (1990). Human photoreceptor topogra-
ology. Part A, Molecular & Integrative Physiology, 161, phy. The Journal of Comparative Neurology, 292, 497–523.
89–94. Daan, S. (1981). Adaptive daily strategies in behavior. In
Brainard, G. C., Barker, F. M., Hoffman, R. J., Stetson, M. H., J. Aschoff (Ed.), Handbook of behavioral neurobiology
Hanifin, J. P., Podolin, P. L., et al. (1994). Ultraviolet vol. 4: Biological rhythms (pp. 275–297). New York:
regulation of neuroendocrine and circadian physiology in Plenum Publishing Cooperation.
rodents. Vision Research, 34, 1521–1533. Daan, S., & Aschoff, J. (1997). The entrainment of circadian
Bubenik, A. B. (1960). Rhythme nicthéméral des ongulés systems. In J. S. Takahashi, F. W. Turek & R. Y. Moore
souvages. Mammalia, 24, 1–58. (Eds.), Handbook of behavioral neurobiology vol. 4: Circa-
Cameron, M. A., Barnard, A. R., Hut, R. A., Bonnefont, X., dian clocks (pp. 1–28). New York: Plenum Press.
van der Horst, G. T., Hankins, M. W., et al. (2008). Electro- Daan, S., Spoelstra, K., Albrecht, U., Schmutz, I., Daan, M.,
retinography of wild-type and Cry mutant mice reveals Daan, B., et al. (2011). Lab mice in the field: Unorthodox
circadian tuning of photopic and mesopic retinal responses. daily activity and effects of a dysfunctional circadian clock
Journal of Biological Rhythms, 23, 489–501. allele. Journal of Biological Rhythms, 26, 118–129.
301

DeCoursey, P. J. (1960). Daily light sensitivity rhythm in a Fulk, G. W. (1976). Notes on activity, reproduction, and social-
rodent. Science, 131, 33–35. behavior of Octodon-Degus. Journal of Mammalogy, 57,
DeCoursey, P. J. (1986). Light-sampling behavior in photo- 495–505.
entrainment of a rodent circadian rhythm. Journal of Com- Gaillard, F., Bonfield, S., Gilmour, G. S., Kuny, S.,
parative Physiology A, Neuroethology, Sensory, Neural, Mema, S. C., Martin, B. T., et al. (2008). Retinal anatomy
and Behavioral Physiology, 159, 161–169. and visual performance in a diurnal cone-rich laboratory
Doyle, S. E., Yoshikawa, T., Hillson, H., & Menaker, M. rodent, the nile grass rat (Arvicanthis niloticus). The Journal
(2008). Retinal pathways influence temporal niche. Pro- of Comparative Neurology, 510, 525–538.
ceedings of the National Academy of Sciences of the United Garcia-Allegue, R., Lax, P., Madariaga, A. M., &
States of America, 105, 13133–13138. Madrid, J. A. (1999). Locomotor and feeding activity
Dubruille, R., & Emery, P. (2008). A plastic clock: How circa- rhythms in a light-entrained diurnal rodent, Octodon degus.
dian rhythms respond to environmental cues in Drosophila. American Journal of Physiology—Regulatory, Integrative
Molecular Neurobiology, 38, 129–145. and Comparative Physiology, 277, R523–R531.
Eriksson, L. O. (1973). Spring inversion of the diel rhythm of Gattermann, R., Johnston, R. E., Yigit, N., Fritzsche, P.,
locomotor activity in young sea-going brown trout, Salmo Larimer, S., Ozkurt, S., et al. (2008). Golden hamsters are
trutta trutta L., and Atlantic salmon, Salmo salar L. Aquilo nocturnal in captivity but diurnal in nature. Biology Letters,
Series Zoologica, 14, 68–79. 4, 253–255.
Erkert, H. G. (1974). Der Einfluss des Mondlichtes auf die Glösmann, M., Steiner, M., Peichl, L., & Ahnelt, P. K. (2008).
Aktivitätsperiodik nachtaktiver Säugetiere. Oecologia, 14, Cone photoreceptors and potential UV vision in a subterra-
269–287. nean insectivore, the European mole. Journal of Vision, 8
Erkert, H. G. (2008). Diurnality and nocturnality in nonhuman (23), 1–12.
primates: Comparative chronobiological studies in labora- Govardovskii, V. I., Röhlich, P., Szél, A., & Khokhlova, T. V.
tory and nature. Biological Rhythm Research, 39, 229–267. (1992). Cones in the retina of the Mongolian gerbil,
Erkinaro, E. (1969). Der Phasenwechsel der lokomotorischen Meriones unguiculatus: An immunocytochemical and
Aktivität bei Microtus agrestis (L.), M. arvalis (Pall.) und electrophysiological study. Vision Research, 32, 19–27.
M. oeconomus (Pall.). Aquilo Series Zoologica, 8, 1–31. Grima, B., Chelot, E., Xia, R., & Rouyer, F. (2004). Morning
Feldman, J. L., & Phillips, C. J. (1984). Comparative retinal and evening peaks of activity rely on different clock neurons
pigment epithelium and photoreceptor ultrastructure in noc- of the Drosophila brain. Nature, 431, 869–873.
turnal and fossorial rodents—the Eastern woodrat, Neotoma Gwinner, E. (1996). Circadian and circannual programmes in
floridana, and the plains pocket gopher, Geomys bursarius. avian migration. Journal of Experimental Biology, 199,
Journal of Mammalogy, 65, 231–245. 39–48.
Fenn, M. G. P., & Macdonald, D. W. (1995). Use of middens Hagenauer, M. H., & Lee, T. M. (2008). Circadian organiza-
by red foxes—Risk reverses rhythms of rats. Journal of tion of the diurnal Caviomorph rodent, Octodon degus.
Mammalogy, 76, 130–136. Biological Rhythm Research, 39, 269–289.
Fernandez-Duque, E., de la Iglesia, H., & Erkert, H. G. Häussler, U., & Erkert, H. (1978). Different direct effects of
(2010). Moonstruck primates: Owl monkeys (Aotus) need light intensity on the entrained activity rhythm in neotropi-
moonlight for nocturnal activity in their natural environ- cal bats (Chiroptera, Phyllostomidae). Behavioural Processes,
ment. PLoS One, 5, e12572. 3, 223–239.
Foster, R. G., & Menaker, M. (1993). Circadian photorecep- Heesy, C. P., & Hall, M. I. (2010). The nocturnal bottleneck
tion in mammals and other vertebrates. In L. Wetterberg and the evolution of mammalian vision. Brain, Behavior
(Ed.), Light and biological rhythms in man (pp. 73–91). and Evolution, 75, 195–203.
Oxford: Pergamon. Hendrickson, A., & Hicks, D. (2002). Distribution and density
Fox, R. J., & Bellwood, D. R. (2011). Unconstrained by the of medium- and short-wavelength selective cones in the
clock? Plasticity of diel activity rhythm in a tropical reef fish, domestic pig retina. Experimental Eye Research, 74,
Siganus lineatus. Functional Ecology, 25, 1096–1105. 435–444.
Fraser, N. H. C., Heggenes, J., Metcalfe, N. B., & Thorpe, J. E. Honma, S., & Hiroshige, T. (1977). Pubertal manifestation of
(1995). Low summer temperatures cause juvenile Atlantic sex difference in circadian rhythm of corticotrophin-releasing
salmon to become nocturnal. Canadian Journal of activity in the rat hypothalamus. Acta Endocrinologica, 86,
Zoology—Revue Canadienne de Zoologie, 73, 446–451. 225–234.
Fraser, N. H. C., & Metcalfe, N. B. (1997). The costs of becom- Hut, R. A., Pilorz, V., Boerema, A. S., Strijkstra, A. M., &
ing nocturnal: Feeding efficiency in relation to light intensity Daan, S. (2011). Working for food shifts nocturnal mouse
in juvenile Atlantic Salmon. Functional Ecology, 11, 385–391. activity into the day. PLoS One, 6, e17527.
302

Hut, R. A., & Scharff, A. (1998). Endoscopic observations on Kronfeld-Schor, N., & Dayan, T. (2003). Partitioning of time
tunnel blocking behaviour in the European ground squirrel as an ecological resource. Annual Review of Ecology, Evolu-
(Spermophilus citellus). Z Säugetierkunde, 63, 377–380. tion, and Systematics, 34, 153–181.
Hut, R. A., Scheper, A., & Daan, S. (2000). Can the circadian Kronfeld-Schor, N., & Dayan, T. (2008). Activity patterns of
system of a diurnal and a nocturnal rodent entrain to ultra- rodents: The physiological ecology of biological rhythms.
violet light? Journal of Comparative Physiology A, 186, Biological Rhythm Research, 39, 193–211.
707–715. Kronfeld-Schor, N., Dayan, T., & Jones, M. E. (2001). Retinal
Hut, R. A., Van Oort, B. E. H., & Daan, S. (1999). Natural structure and foraging microhabitat use of the golden spiny
entrainment without dawn and dusk: The case of the Euro- mouse (Acomys russatus). Journal of Mammalogy, 82,
pean ground squirrel (Spermophilus citellus). Journal of 1016–1025.
Biological Rhythms, 14, 290–299. Kryger, Z., Galli-Resta, L., Jacobs, G. H., & Reese, B. E. (1998).
Hutchinson, G. E. (1957). Population studies—Animal ecology The topography of rod and cone photoreceptors in the retina
and demography—Concluding remarks. Cold Spring Harbor of the ground squirrel. Visual Neuroscience, 15, 685–691.
Symposia on Quantitative Biology, 22, 415–427. Kureck, A. (1979). Two circadian eclosion times in
Jacobs, G. H., Calderone, J. B., Fenwick, J. A., Krogh, K., & Chironomus thummi (Diptera), alternately selected with
Williams, G. A. (2003). Visual adaptations in a diurnal different temperatures. Oecologia, 40, 311–323.
rodent, Octodon degus. Journal of Comparative Physiology Kushnirov, D., Beolchini, F., Lombardini, F., & Nevo, E.
A, Neuroethology, Sensory, Neural, and Behavioral Physiol- (1998). Radiotracking studies in the blind mole rat. In:
ogy, 189, 347–361. Abstracts Euro-American mammal congress, Santiago de
Jacobs, G. H., & Yolton, R. L. (1971). Visual sensitivity and Compostella, Spain, 381.
color vision in ground squirrels. Vision Research, 11, LaVail, M. M. (1976). Survival of some photoreceptor cells in
511–537. albino-rats following long-term exposure to continuous light.
Jesseau, S. A., Holmes, W. G., & Lee, T. M. (2009). Commu- Investigative Ophthalmology, 15, 64–70.
nal nesting and discriminative nursing by captive degus, Lehmann, U. (1976). Short-term and circadian rhythms in the
Octodon degus. Animal Behavior, 78, 1183–1188. behaviour of the vole, Microtus agrestis (L.). Oecologia, 23,
Johnson, C. H., Elliott, J. A., & Foster, R. (2003). Entrainment 185–199.
of circadian programs. Chronobiology International, 20, Lehmann, U., & Sommersberg, C. (1980). Activity patterns of
741–774. the common vole, Microtus arvalis—Automatic recording of
Kalsbeek, A., Verhagen, L. A., Schalij, I., Foppen, E., behavior in an enclosure. Oecologia, 47, 61–75.
Saboureau, M., Bothorel, B., et al. (2008). Opposite actions Levy, O., Dayan, T., & Kronfeld-Schor, N. (2007). The relation-
of hypothalamic vasopressin on circadian corticosterone ship between the golden spiny mouse circadian system and its
rhythm in nocturnal versus diurnal species. European Journal diurnal activity: An experimental field enclosures and labora-
of Neuroscience, 27, 818–827. tory study. Chronobiology International, 24, 599–613.
Kas, M. J., & Edgar, D. M. (1999). A nonphotic stimulus Levy, O., Dayan, T., & Kronfeld-Schor, N. (2011a). Interspe-
inverts the diurnal-nocturnal phase preference in Octodon cific competition and torpor in golden spiny mice: Two sides
degus. Journal of Neuroscience, 19, 328–333. of the energy-acquisition coin. Integrative and Comparative
Kempinger, L., Dittmann, R., Rieger, D., & Helfrich- Biology, 51, 441–448.
Förster, C. (2009). The nocturnal activity of fruit flies Levy, O., Dayan, T., & Kronfeld-Schor, N. (2011b). Adaptive
exposed to artificial moonlight is partly caused by direct thermoregulation in golden spiny mice: The influence of
light effects on the activity level that bypass the endogenous season and food availability on body temperature. Physio-
clock. Chronobiology International, 26, 151–166. logical and Biochemical Zoology, 84, 175–184.
Kitchen, A. M., Gese, E. M., & Schauster, E. R. (2000). Lockard, R. B. (1978). Seasonal change in activity pattern of
Changes in coyote activity patterns due to reduced exposure Dipodomys-Spectabilis. Journal of Mammalogy, 59, 563–568.
to human persecution. Canadian Journal of Zoology, 78, Lourens, S., & Nel, J. A. J. (1990). Winter activity of bat-eared
853–857. foxes Otocyon megalotis on the cape west-coast. South
Kleitman, N., & Engelmann, T. G. (1953). Sleep characteristics African Journal of Zoology, 25, 124–132.
of infants. Journal of Applied Physiology (Bethesda, MD: Lovegrove, B. G., Raman, J., & Perrin, M. R. (2001a). Daily
1985), 6, 269–282. torpor in elephant shrews (Macroscelidea: Elephantulus
Koskela, T. K., Reiss, G. R., Brubaker, R. F., & Ellefson, R. D. spp.) in response to food deprivation. Journal of Compara-
(1989). Is the high concentration of ascorbic acid in the tive Physiology B, 171, 11–21.
eye an adaptation to intense solar irradiation? Investigative Lovegrove, B. G., Raman, J., & Perrin, M. R. (2001b). Hetero-
Ophthalmology & Visual Science, 30, 2265–2267. thermy in elephant shrews, Elephantulus spp. (Macroscelidea):
303

Daily torpor or hibernation? Journal of Comparative Physiol- Peichl, L., Chavez, A. E., Ocampo, A., Mena, W., Bozinovic, F.,
ogy B, 171, 1–10. & Palacios, A. G. (2005). Eye and vision in the subterranean
Majercak, J., Chen, W. F., & Edery, I. (2004). Splicing of the rodent cururo (Spalacopus cyanus, Octodontidae). The Journal
period gene 30 -terminal intron is regulated by light, circadian of Comparative Neurology, 486, 197–208.
clock factors, and phospholipase C. Molecular and Cellular Peichl, L., & González-Soriano, J. (1994). Morphological types
Biology, 24, 3359–3372. of horizontal cell in rodent retinae: A comparison of rat,
Majercak, J., Sidote, D., Hardin, P. E., & Edery, I. (1999). mouse, gerbil, and guinea pig. Visual Neuroscience, 11,
How a circadian clock adapts to seasonal decreases in tem- 501–517.
perature and day length. Neuron, 24, 219–230. Peichl, L., Künzle, H., & Vogel, P. (2000). Photoreceptor types
Martin, R. D., & Ross, C. F. (2005). J. Kremers (Ed.), The evo- and distributions in the retinae of insectivores. Visual Neu-
lutionary and ecological context of primate vision (pp. 1–36). roscience, 17, 937–948.
New York: John Wiley & Sons Ltd. Peichl, L., Nemec, P., & Burda, H. (2003). S-Cone dominance
Metcalfe, N. B., Fraser, N. H. C., & Burns, M. D. (1998). in the retinae of subterranean African mole-rats (Rodentia,
State-dependent shifts between nocturnal and diurnal activ- Bathyergidae). Investigative Ophthalmology & Visual Sci-
ity in salmon. Proceedings of the Royal Society of London. ence, 44, U457.
Series B, 265, 1503–1507. Pereira, J. A. (2010). Activity pattern of Geoffroy’s cats
Metcalfe, N. B., Fraser, N. H. C., & Burns, M. D. (1999). Food (Leopardus geoffroyi) during a period of food shortage.
availability and the nocturnal vs. diurnal foraging trade-off Journal of Arid Environments, 74, 1106–1109.
in juvenile salmon. Journal of Animal Ecology, 68, 371–381. Perrigo, G. (1987). Breeding and feeding strategies in deer
Morton, S. R. (1995). Fat-tailed dunnart Sminthopsis mice and house mice when females are challenged to work
crassicaudata (Gould, 1844). In R. Strahan (Ed.), The for their food. Animal Behavior, 35, 1298–1316.
mammals of Australia (pp. 129–131). Sydney: Reed Books. Perrigo, G. (1990). Food, sex, time, and effort in a small mam-
Mrosovsky, N., & Hattar, S. (2005). Diurnal mice (Mus mal: Energy allocation strategies for survival and reproduc-
musculus) and other examples of temporal niche switching. tion. Behaviour, 114, 191–205.
Journal of Comparative Physiology A, Neuroethology, Sen- Pizzatto, L., Child, T., & Shine, R. (2008). Why be diurnal?
sory, Neural, and Behavioral Physiology, 191, 1011–1024. Shifts in activity time enable young cane toads to evade can-
Müller, B., & Peichl, L. (1989). Topography of cones and rods nibalistic conspecifics. Behavioral Ecology, 19, 990–997.
in the tree shrew retina. The Journal of Comparative Neurol- Potvin, C. L., & Bovet, J. (1975). Annual cycle of patterns of
ogy, 282, 581–594. activity rhythms in beaver colonies (Castor canadensis).
Nespolo, R. F., Bacigalupe, L. D., Figueroa, C. C., Koteja, P., Journal of Comparative Physiology, 98, 243–256.
& Opazo, J. C. (2011). Using new tools to solve an old prob- Refinetti, R. (2006). Variability of diurnality in laboratory
lem: The evolution of endothermy in vertebrates. Trends in rodents. Journal of Comparative Physiology, 192, 701–714.
Ecology & Evolution, 26, 414–423. Refinetti, R. (2008). The diversity of temporal niches in
Nyholm, E. S. (1965). Zur Ökologie von Myotis mystacinus mammals. Biological Rhythm Research, 39, 173–192.
(Leisl.) und M. daubentoni (Leisl.) (Chiroptera). Annales Regular, P. M., Hedd, A., & Montevecchi, W. A. (2011). Fish-
Zoologici Fennici, 2, 77–123. ing in the dark: A pursuit-diving seabird modifies foraging
Ocampo-Garcès, A., Mena, W., Hernández, F., Cortès, N., & behaviour in response to nocturnal light levels. PLoS One,
Palacios, A. G. (2006). Circadian chronotypes among wild- 6, e26763.
captured west Andean Octodontids. Biological Research, Reiss, G. R., Werness, P. G., Zollman, P. E., &
39, 209–220. Brubaker, R. F. (1986). Ascorbic acid levels in the aqueous
Oster, H., Avivi, A., Joel, A., Albrecht, U., & Nevo, E. (2002). humor of nocturnal and diurnal mammals. Archives of Oph-
A switch from diurnal to nocturnal activity in S. ehrenbergi thalmology, 104, 753–755.
is accompanied by an uncoupling of light input and the cir- Richter, C. P. (1978). Dark-active rat transformed into light-
cadian clock. Current Biology, 12, 1919–1922. active rat by destruction of 24-hr clock—Function of 24-hr
Palacios, A. G., Bozinovic, F., Vielma, A., Arrese, C. A., clock and synchronizers. Proceedings of the National Acad-
Hunt, D. M., & Peichl, L. (2010). Retinal photoreceptor emy of Sciences of the United States of America, 75,
arrangement, SWS1 and LWS opsin sequence, and electro- 6276–6280.
retinography in the South American marsupial Thylamys Ringvold, A. (1980). Aqueous humour and ultraviolet radia-
elegans (Waterhouse, 1839). The Journal of Comparative tion. Acta Ophthalmologica, 58, 69–82.
Neurology, 518, 1589–1602. Ringvold, A., Anderssen, E., & Kjönniksen, I. (2000a). Distri-
Peichl, L. (1997). Die Augen der Säugetiere: Unterschiedliche bution of ascorbate in the anterior bovine eye. Investigative
Blicke in die Welt. Biologie in unserer Zeit, 27, 96–105. Ophthalmology & Visual Science, 41, 20–23.
304

Ringvold, A., Anderssen, E., & Kjönniksen, I. (2000b). UV Szél, A., & Röhlich, P. (1988). Four photoreceptor types in the
absorption by uric acid in diurnal bird aqueous humor. Inves- ground squirrel retina as evidenced by immunocytochemistry.
tigative Ophthalmology & Visual Science, 41, 2067–2069. Vision Research, 28, 1297–1302.
Rocha, F. A., Ahnelt, P. K., Peichl, L., Saito, C. A., Szél, A., & Röhlich, P. (1992). Two cone types of rat retina
Silveira, L. C., & De Lima, S. M. (2009). The topography detected by anti-visual pigment antibodies. Experimental
of cone photoreceptors in the retina of a diurnal rodent, Eye Research, 55, 47–52.
the agouti (Dasyprocta aguti). Visual Neuroscience, 26, Tan, Y., Yoder, A. D., Yamashita, N., & Li, W. H. (2005). Evi-
167–175. dence from opsin genes rejects nocturnality in ancestral
Rodieck, R. W. (1973). The vertebrate retina: Principles of primates. Proceedings of the National Academy of Sciences
structure and function. Oxford, England: W. H. Freeman. of the United States of America, 102, 14712–14716.
Roll, U., Dayan, T., & Kronfeld-Schor, N. (2006). On the role Tomotani, B. M., Flôres, D. E. F. L., Paliza, J. D., Oda, G. A.,
of phylogeny in determining activity patterns of rodents. & Valentinuzzi, V. S. (2011). Field and laboratory studies
Evolutionary Ecology, 20, 479–490. provide insights into the meaning of day-time activity in a
Rowsemitt, C. N., Petterborg, L. J., Claypool, L. E., subterranean rodent (Ctenomys aff. Knighti), the tuco-tuco.
Hoppensteadt, F. C., Negus, N. C., & Berger, P. J. PLoS ONE 7, e37918.
(1982). Photoperiodic induction of diurnal locomotor Tscherkowitsch, G. M. (1953). Ontogenese der tagesperiodik
activity in Microtus montanus, the montane vole. Canadian bei Affen. In Bykow Regulation physiologischer Functionen.
Journal of Zoology—Revue Canadienne de Zoologie, 60, Moskau, Band, 2, (pp. 187–198).
2798–2803. Turner, D. C. (1975). The vampire bat: A field study in behav-
Ruby, N. F., & Zucker, I. (1992). Daily torpor in the absence ior and ecology. Baltimore: Johns Hopkins University Press.
of the suprachiasmatic nucleus in Siberian hamsters. Ameri- Van Oort, B. E. H., Tyler, N. J. C., Gerkema, M. P., Folkow, L.,
can Journal of Physiology, 263, R353–R362. & Stokkan, K. A. (2007). Where clocks are redundant: Weak
Scheibler, E., & Wollnik, F. (2011). Interspecific contact circadian mechanisms in reindeer living under polar photic
affects time use in free-living Desert hamsters in Alashan conditions. Naturwissenschaften, 94, 183–194.
desert, China. In: 12th meeting European biological rhythms Vivanco, P., Rol, M. A., & Madrid, J. A. (2010a). Temperature
society, Oxford, U.K. cycles trigger nocturnalism in the diurnal homeotherm
Schleich, C. E., Vielma, A., Glösmann, M., Palacios, A. G., & Octodon degus. Chronobiology International, 27, 517–534.
Peichl, L. (2010). Retinal photoreceptors of two subterra- Vivanco, P., Rol, M. A., & Madrid, J. A. (2010b). Pacemaker
nean tuco-tuco species (Rodentia, Ctenomys): Morphology, phase control versus masking by light: Setting the circadian
topography, and spectral sensitivity. The Journal of Com- chronotype in dual Octodon degus. Chronobiology Interna-
parative Neurology, 518, 4001–4015. tional, 27, 1365–1379.
Schmitz, L., & Motani, R. (2011). Nocturnality in dinosaurs Warnecke, L., Turner, J. M., & Geiser, F. (2008). Torpor and
inferred from scleral ring and orbit morphology. Science basking in a small arid zone marsupial. Naturwissenschaften,
(New York, NY), 332, 705–708. 95, 73–78.
Schmitz, L., & Wainwright, P. C. (2011). Nocturnality con- Weinert, D., Weinandy, R., & Gattermann, R. (2007). Photic
strains morphological and functional diversity in the eyes and non-photic effects on the daily activity pattern of Mon-
of reef fishes. BMC Evolutionary Biology, 11, 338. golian gerbils. Physiology & Behavior, 90, 325–333.
Shkolnik, A. (1971). Diurnal activity in a small desert rodent. West, R. W., & Dowling, J. E. (1975). Anatomical evidence for
International Journal of Biometeorology, 15, 115–120. cone and rod-like receptors in the gray squirrel, ground
Smale, L., Lee, T., & Nunez, A. A. (2003). Mammalian squirrel, and prairie dog retinas. The Journal of Compara-
diurnality: Some facts and gaps. Journal of Biological tive Neurology, 159, 439–460.
Rhythms, 18, 356–366. Whitford, W. G., Depree, D. J., Hamilton, P., &
Stockman, A., Sharpe, L. T., & Fach, C. (1999). The spectral Ettershank, G. (1981). Foraging ecology of seed-harvesting
sensitivity of the human short-wavelength sensitive cones ants, Pheidole spp. in a Chihuahuan desert ecosystem.
derived from thresholds and color matches. Vision Research, American Midland Naturalist, 105, 159–167.
39, 2901–2927. Yolton, R. L., Yolton, D. P., Renz, J., & Jacobs, G. H. (1974).
Stokkan, K. A., Yamazaki, S., Tei, H., Sakaki, Y., & Preretinal absorbance in sciurid eyes. Journal of Mammal-
Menaker, M. (2001). Entrainment of the circadian clock in ogy, 55, 14–20.
the liver by feeding. Science, 291, 490–493. Yoshii, T., Rieger, D., & Helfrich-Förster, C. (2012). Two
Stoleru, D., Nawathean, P., Fernandez, M. P., Menet, J. S., Clocks in the Brain—An Update of the Morning and Even-
Ceriani, M. F., & Rosbash, M. (2007). The Drosophila circa- ing Oscillator Model in Drosophila. Progress in Brain
dian network is a seasonal timer. Cell, 129, 207–219. Research, 199, 56–82.
Stoleru, D., Peng, Y., Agosto, J., & Rosbash, M. (2004). Young, J. Z. (1962). The life of the vertebrates (2nd ed.).
Coupled oscillators control morning and evening locomotor Oxford, UK: The Clarendon Press.
behaviour of Drosophila. Nature, 431, 862–868. Zigman, S. (1983). Effects of near ultraviolet radiation on the
Strubbe, J. H., & Gorissen, J. (1980). Meal patterning in the lens and retina. Documenta ophthalmologica—Advances in
lactating rat. Physiology & Behavior, 25, 775–777. ophthalmology, 55, 375–391.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 18

Feedback actions of locomotor activity to the


circadian clock

Alun T. L. Hughes* and Hugh D. Piggins

Faculty of Life Sciences, University of Manchester, Manchester, United Kingdom

Abstract: The phase of the mammalian circadian system can be entrained to a range of environmental
stimuli, or zeitgebers, including food availability and light. Further, locomotor activity can act as an
entraining signal and represents a mechanism for an endogenous behavior to feedback and influence
subsequent circadian function. This process involves a number of nuclei distributed across the brain
stem, thalamus, and hypothalamus and ultimately alters SCN electrical and molecular function to
induce phase shifts in the master circadian pacemaker. Locomotor activity feedback to the circadian
system is effective across both nocturnal and diurnal species, including humans, and has recently been
shown to improve circadian function in a mouse model with a weakened circadian system. This raises
the possibility that exercise may be useful as a noninvasive treatment in cases of human circadian
dysfunction including aging, shift work, transmeridian travel, and the blind.

Keywords: circadian; suprachiasmatic; locomotor activity; wheel-running; exercise; clock gene; VIP;
aging; arousal; NPY; serotonin.

Introduction physiological functions. A feature of these daily, or


circadian, rhythms (derived from the Latin circa,
The rotation of the Earth about its axis imposes meaning about, and dies, a day) is that they are gen-
upon the vast majority of living organisms predict- erated by endogenous circadian clocks expressed
able daily cycles with a period of  24 h. In order within the cells and tissues of an organism and
to perform optimally in an appropriate temporal continue to oscillate with a period of close to 24 h
niche, organisms have developed corresponding even when the organism is isolated from all
cyclic rhythms in their own behavioral and external timing influences. Endogenous circadian
oscillators are thought to confer a selective advan-
*Corresponding author. tage on organisms through predictive homeostasis;
Tel.: þ44-161-306-0502; Fax: þ44-161-275-3938 organisms are able to anticipate predictable envi-
E-mail: alun.t.hughes@manchester.ac.uk ronmental events, such as dusk and dawn or times

http://dx.doi.org/10.1016/B978-0-444-59427-3.00018-6 305
306

of high predator/prey activity, before they occur and vasoactive intestinal polypeptide (VIP) that
can therefore prepare appropriately (Piggins and receives heavy external input from the retina and
Guilding, 2011). Indeed, the lack of a circadian clock other regions of the brain, and the remainder of
can influence survival in a natural environment the SCN that receives somewhat lighter extra-
(DeCoursey et al., 2000), and possession of a clock SCN innervation (Abrahamson and Moore,
that runs with an inappropriate cycle length is disad- 2001; Morin, 2007). The SCN generates rhythms
vantageous in laboratory-based competition studies of high- and low-neuronal discharge (Brown and
(Woelfle et al., 2004). Piggins, 2007) driven by an interlocking network
The mammalian circadian system consists of of core clock genes and other rhythmic processes
input pathways (Fig. 1), a central “master” pace- (Glossop, 2011; Hastings et al., 2008; Ko and Taka-
maker located in the suprachiasmatic nuclei hashi, 2006), though recent evidence indicates
(SCN) of the hypothalamus, numerous extra-SCN that the electrical output of the SCN is more com-
oscillators distributed throughout the brain and plex than this simple model suggests (Belle et al.,
body, and output pathways to effect changes in 2009). Under normal environmental and physio-
the temporal organization of physiology and logical conditions, the SCN dominates the circadian
behavior (Golombek and Rosenstein, 2010; system of mammals, coordinating the phase of
Kalsbeek et al., 2006). Anatomically, the SCN can rhythms in the various tissue-specific extra-SCN
be loosely divided into two subregions, one oscillators (see Dibner et al., 2010; Guilding and
characterized by expression of the neuropeptide Piggins, 2007).
To benefit an organism, endogenous circadian
rhythms must be maintained in phase with envi-
ronmental time cues. In addition to specific chem-
IGL

Serotonin
ical agents such as melatonin (van Geijlswijk
DR
et al., 2010), certain stimuli including light and a
Orexin
HT

variety of light-independent, or nonphotic, cues


G

are capable of phase-dependently phase resetting


AB Y
NP

LHA the circadian system. Repeated exposure to these


G

O
re
xi
n

Eye Glu signals at appropriate times can synchronize, or


Orexin

RH entrain, the circadian system to the environment


T

MR (Golombek and Rosenstein, 2010). The purpose


nin
SCN ro
to of this review is to collate and summarize the
PA

Se
CA

phase resetting and entraining actions of locomo-


P

tor activity feedback to the mammalian circadian


Fig. 1. Schematic “wiring diagram” showing nuclei and system.
neurotransmitters of the mammalian central circadian system
involved with signaling photic and nonphotic entrainment
information. Photic signals are transmitted directly to the
Photic entrainment
SCN via the RHT utilizing glutamate and PACAP.
Nonphotic information, however, reaches the SCN via a
network of connections involving the DR and MR, the IGL, and The most ubiquitous entraining stimulus, or zeitge-
the LHA. These pathways use serotonin, NPY and GABA, and ber, encountered in daily life for the majority of
orexin, respectively as their main transmitters. DR, dorsal raphe; organisms is that of the light–dark (LD) cycle of
GABA, g-amino butyric acid; GHT, geniculohypothalamic
the solar day; therefore, to fully appreciate non-
tract; Glu, glutamate; IGL, intergeniculate leaflet; LHA, lateral
hypothalamic area; MR, median raphe; PACAP, pituitary photic influences on the circadian system, we must
adenylate-cyclase-activating peptide; RHT, retinohypothalamic first briefly consider photic entrainment. As the
tract; SCN, suprachiasmatic nucleus. endogenous SCN-driven period of circadian
307

rhythms for most mammals differs slightly from responses to light encountered during an animal’s
24 h, rhythms must be periodically shifted back “subjective” night (i.e., night according to endoge-
toward external environmental time in order to nous circadian phase) are similar in both nocturnal
maintain an appropriate phase relationship with and diurnal species (e.g., Slotten et al., 2005) and
the environment (Challet, 2007). Light encoun- can be mimicked in a controlled laboratory envi-
tered during the early part of the night phase of ronment. Brief (5–10 min) pulses of light delivered
an animal’s endogenous rhythm, for instance, if a to animals otherwise housed in constant darkness
nocturnal rodent were to become active before (DD) produce phase-dependent shifts in the timing
dusk, delays the clock. This shifts activity to a of the rest-activity cycle defining a phase response
slightly later time on subsequent cycles, realigning curve (PRC; see Fig. 2) to photic stimulation (Daan
it with the external environment. Conversely, light and Pittendrigh, 1976; Schwartz and Zimmerman,
encountered during the late part of an animal’s 1990). Such procedures, in addition to confirming
endogenous night phase, if a nocturnal rodent were the advance and delay portions of the photic
to remain active after dawn, advances the phase of PRC, also reveal a photically nonresponsive zone
the clock. This shifts activity to a slightly earlier during the subjective day, a time at which light
time on subsequent days, again realigning behavior would be normal, so phase shifts would be inappro-
and physiology with the external world. These priate (Fig. 2).

(a) (b)
Nonphotic PRC
0 +3.0
Phase shift (h)

18
–3.0
Subjective day Subjective night
Days

(c) Photic PRC


+3.0

36
Phase shift (h)

–3.0
Subjective day Subjective night
54

Fig. 2. Stylized schematic representation of phase shift data for a laboratory rodent in constant darkness. (a) Bouts of induced
locomotor activity (open ellipses), stimulated by mechanisms such as novel wheel confinement or triazolam injection, produce
marked phase advances during the mid-to-late subjective day and modest phase delays during the late subjective night. Brief
light pulses (open circles) produce differential phase-dependent phase shifts in behavior when presented during the early and
late subjective night and fail to shift behavior during the subjective day. Each row represents 24 h and black bars represent
voluntary home-cage locomotor activity. (b) Phase response curve (PRC) summarizing the phase-dependent phase responses of
nocturnal rodents to a generalized “nonphotic” zeitgeber such as 3 h of induced locomotor activity. (c) PRC summarizing the
phase-shift responses of nocturnal rodents to brief (5–60 min) light pulses. Reflecting the requirements of different circadian
systems, in species where tau is shorter than 24 h (as shown), the phase delay portion of the PRC is usually larger than the phase
advance portion. In species where tau is longer than 24 h, the advance portion is usually larger.
308

Phase shifts such as those described above in birdsong (Gwinner, 1966; Menaker and Eskin,
response to light are typically assessed by measure- 1966). These experiments, along with early reports
ment of the daily onset or offset of locomotor or of social entrainment in humans (Aschoff, 1979;
consumptive activities of an animal, however, Wever, 1979) elegantly revealed that circadian
these shifts in behavior reflect changes to the phase rhythms could be altered using so-called nonphotic
of the underlying SCN oscillator. Light signals are mechanisms; that is, any phase shifting or
delivered to the central SCN oscillator via a gluta- entraining stimulus independent of light itself or
mate and pituitary adenylate-cyclase-activating the specific neural pathways activated by light. A
peptide-expressing projection (Fig. 1) from a wide variety of different stimuli that fall into this
subset of intrinsically photosensitive melanopsin- category are capable of signaling environmental
expressing retinal ganglion cells (Hannibal and information to the circadian system and these
Fahrenkrug, 2006; Morin and Allen, 2006). Gluta- nonphotic stimuli can be broadly divided into
mate agonists and antagonists infused or injected two major subclasses with very different functions
into the third ventricle in vivo or applied to SCN and behavioral responses: those associated with
brain slices in vitro, mimic and block the effects of feeding and those that induce locomotor activity
light and optic nerve stimulation, respectively and/or arousal (Mistlberger and Antle, 2011).
(Abe et al., 1991; Ebling et al., 1991; Mintz and
Albers, 1997; Shibata et al., 1994; Vindlacheruvu
et al., 1992), confirming the role of glutamate in Entrainment to restricted feeding and the
signaling photic information to the SCN. food-entrainable oscillator

The importance of regular food availability and


Negative masking by light dire consequences of failure to acquire food
places special import on the ability of an animal
In addition to entrainment of a functional oscillator, to track predictable feeding opportunities. As
light can directly affect the expression of locomotor such, animals have evolved a food-entrainable
activity in nocturnal species through an alternative, oscillator (FEO) which functions independently
more direct route: masking. Negative masking is of the SCN and can become decoupled from it
the direct suppression of locomotor activity by light (see Davidson, 2006 and Mistlberger and Antle,
in nocturnal species (Doyle and Menaker, 2007) and 2011 for review). FEO-driven food anticipatory
is mediated by the same melanopsin-expressing behavior can be driven out of phase from the
retinal ganglion cells mentioned above, however, SCN-controlled main active phase by the restric-
the neural and neurochemical processes that medi- tion of food presentation to a time during an
ate masking in the brain are currently poorly under- animal’s normal inactive phase (Boulos et al.,
stood (Goz et al., 2008; Hattar et al., 2003; Li et al., 1980; Honma et al., 1983). The anatomical and bio-
2005; Panda et al., 2003). chemical bases of this FEO are not yet known,
though have been the focus of much debate and
controversy in recent years (e.g., Fuller et al.,
Nonphotic entrainment 2009; Mistlberger et al., 2009). We do know, how-
ever, that this oscillator persists in the absence of
The concept that circadian rhythms could be the SCN (Stephan et al., 1979a,b), in the absence
influenced by mechanisms other than light came of a critical core clock gene (bmal1; Storch and
to prominence in the mid-1960s through a series Weitz, 2012) and that daytime restricted feeding
of experiments demonstrating that the rhythms of not only induces food anticipatory activity prior
birds could be entrained to the daily playback of to food presentation, but also potently entrains
309

the phase of liver and other peripheral oscillators Locomotor activity as a nonphotic stimulus
associated with metabolism, while the SCN
remains entrained to the LD cycle (Damiola A number of techniques can be used to induce
et al., 2000; Stokkan et al., 2001). locomotor activity in laboratory rodents including
confinement to a novel-running wheel, injections
of some benzodiazepines, and forced treadmill
Nonphotic stimuli based on locomotor activity running (e.g., Mistlberger, 1991a,b; Mrosovsky
and/or arousal and Salmon, 1987; Turek and Losee-Olson, 1986).
Each of these, along with scheduled voluntary
The second major subclass of nonphotic stimuli is wheel access (Edgar and Dement, 1991), has been
composed of those associated with locomotor used extensively to investigate the effects of overt
activity and/or arousal. A great variety of different locomotor activity on circadian function.
stimuli of this nature can influence circadian func- The involvement of induced locomotor activity
tion, including the introduction of novelty to the in benzodiazepine-induced circadian effects was
environment, for instance, the presentation of a originally the subject of some controversy. Initial
new running wheel (Mrosovsky and Salmon, studies using triazolam, a short-acting benzodiaze-
1987); social interactions with conspecifics, and pine, were designed to investigate the role of
in particular, proximity of inaccessible reproduc- GABA signaling in circadian biology (e.g., Turek
tively active females to male test animals (Honrado and Losee-Olson, 1986, 1987a) and the realization
and Mrosovsky, 1989); forced or voluntary loco- that triazolam-mediated effects were reliant on
motor activity (Edgar and Dement, 1991; induced locomotor activity came later: hamsters
Marchant and Mistlberger, 1996); daytime pulses denied wheel access following injections show no
of darkness to nocturnal animals (Boulos and circadian effects, and crucially, the magnitude of
Rusak, 1982); and gentle handling (Antle and Mis- triazolam-mediated effects correlates with the
tlberger, 2000). This wide range of stimuli involves amount of time spent wheel-running during the
a common element, however: the induction of vig- first 6 h after an injection (Mrosovsky and Salmon,
orous locomotor activity and/or the induction of 1990; Van Reeth and Turek, 1989a,b). It should be
arousal. Moreover, injection of chemicals that noted however that, while locomotor activity
induce locomotor exercise, for instance, certain undoubtedly feeds back to the circadian system,
benzodiazepines (see below) or morphine injections of chlordiazepoxide, a benzodiazepine
(Marchant and Mistlberger, 1995), can act as non- that suppresses locomotor activity, also cause
photic zeitgebers. This group of general locomotor nonphotic-like phase shifts, demonstrating that
activity/arousal-associated stimuli differs funda- benzodiazepines in their own right can directly
mentally to feeding-related time cues and the confer some nonphotic information, presumably
FEO in that they function through mechanisms through their effect on GABA signaling (Hastings
that, similarly to photic zeitgebers, converge and et al., 1998; Meyer et al., 1993). Indeed, injections
act on the central SCN oscillator. Locomotor of muscimol, a specific GABAA agonist, produce
activity/arousal-associated events induce a PRC a nonphotic PRC (Smith et al., 1989).
completely different to that of light, characterized
by large phase advances during the mid-to-late
subjective day and relatively small phase delays in Circadian responses to discrete “pulses” of
the late subjective night (Mrosovsky et al., 1989; locomotor activity
Fig. 2). It is this second group of nonphotic stimuli,
and, in particular, locomotor activity/exercise that Single, discrete, locomotor activity pulses induced
will form the main focus of this review. during the mid-to-late part of the subjective day
310

phase advance free-running behavioral rhythms in which is then permanently left in place (Gannon
hamsters. Both triazolam injections and confine- and Rea, 1995).
ment to a novel-running wheel produce this effect, The relatively long exposures to exercise
in both DD and constant light (LL; Reebs and required for large shifts indicate that the circadian
Mrosovsky, 1989a; Reebs et al., 1989; Turek and system of nocturnal rodents has a relatively high
Losee-Olson, 1986, 1987a). Triazolam injections, threshold for exercise to alter circadian function,
in addition to shifting the phase of locomotor activ- at least compared to light, which produces measur-
ity rhythms, have also been shown to phase shift able phase shifts after as little as a few minutes of
the rhythm of LH release in female hamsters exposure (Sharma and Chandrashekaran, 1997;
(Turek and Losee-Olson, 1988). The efficacy of Takahashi et al., 1984). This relatively low sensitiv-
single locomotor activity pulses to shift rhythms in ity to locomotor activity may represent a buffering
other species is less clear, however. For instance, system to prevent inappropriate phase shifting
“pulses” of novelty-induced activity and 3-h forced to small amounts of activity which may be normal
treadmill running produce only modest and incon- outside the main active phase.
sistent phase shifts in mice (Challet et al., 2000;
Marchant and Mistlberger, 1996).
PRCs to both triazolam-induced (Turek and Interactions between nonphotic and photic
Losee-Olson, 1986) and novel wheel-induced entraining stimuli
locomotor activity pulses (Reebs and Mrosovsky,
1989a) similarly define a classical nonphotic pro- Previously in this review, we have considered
file, with maximal phase advances, of  2–3 h, the well-characterized effects of both light and
when activity is induced during the mid-to-late nonphotic influences on the phase of behavioral
subjective day and somewhat smaller phase circadian rhythms. While in the laboratory we aim
delays seen during the late subjective night to control variables in order to determine
(Reebs and Mrosovsky, 1989a,b; Wickland and responses to specific stimuli, in nature, it would
Turek, 1991; Fig. 2). Following initial contradic- be more common for different zeitgeber modalities
tory reports, a careful assessment of the literature to be encountered in combination. Responses of
including consideration of confounding factors in the mammalian circadian system to photic and
individual experiments supports the notion that nonphotic stimuli are both mediated by effects on
the amount of locomotor activity performed does, the central SCN oscillator (Challet, 2007). As such,
in the majority of cases, correlate with shift mag- it is important to consider the convergent, yet dis-
nitude (elegantly reviewed in Mrosovsky, 1996). tinct, responses of animals to photic and nonphotic
Shift magnitude in response to novelty-induced signals and to examine the interactions between
locomotor activity correlates with the amount of them when presented either together or in series.
locomotor activity performed (Bobrzynska and
Mrosovsky, 1998; Janik and Mrosovsky, 1993;
Weisgerber et al., 1997; Wickland and Turek, Dark pulses
1991) and a minimum of  3 h of running is neces-
sary to induce maximal responses (Reebs and Behavioral responses to pulses of darkness pres-
Mrosovsky, 1989b; Wickland and Turek, 1991). ented to nocturnal rodents housed otherwise in
Individuals that neglect to run vigorously tend to LL (Boulos and Rusak, 1982; Ellis et al., 1982)
show minimal phase shifts. Further, the induction produce a PRC that is somewhat similar to those
of extremely large phase shifts, in the order of described for typical locomotor activity/arousal-
 12 h, is possible with the combined stimulus of associated nonphotic stimuli (Canal and Piggins,
transfer to DD and presentation of a novel wheel 2006; Marston et al., 2008). Originally considered
311

a mirror image of the photic PRC (Subbaraj and and overt behavior with the environment. Under
Chandrashekaran, 1978), these responses are normal conditions, reentrainment to an 8-h
now thought to arise from a combination of both advance such as this would require  8 days.
nonphotic and photic components (Coogan and However, this latency to reentrainment can be
Piggins, 2005; Rosenwasser and Dwyer, 2001), dramatically reduced in laboratory rodent models
though the contribution of locomotor activity to by locomotor activity induced at the beginning of
dark pulse phase shifts has been debated. Total the dark phase on the first day in the shifted LD
wheel revolutions during a pulse do not correlate cycle, by either 3 h of novel-running wheel con-
significantly with shift magnitude, but a trend finement or triazolam injection (Mrosovsky and
toward more revolutions resulting in larger shifts Salmon, 1987, 1990; Van Reeth and Turek, 1987,
is evident (Coogan and Piggins, 2005; Marston 1989a,b). These locomotor activity pulses result
et al., 2008). However, specific assessment of in a rapid advance in the onset of subsequent
locomotor activity induced within an individual home-cage activity, with a shift of  7 h clearly
by subjective day dark pulses, accounting for visible by the second cycle in the new lighting
baseline activity on previous days, does correlate schedule. In addition to accelerating behavioral
well with shift magnitude (Canal and Piggins, reentrainment, locomotor activity pulses such as
2006). Further, the magnitude of phase shifts is this also accelerate the reentrainment of some
attenuated by restriction of wheel-running activity peripheral oscillators (Yamanaka et al., 2008).
during the pulse (Reebs et al., 1989). These data Rapid, locomotor activity-induced, phase
suggest that locomotor activity does indeed form reentrainment to an LD shift and the previously
the nonphotic component of a dark pulse, but described PRCs to novel wheel/triazolam-induced
that a change in the level of activity within locomotor activity in DD provide an interesting
an individual is important, rather than a pan- opportunity to examine the interaction of loco-
individual threshold of activity. A photic compo- motor exercise and photic cues. The large  7h
nent arising from the resumption of LL at the phase advance seen in response to induced day-
end of a pulse is postulated to give rise to the time exercise shortly after an LD shift cannot be
differences between typical nonphotic PRCs and explained by the peak of the advance portion of
the dark pulse PRC (Rosenwasser and Dwyer, the PRC to exercise alone, creating an apparent
2001) and may be mediated at the molecular level discrepancy between the large phase shifts neces-
by interactions between photic and nonphotic sary for rapid reentrainment and the 3-h peak
resetting as they converge at the SCN (Maywood shift for discrete pulses of locomotor activity seen
and Mrosovsky, 2001). in PRCs. This can be reconciled, however, by view-
ing the phase response of a hamster to a combina-
tion of an LD shift and exercise as synergistic of
Reentrainment to an LD cycle shift and photic/ both the nonphotic cue of induced activity and
nonphotic conflict studies the photic cue of light falling at inappropriate
phases of the circadian cycle (as a result of the
Sudden shifts in the phase of the lighting regime LD shift; Mrosovsky 1991; Reebs and Mrosovsky,
under LD simulate the jetlag encountered by 1989b). Indeed, presenting light pulses at different
humans following transmeridian air travel. An phases relative to a phase-advancing exercise pulse
8-h advance of the LD cycle approximates the produces variable phase shifts depending on the
temporal disruption experienced following a flight temporal relationship between locomotor activity
from California to the United Kingdom and and light (Joy and Turek, 1992; Mrosovsky, 1991;
requires a series of stepwise phase-advancing Reebs and Mrosovsky, 1989b). For instance, a light
shifts to resynchronize endogenous oscillators pulse following within 3 h of a late subjective
312

daytime activity pulse would fall in the delay portion performed (Shioiri et al., 1991; Yamada et al.,
of the photic PRC and all but cancels the phase- 1990), an effect also seen, though somewhat
advancing effect of locomotor activity, while a light inconsistently, in hamsters (Mrosovsky, 1999;
pulse presented 8 h after subjective daytime activity Weisgerber et al., 1997).
enhances the resultant phase shift as a summation of
the two phase-advancing stimuli (Mrosovsky, 1991;
Reebs and Mrosovsky, 1989b). Entrainment to scheduled locomotor activity
Similarly, inducing locomotor activity during
subjective nighttime light pulses dramatically To be of functional use for entrainment, however,
attenuates phase shifts compared to those in feedback information from locomotor activity must
response to light alone (Ralph and Mrosovsky, demonstrate the ability to synchronize free-running
1992) and parallel treatment of SCN-containing rhythms through repeated phase adjustments when
brain slices in vitro with neurochemicals that the opportunity to indulge in locomotor activity
mediate photic and nonphotic resetting recipro- is either offered, induced, or forced repeatedly on
cally blocks the phase-resetting effects of the sequential cycles.
other (Biello et al., 1997). This spectrum of Induced and scheduled locomotor activity can
responses to the combined effects of photic and alter the phase angle of entrainment to an LD cycle
nonphotic zeitgebers defines a complex relation- (Mistlberger, 1991b; Mistlberger and Holmes,
ship between them, reflective of the complexity 2000; Salgado-Delgado et al., 2008; Turek and
of temporal signals organisms are exposed to in Losee-Olson, 1987b). The direction of phase angle
the real world and the underlying complexity of changes in these studies is predictable based on
the extended circadian system itself. the PRCs to induced locomotor activity. Activity
induced near the light-to-dark transition, proximal
to the advance portion of the locomotor PRC,
Entrainment to exercise: Longer-term effects of results in a more positive (advanced) phase angle
locomotor activity of entrainment and activity late in the dark
phase results in a delayed phase angle. Other
We have, up to now, considered the short-term reports, however, have suggested that under
effects of acute “pulses” of locomotor activity on some circumstances a more complex relationship
free-running circadian rhythms in constant between photic and nonphotic zeitgebers may exist
conditions and the ability of such pulses to modify (Sinclair and Mistlberger, 1997).
the reentrainment of behavior following LD Stable entrainment of rodent behavioral rhythms
shifts. In addition to these acute effects, ad libitum in the absence of an LD cycle has also been
access to a running wheel, and hence daily, shown. Serial daily injections of triazolam to
though unscheduled, vigorous locomotor exercise enucleated hamsters and daily novel wheel
also feeds back to regulate central circadian func- confinement of hamsters in DD both entrain
tion. The presence of a home-cage running wheel behavioral rhythms (Reebs and Mrosovsky,
shortens tau in rats and mice (Benus et al., 1988; 1989a; Van Reeth and Turek, 1989a,b, 1990).
Edgar et al., 1991; Mistlberger and Holmes, Similarly, mice entrain to both voluntary and
2000; Mistlberger et al., 1998; Yamada et al., forced scheduled locomotor exercise of between
1986, 1988) and has been suggested, though not 2 and 12 h each day (Dallmann et al., 2007; Edgar
proven, to do so in hamsters (Pratt and Goldman, and Dement, 1991; Edgar et al., 1997; Laemle and
1986; though see also Aschoff et al., 1973). In rats, Ottenweller, 1999; Marchant and Mistlberger,
the magnitude of tau shortening by wheel-running 1996; Marchant et al., 1997; Power et al., 2010;
correlates with the number of wheel revolutions Figs. 3 and 4a) and while apparently a weak
313

stimulus in rats, forced treadmill running entrains with fetal SCN transplants, that lack extensive
free-running rhythms in this species too neural connections with the host, fail to phase
(Mistlberger, 1991a). “Aftereffects” are com- advance in response to induced subjective day activ-
monly observed such that following scheduled ity (Canbeyli et al., 1991).
exercise, free-running tau is closer to the period
of the entraining signal than in initial free
run (Dallmann et al., 2007; Laemle and Phase angle of entrainment to scheduled
Ottenweller, 1999; Marchant and Mistlberger, locomotor activity and the nonphotic PRC
1996; Power et al., 2010; Fig. 4a; see also Reebs
and St-Coeur, 1994). Radio frequency lesions in Importantly, in all successful locomotor activity
mice have shown the SCN to be essential for entrainment studies to date, the phase angle of
entrainment to scheduled locomotor activity; entrainment of behavioral rhythms to repeated
mice rendered behaviorally arrhythmic by com- locomotor activity, be it voluntary, induced, or
plete SCN lesions do run on the treadmill, but forced, is dependent on the relative periods of the
scheduled forced treadmill activity fails to either entraining stimulus and endogenous free-running
induce or sustain behavioral circadian rhythms rhythms of subject animals (e.g., Dallmann et al.,
(Marchant et al., 1997). Further, SCN efferent 2007). When entrainment by phase delay is
connections appear to be essential for activity- required, that is, endogenous period, tau, is
induced phase modulations; SCN-lesioned hamsters shorter than the entraining period, T, rhythms

(a) (b)
0 0

18 18
Days

Days

36 36

54 54
Entrainment by delay Entrainment by advance

Fig. 3. Stylized schematic representation of entrainment to scheduled locomotor activity. The scheduled opportunity for vigorous
locomotor activity, be it voluntary wheel-running or forced treadmill running, entrains behavioral rhythms in most species
examined. The black outline boxes in (a) and (b) indicate timing of scheduled activity and black bars represent behavior
recorded using a nonrunning wheel mechanism such as drinking or general locomotor activity. Each row represents 24 h. (a)
When an animal’s endogenous tau is shorter than the zeitgeber period (T) entrainment is achieved through daily phase delays.
Synchronizing the opportunity to run with the delay portion of the nonphotic phase response curve results in a positive phase
angle of entrainment. (b) When an animal’s endogenous tau is longer than T, entrainment is achieved through daily
phase advances, resulting in a neutral or small negative phase angle of entrainment. These characteristics are consistent with the
phase response curve shown in Fig. 2b and have been described across a range of both nocturnal and diurnal species.
314

(a) WT (b) Vipr2–/–


Wheel-running Drinking Wheel-running Drinking
LD
DD1
locomotor activity
Scheduled
DD2

0 24 48 0 24 48 0 24 48 0 24 48
Time (h)

Fig. 4. Scheduled locomotor activity acts as a potent zeitgeber to entrain behavior in a mouse model with a weakened SCN
pacemaker and induces lasting improvements in circadian function. (a) Wheel-running and drinking behavior for a wild-type
(C57BL/6J) mouse under LD, free-running in DD then subjected to 50 days of scheduled voluntary locomotor activity before a
second DD free run. (b) Wheel-running and drinking behavior for a mouse lacking functional expression of the VPAC2 receptor
(Vipr2/) under the same experimental protocol as (a). Black outline boxes indicate timing of scheduled activity, data are
double plotted such that each row shows 48 h. Wild-type mice gradually entrain to 6 h of daily scheduled voluntary locomotor
activity, visible on the drinking panel, and subsequently free-run with a modestly lengthened tau. Vipr2/ mice rapidly entrain
to scheduled activity and display a profound and persistent change in subsequent free-running tau of 24 h. Note that behavior
in (a) entrains according to the rules described in Fig. 3 and consistent with the PRC in Fig. 2b. Postscheduled activity behavior
in (b) free runs in phase with the previous opportunity to exercise. See Power et al. (2010) for further examples and interpretation.

entrain with a phase angle that places activity at what was previously the light phase, the subjec-
the end of the subjective night (Fig. 3a), the delay tive day phase advance portion of the nonphotic
portion of the nonphotic PRC (Fig. 2b). When tau PRC (e.g., Mrosovsky and Salmon, 1987; van
is longer than T, entrainment by advance is Reeth and Turek, 1987; Fig. 2b).
required resulting in a phase angle that places
activity at the end of the subjective day/beginning
of the subjective night (Fig. 3b); the advance por- The roles of arousal and physiological correlates
tion of the PRC to locomotor activity (Fig. 2b). of locomotor activity in entrainment
The modulation of phase angle of entrainment
to LD cycles by scheduled locomotor activity The role of exercise/locomotor activity in entrain-
described above (Mistlberger, 1991b; Turek and ment is intimately linked with the arousal state of
Losee-Olson, 1987b) is also consistent with this an animal (Mrosovsky, 1996), and to date, even
model, as is the acceleration of reentrainment to 25 years after the first reports of locomotor activity
an LD cycle advance by activity induced during feeding back to the central oscillator, the precise
315

nature of this signal has yet to be resolved. It et al. (1996) report that with an appropriate exper-
is clear that locomotor activity feedback can alter imental design, cold-induced running can be as
circadian phase, both acutely and as stable entrain- effective at producing phase shifts as novelty-
ment, however, it is unknown which aspect or com- induced running. Indeed, scheduled forced tread-
bination of aspects of locomotor activity are crucial mill running, which potentially lacks the a priori
for this effect. Locomotor activity represents a motivation to run, is as effective at inducing
complex amalgam of behavioral, physiological, entrainment in mice as the scheduled opportunity
and emotional factors. These include an underlying for voluntary wheel-running (Marchant and
arousal state related to both the motivation to Mistlberger, 1996; Marchant et al., 1997).
exercise and associated rewarding feedback,
physiological adaptations to demanding physical
activity, increases in body temperature during vig- Physiological correlates of locomotor activity
orous exercise, physical exhaustion, and increases
in consumptive behaviors inevitably associated Few studies have examined the roles of physio-
with increases in energy expenditure. A number logical correlates of locomotor activity in entrain-
of studies have attempted to identify the key caus- ment, though a number of reports have been
ative elements of locomotor activity-induced phase published. Inadvertent scheduling of water intake
shifts though no consensus has yet been reached. is an inevitable consequence of ad libitum water
availability during scheduled activity studies (e.g.,
see Edgar and Dement, 1991; Marchant and
Arousal and motivation Mistlberger, 1996; Power et al., 2010), but is not suf-
ficient for entrainment: Marchant and Mistlberger
In general, the number of wheel revolutions (1996) report that scheduled drinking without
performed correlates with shift magnitude at a par- scheduled exercise fails to entrain behavior.
ticular circadian phase (e.g., Biello et al., 1994; The possible role of inadvertent scheduled feed-
Mrosovsky and Salmon, 1990; see Mrosovsky, ing during scheduled exercise has yet to be specifi-
1996 for a detailed interpretation of different cally examined. While daytime restricted feeding
methods of inducing activity). It has, however, under an LD cycle is effective in entraining a com-
been suggested that locomotor activity per se may ponent of behavioral rhythms and can alter SCN
not be sufficient for phase shifts to occur. One function (Verwey and Amir, 2009), it does not
alternative method to those discussed previously entrain the SCN (Damiola et al., 2000). However,
that can induce wheel-running is lowering ambient daytime restricted feeding does entrain peripheral
housing temperature, though increasing wheel- oscillators (Damiola et al., 2000; Stokkan et al.,
running during a novelty-induced activity bout by 2001) and scheduled feeding in DD in the absence
lowering ambient temperature does not necessarily of other timing cues does entrain the SCN and
result in increased shift magnitude (Janik and major SCN-controlled circadian systems, at least
Mrosovsky, 1993; Mrosovsky and Biello, 1994). in some mouse strains (Abe et al., 1989, 2007;
This has been interpreted as a lack of some neces- Castillo et al., 2004). The timing of feeding during
sary element of motivation or arousal associated scheduled locomotor activity entrainment has yet
with cold-induced wheel-running increases; run- to be recorded though is likely to be high around
ning in the cold may be necessary for thermoregu- the times of increased locomotor activity. Indeed,
lation whereas running at normal temperatures forced activity during the light phase of an LD
may simply be rewarding (Janik and Mrosovsky, cycle is reported to increase feeding during this
1993; Mrosovsky and Biello, 1994). This methodol- time (Salgado-Delgado et al., 2008) and temporally
ogy has, however, been questioned: Mistlberger targeted feeding hastens reentrainment to an LD
316

shift (Angeles-Castellanos et al., 2011). As such, a PER1 expression in the SCN throughout the circa-
role for the inadvertent scheduling of feeding in dian cycle, but fails to alter the phase of this rhythm
entrainment to exercise is a possibility that cannot (Salgado-Delgado et al., 2010).
be discounted. A role for body temperature While the mechanisms of signal transduction
rhythms, which will be higher during vigorous run- that mediate this suppressive action of locomotor
ning also remains a possibility though has yet to be activity on per gene expression are poorly under-
systematically investigated. stood, a number of reports indicate the involve-
ment of the extracellular signal-regulated kinase
(ERK) pathway in response to this and other
Molecular, neural, and peripheral bases of nonphotic stimuli. Dark pulses during the subjec-
nonphotic phase-shifting and entrainment tive day, that induce vigorous wheel-running in
nocturnal rodents, suppress ERK phosphoryla-
Molecular basis of phase shifts tion in the SCN (Coogan and Piggins, 2005),
while sleep deprivation, a (nonexercise related)
At the molecular level, circadian rhythms are gen- nonphotic arousal stimulus, suppresses ERK
erated within cell autonomous pacemakers phosphorylation in most areas of the SCN. In par-
through the rhythmic expression of a network of allel to this general suppression, this stimulus
core clock genes and their proteins, including, increases phospho-ERK levels in a distinct subre-
among others, the period (per) genes, per1 and gion that corresponds with the area that receives
per2, the cryptochrome 1 and 2 genes and bmal1. both retinal and geniculothalamic input (see
Our current understanding of the components Section “Neuropeptide Y and the IGL” below;
and workings of this core molecular oscillator has Antle et al., 2008).
been extensively reviewed elsewhere (Glossop, Further clues as to the mechanisms of nonphotic
2011; Ko and Takahashi, 2006), so this discussion control of per expression can be gleaned from the
shall be limited to the role of per/PER in the con- somewhat better described mechanisms of signal
text of nonphotic phase shifting and entrainment. transduction for photic information. Representing
The per genes are the main targets within the a point of convergence between the broadly oppo-
core molecular oscillator for stimuli that phase shift site effects of photic and nonphotic stimuli on the
the clock (though photic influences on other clock circadian system, per1 and per2, and their proteins,
genes have been described, see Challet et al., are induced by phase-shifting light pulses delivered
2003 for review). Locomotor activity induced by at appropriate nighttime phases (Albrecht et al.,
both novel wheel confinement and brotizolam 1997; Shigeyoshi et al., 1997; Yan and Silver,
(a benzodiazepine with similar phase-shifting 2002). These photic signals induce per expression
actions to triazolam) suppresses the expression of through activating a number of second messenger
per1 and per2 mRNA in the SCN at times that cascades, including ERK, protein kinase A, and
would result in a behavioral phase shift (Maywood Ca2 þ/calmodulin-dependent protein kinase II sig-
et al., 1999; Yokota et al., 2000). Further, and cru- naling and ultimately lead to transcriptional activa-
cially, injection of per1 antisense oligonucleotides tion of the per genes via activation of cAMP
in the mid-subjective day, the approximate time response elements in their promoters (Golombek
of maximal nonphotic phase shifts and maximal and Ralph, 1995; Tischkau et al., 2000; von Gall
per1 expression, suppresses per1 levels and induces et al., 1998). Given the synergy between per sup-
nonphotic-like phase advances (Hamada et al., pression and induction by nonphotic and photic
2004). Interestingly, forced activity during the light stimuli, respectively, and the demonstration of
phase of an LD cycle, therefore in conflict with the ERK pathway involvement in responses to both,
photic zeitgeber, appears to suppress levels of it is likely that at least some of the photic signaling
317

pathways mentioned above, and specifically Neuropeptide Y and the IGL


their suppression, may also function in mediating
nonphotic signals including those arising due to The intergeniculate leaflet (IGL) is a narrow
locomotor activity. band of the visual thalamus that lies between
the dorsal and ventral parts of the lateral genicu-
late nucleus and abundantly expresses neuropep-
Locomotor activity suppresses electrical output tide Y (NPY) and GABA, both of which are
of the SCN also found in the major IGL efferent projection
to the SCN, the geniculohypothalamic tract
The mechanistic connections between the core (GHT; Harrington, 1997; Fig. 1).
molecular oscillator and membrane-level electri- Evidence supporting a role for the IGL and NPY
cal activity have yet to be fully elucidated in mediating nonphotic signals to the central circa-
(Colwell, 2011). We cannot, therefore, describe dian system is extensive and includes convincing
how the above effects of locomotor activity on data demonstrating a role specifically in the feed-
the core molecular oscillator are transduced into back effects of locomotor activity. Neurotoxic and
long-term changes to the main output function electrolytic IGL lesions attenuate locomotor activ-
of the SCN, the phase of its electrical activity, or ity-induced behavioral phase shifts in both
indeed comment on information exchange in the hamsters and mice (Janik and Mrosovsky, 1994;
opposite direction. However, daily changes of Johnson et al., 1988; Koletar et al., 2011; Wickland
electrical activity in the SCN correlate with the and Turek, 1994). A potential confound for the
rest-activity cycle (Inouye and Kawamura, 1979) interpretation of these studies, however, is that
and an acute effect of locomotor activity on IGL lesions also reduce wheel-running, though this
SCN electrical firing in vivo has been reported. concern has been addressed: radio frequency
In vivo recordings of extracellular firing in the lesions of the IGL completely block entrainment
SCN of freely moving rats and hamsters show that of mice to scheduled forced treadmill running with
bouts of increased locomotor activity are followed no associated reduction in activity (Marchant et al.,
by acute reductions in SCN firing rate that 1997). In vivo electrical stimulation of the IGL
recover after an activity bout ceases (Meijer (Rusak et al., 1989) and NPY infusion into the third
et al., 1997; Schaap and Meijer, 2001; Yamazaki ventricle (Albers and Ferris, 1984; Biello and
et al., 1998). These suppressions are not induced Mrosovsky, 1996) both result in behavioral phase
by less vigorous activities such as feeding and shifts that resemble the traditional nonphotic
grooming (Schaap and Meijer, 2001) and the PRC seen in response to induced locomotor activ-
magnitude of suppressions correlates with the ity. Finally, mice lacking Dexras-1, an intracellular
intensity of wheel-running (Yamazaki et al., signaling component proposed to inhibit responses
1998). While further investigation will be required to NPY receptor activation in vivo, are reported to
to establish the causal nature of this interaction, exhibit larger phase shifts to novel wheel-induced
that is, whether locomotor activity suppresses exercise than WT mice (Cheng et al., 2004; Koletar
electrical activity or lower electrical activity et al., 2011), though others have failed to replicate
induces locomotor activity, reports that sup- this finding (Dallmann and Mrosovsky, 2007).
pressions of electrical activity follow increased Further evidence also exists to place the IGL and
locomotor activity indicate the former (Meijer NPY specifically downstream of locomotor activity
et al., 1997; Schaap and Meijer, 2001). This raises in phase-shifting processes. Confining hamsters to
the possibility that locomotor activity suppression a nest box, and hence preventing wheel-running,
of SCN electrical activity may, at least in part, fails to block phase advances in response to NPY
contribute to its phase-shifting effects. infusion to the third ventricle during the day and
318

NPY antiserum injections severely attenuate novel in mediating locomotor activity-induced phase-
wheel-running-induced phase advances without shifting effects on the clock, with the lesion study
reducing wheel revolutions during the novelty of Marchant et al. (1997) concluding that the IGL
pulse (Biello et al., 1994). Moreover, novelty- may even be essential for locomotor activity-
induced wheel-running at times when phase shifts mediated entrainment.
would occur activates NPY neurons in the IGL,
demonstrated by cFOS induction (Janik and
Mrosovsky, 1992; Janik et al., 1995; Webb et al., Serotonin and the raphe nuclei
2008; though see Zhang et al., 1993). Interestingly,
though the IGL also receives photic input due to its The monoamine serotonin (5-HT; 5-hydroxy-tryp-
location in the visual thalamus and is activated by tophan) has long been reported to function in the
phase-shifting light pulses, photically induced control of circadian responses to nonphotic stimuli
cFOS in the IGL is not found in NPY-positive though specific details of the role played by seroto-
neurons (Janik et al., 1995). Further, light nin in mediating nonphotic-like and, in particular,
attenuates phase shifts due to both exercise and locomotor activity-induced phase shifts are, despite
NPY injection during the day (Biello and 22 years of research, still unclear (Mistlberger and
Mrosovsky, 1995) and, in addition to effects Antle, 2011). This perhaps surprising lack of clarity
directly in the SCN, this integration of photic and arises due to a number of factors including the
nonphotic stimuli is likely to be mediated, at least complexity of serotonin receptor pharmacology,
in part, in the IGL as behavioral phase shifts differences in lesioning methods, and lesion extent
induced by activation of the IGL are blocked by between studies, species differences, and potential
light exposure (Challet et al., 1998). differences in the role of serotonin in mediating
Increased locomotor activity results in increased phase shifts resulting from locomotor activity
NPY levels in the SCN (Glass et al., 2010) and NPY induced in different ways (e.g., triazolam injection
alters circadian function through changing clock vs. novel wheel). That serotonin is involved in
gene expression and electrical activity in the cen- these processes in some capacity, however, is
tral oscillator. Infusion of NPY into the region of beyond doubt.
the SCN in vivo and directly onto SCN slices Serotonin is released in the hypothalamus in
in vitro suppresses per1 and per2 expression response to locomotor activity (Dudley et al.,
(Fukuhara et al., 2001; Maywood et al., 2002); an 1998) and serotonergic projections from the
effect mediated by NPY Y1/Y5 and Y2 receptors. median raphe (MR) of the brainstem innervate
NPY application to cultured SCN slices in vitro also the SCN, while the IGL is innervated by seroto-
induces nonphotic-like phase shifts in electrical nergic projections from the dorsal raphe (DR;
activity as well as acute suppression of SCN firing Abrahamson and Moore, 200l; Fig. 1). Further,
(Cutler et al., 1998; Gribkoff et al., 1998; Medanic circadian variation in serotonin content in the
and Gillette, 1993; Shibata and Moore, 1993; van SCN correlates with wheel revolutions performed
den Pol et al., 1996). The GABAergic nature of at different times of the day (Shioiri et al., 1991);
IGL neurons, including those that give rise to the a number of markers of serotonin function in the
GHT, further supports a role for the IGL/GHT in raphe nuclei are regulated by locomotor activity
nonphotic signaling, given, as mentioned earlier, (Greenwood et al., 2005; Malek et al., 2007) and
that the GABAA agonist muscimol also produces electrical stimulation of the raphe nuclei results
nonphotic phase-shifting responses when infused in serotonin release in the SCN, suppression of
centrally to hamsters (Smith et al., 1989). SCN cFOS expression, and behavioral phase
Collectively, the above studies indicate a cen- shifts (Dudley et al., 1999; Meyer-Bernstein and
tral role for the IGL, utilizing NPY and GABA, Morin, 1999).
319

Serotonin itself and a number of serotonin Taken together, these studies show that serotonin
agonists, effective variously against different receptor acts downstream of locomotor activity in mediating
subtypes (e.g., 8-hydroxy-2-di-n-propylaminotetralin nonphotic phase shifts. Finally, serotonin is a likely
(8-OH-DPAT), 5-carboxamidotryptamine, qui- candidate for involvement in nonphotic modula-
pazine), administered to SCN slices in vitro and tion of photic circadian control (Sanggaard et al.,
whole animals in vivo broadly produce typical 2003; reviewed in Yannielli and Harrington, 2004).
nonphotic style PRCs, with large phase advances Serotonin signaling also alters core clock gene
during the day and somewhat smaller phase expression and electrical activity in the SCN, fur-
delays at night, though not all studies report ther confirming that serotonin is functionally
the latter (Bobrzynska et al, 1996a; Challet involved in mediating nonphotic and locomotor-
et al., 1998; Edgar et al., 1993; Ehlen et al., mediated phase shifts and entrainment. 8-OH-
2001; Horikawa and Shibata, 2004; Lovenberg DPAT injection during the subjective day, the
et al., 1993; Medanic and Gillette, 1992; Prosser time at which this treatment produces phase
et al., 1990, 1993; Shibata et al., 1992; Tominaga advances, markedly suppresses per1 and per2
et al., 1992; though see Antle et al., 2003). More- mRNA levels (Duncan et al., 2005; Horikawa
over, suppressing endogenous serotonergic tone et al., 2000) and cotreatment with both 8-OH-
with short-term constant light exposure, which DPAT and brotizolam potentiates both behav-
creates a state of serotonin-hypersensitivity, ioral phase shifts and the suppression of per
strikingly potentiates phase shifts in response to (Yokota et al., 2000). Moreover, in addition to
a locomotor activity pulse (Knoch et al., 2004, phase-shifting SCN firing rate rhythms in vitro
2006; Landry and Mistlberger, 2005; Mistlberger (see above), serotonin also acutely suppresses
et al., 2002). Both 5-HT1a and 5-HT7 receptors, SCN electrical activity, both in vitro (Medanic
potentially acting in consort, are likely involved and Gillette, 1992) and in vivo (Mason, 1986;
as transgenic mice lacking either subtype fail to Nishino and Koizumi, 1977).
respond to 8-OH-DPAT injection (Gardani and Inconsistencies, however, do exist. While
Biello, 2008; Smith et al., 2008). wheel-running does induce modest increases in
Depletion of serotonin in the mouse SCN blocks cFOS expression in the DR, this is not in seroto-
the tau-shortening effect of a home-cage running nergic neurons (Webb et al., 2010). Neurotoxic
wheel (Mistlberger et al., 1998), and general hypo- lesions of serotonin neurons in the MR, which
thalamic serotonin depletion in hamster severely result in complete depletion of serotonin fibers
attenuates locomotor activity-mediated phase in the SCN, have been reported to block
shifts induced by triazolam injection, without a triazolam-induced but not novel wheel-induced
reduction in induced locomotor activity (Penev shifts (Meyer-Bernstein and Morin, 1998). More-
et al., 1995a). Further, and similar to data for over, other studies have found serotonin deple-
NPY, preventing hamsters from exercising after tion in the SCN and serotonin antagonist
injections of 8-OH-DPAT does not affect the injections to neither attenuate phase shifts
amplitude of phase shifts (Bobrzynska et al., resulting from novelty-induced wheel-running
1996a). Perhaps most importantly, disruption of (Antle et al., 1998; Bobrzynska et al., 1996b) nor
serotonin signaling impacts on stable entrainment alter the acceleration of reentrainment mediated
to locomotor activity feedback; lesioning serotonin by wheel-running (Smale et al., 1990). However,
terminals in and around the SCN blocks entrain- Bobrzynska et al. (1996b) report only 88–95%
ment of mice to scheduled voluntary exercise with- depletion of serotonin in and around the SCN
out reducing running (Edgar et al., 1997) and and the serotonergic DR projection to the IGL,
severely disrupts the ability of mice to entrain to and IGL projection to the SCN, would have been
forced treadmill running (Marchant et al., 1997). intact in this study and may account for the failure
320

to block shifts. Indeed, serotonin agonist injection phase shifts or part of the mechanism controlling
into the IGL during subjective day does induce them. However, demonstration of increased num-
phase advances (Challet et al., 1998). bers of orexin-positive neurons in the LHA during
Methodological, pharmacological, and species the “activity phase” of a forced daytime activity par-
differences notwithstanding, serotonin signaling adigm suggests that these neurons and the actions of
from the raphe nuclei, acting either directly or orexin function downstream of locomotor activity,
via the IGL appears, at least in part, to mediate potentially to mediate influences on the circadian
nonphotic-like phase shifts and entrainment to system (Salgado-Delgado et al., 2010).
scheduled exercise by suppressing clock gene
expression in the SCN.
Locomotor activity and glucocorticoid signaling

Orexin and the lateral hypothalamic area The hypothalamic–pituitary–adrenal (HPA) axis
and sympathico-adrenomedullary system are the
The lateral hypothalamic area (LHA) contains main neuroendocrine pathways that are activated
orexin-expressing neurons that are critical for in response to stressful stimuli and physical
appropriate control of arousal (Sakurai, 2007) challenges. Both are potently stimulated by loco-
and innervate a number of the previously discussed motor activity, resulting in extensive physiological
nuclei implicated in locomotor activity-mediated adaptations that are now recognized not as stress
entrainment (IGL, MR, and the periphery of the responses, but as normal responses to support
SCN; Cutler et al., 1999; Peyron et al., 1998; demanding behaviors (Beerling et al., 2011;
McGranaghan and Piggins, 2001; Fig. 1). Orexin Koolhaas et al., 2011). The contributions of these
neurons themselves receive indirect afferent locomotor activity-induced physiological responses
projections arising from the SCN (Abrahamson to phase shifting by, and entrainment to, exercise
and Moore, 2001; Deurveilher and Semba, 2005), have yet to be thoroughly investigated. A key HPA
suggesting reciprocal modulation between circa- response to locomotor activity is increased cir-
dian and arousal-promoting circuits and raising culating glucocorticoid levels (Stranahan et al.,
the possibility that orexin neurons in the LHA 2008) and glucocorticoids have been postulated to
may play a role in mediating the actions of locomo- play a role in mediating some activity-independent
tor activity on central oscillator mechanisms. nonphotic effects on the circadian system (Mis-
Indeed, orexin acutely alters both SCN and IGL tlberger et al., 2003; Sumova et al., 1994).
neuronal activity in vitro (Brown et al., 2008; A specific role in entrainment to locomotor activity
Farkas et al., 2002; Klisch et al., 2009; Pekala has not yet been determined however, but will be
et al., 2011), activation of orexin neurons in the required for a full understanding of the
LHA precedes the onset of the main activity bout mechanisms of entrainment to locomotor activity
(Marston et al., 2008), and electrical activity of feedback. Indeed, glucocorticoids play a key role
orexin neurons correlates with locomotor activity in maintaining internal synchrony between the var-
(Mileykovskiy et al., 2005). Further, running in a ious central and peripheral oscillators of the circa-
novel wheel and locomotor activity-promoting dian system (Dibner et al., 2010). It is likely that
dark pulses, which result in subjective day phase glucocorticoid signaling will function downstream
advances, induce cFOS expression in LHA orexin of SCN entrainment to locomotor activity to syn-
neurons while suppressing cFOS in the SCN chronize and entrain other oscillators; alternatively,
(Marston et al., 2008; Webb et al., 2008). It is not it remains possible that, given the direct induction
currently clear, however, whether this activation of of glucocorticoids by locomotor activity and
LHA orexin neurons is a consequence of nonphotic their action to synchronize peripheral oscillators,
321

glucocorticoids may function as part of a mechanism Intriguingly, the degus and Nile grass rat
for organizing peripheral oscillators to allow a (Arvicanthus ansorgei) exhibit diurnal activity
coordinated feedback from peripheral clocks to the when housed without a running wheel but a sub-
central circadian system (Dibner et al., 2010; set switch to nocturnal behavior when a wheel is
Stranahan et al., 2008). present (Kas and Edgar, 1999; Smale et al.,
2001). In the grass rat, this locomotor activity-
induced switch is associated with changes in the
Nonphotic entrainment in diurnal species time of increased NPY neuron activation such
that cFOS levels are always high in NPY neurons
Increasing numbers of studies demonstrating tem- in the IGL during an animal’s active phase
poral niche switching are blurring the distinction (Smale et al., 2001), regardless of whether this
between species traditionally considered in occupies a light or dark niche.
the laboratory to be “nocturnal” or “diurnal”
(reviewed in Kronfeld-Schor and Dayan, 2008).
Nevertheless, the vast majority of nonphotic Feedback of exercise to the human circadian
entrainment research to date, and indeed circa- system
dian research in general, has been carried out on
rodents that exhibit nocturnal activity patterns Regular physical exercise is held as the cornerstone
under normal laboratory conditions. Studies of for good physical and mental health and given the
species that exhibit diurnal activity under normal well-documented entraining effects of locomotor
laboratory conditions, though relatively few, have activity in animal models, exercise is attractive as
however been performed. a cheap and noninvasive intervention in circadian
Regardless of the temporal niche occupied by an misalignment disorders and maladies. However,
animal’s main active phase, central SCN pace- due to the difficulty of maintaining humans in a
maker phase is similar in all species studied; the controlled environment for long periods and in
SCN exhibits high transcriptional, electrical, and some studies a failure to adequately control lighting
metabolic activity during the subjective day and during exercise pulses at times when light levels
low at subjective night (e.g., Sato and Kawamura, should be low, reports of the phase responses of
1984; reviewed in Smale et al., 2003). Entrainment humans to exercise have been variable (reviewed
to scheduled locomotor activity has been reported in Atkinson et al., 2007). These difficulties notwith-
in diurnal European ground squirrels and common standing, a PRC to exercise for humans has been
marmosets (Glass et al., 2001; Hut et al., 1999) as postulated and generally resembles an archetypal
well as weak locomotor activity feedback effects “nonphotic” profile similar to that of both noctur-
in diurnal/crepuscular degus (Octodon degus; Kas nal and other diurnal species; delay shifts occur
and Edgar, 2001). The characteristics of nonphotic during the late night and advances are reported
entrainment in these “diurnal” mammals suggest during the late day and into the early evening
that the diurnal nonphotic PRC is similar in form (Buxton et al., 1997a,b, 2003; Mistlberger and
to that of nocturnal mammals; phase angle of Skene, 2005; Van Reeth et al., 1994).
entrainment in all cases agrees (as described in Sec- Complementary to these studies, “real-world”
tion “Phase angle of entrainment to scheduled attempts have been made to assess the impact
locomotor activity and the nonphotic PRC”) with of exercise on reentrainment following tra-
a mid-to-late day phase advance portion and a nsmeridian travel (Klein and Wegmann, 1974;
late-night delay portion. Nonphotic entrainment is Shiota et al., 1996). Both studies report accelerated
clearly determined by pacemaker phase and not reentrainment with exercise, though with the
by activity phase. caveat that neither study controlled for light
322

exposure or exercise intensity. Night shift studies despite no reduction in induced wheel-running
have reported more variable responses but similar (Van Reeth et al., 1992) and repeated injections
concerns over the control of a number of variables of triazolam fail to stably entrain behavioral
exist (e.g., Baehr et al., 1999; Eastman et al., 1995). rhythms in most old hamsters (Van Reeth et al.,
Nevertheless, reports of beneficial outcomes in a 1993). Dietary treatment of old hamsters with
real-world setting are encouraging and a role for the chronobiotic melatonin, however, improves
exercise in humans as a phase modulator under responses to triazolam-induced locomotor activity
an LD cycle would well suit the use of exercise to (Kolker et al., 2002). Further, novel wheel confine-
assist in reentrainment following transmeridian ment also fails to induce phase shifts in old hamsters,
travel. Studies of blind subjects, under controlled though it also induces less running in the wheel
conditions, performing defined duration and inten- than in young hamsters (Mrosovsky and Biello,
sity exercise protocols will be useful to determine 1994). In this study, increasing activity during novel
more precisely the parameters required for optimal wheel confinement to normal induced levels with
phase modulation by exercise in humans. Following either triazolam injection or positive motivation
additional research, this may also lead to the use of (the presence of an estrus female) recovered
exercise as a component of a combined nonphotic phase shifts. Together, these data indicate a reduced
treatment, along with sleep scheduling, social responsiveness of the aging circadian system to
interactions, and scheduled meal times to improve stimuli of this kind (though dark pulse-induced
synchronization with society in either blind or phase shifts are reported to be unaltered with age;
circadian-compromised individuals. Duncan and Deveraux, 2000). This may arise due
to reduction of one or more of the neurochemical
elements involved in mediating locomotor activity-
Locomotor activity feedback to the circadian induced circadian feedback effects. Indeed, seroto-
system in animals with altered circadian clocks nin levels in the brain have been reported to reduce
with age (Jagota and Kalyani, 2008) and both in vivo
Aging and in vitro responses to 8-OH-DPAT are reduced
as age increases (Biello, 2009; Penev et al., 1995b),
Age-related changes in circadian function in suggesting some age-related deficiency in serotonin
animals are well documented. Phase angle of receptor function may also be a factor. Consistent
entrainment to LD becomes more positive, with this hypothesis, many of the effects of ageing
variability in onsets increases, and behavioral on activity-induced entrainment events can be
rhythms fragment and lose amplitude (Weinert, simulated by monoamine depletion (as discussed
2000). SCN function also appears to weaken, with previously and reviewed in Turek et al., 1995).
lower amplitude oscillations in electrical function
(Biello, 2009; Nakamura et al., 2011; Watanabe
et al., 1995) and reduced responsiveness to photic Core clock gene mutations
stimuli (Kolker et al., 2003). Older animals per-
form fewer spontaneous revolutions in a home- Of the many core clock gene mutant strains that
cage running wheel (e.g., Penev et al., 1995b). exist, only two have been assessed for their
Changes in the circadian responses of aged responses to locomotor feedback to the circadian
animals to locomotor activity and neurochemicals system: the tau mutant hamster (Ralph and
involved in signaling these responses have also Menaker, 1988) and the clock mutant mouse
been reported. (Vitaterna et al., 1994).
Aging hamsters fail to show behavioral phase Both strains exhibit altered responses to locomo-
shifts following single injections of triazolam, tor activity feedback or associated neurochemical
323

treatments. Novel wheel confinement-induced loco- (Vipr2/) suffer severe behavioral circadian
motor activity produces a shifted and exaggerated deficits ranging from arrhythmicity to weak, inco-
PRC in tau mutant hamsters (Mrosovsky et al., herent, short period oscillations (Aton et al.,
1992) with behavioral responses to NPY and 2005; Ciarleglio et al., 2009; Hughes and Piggins,
8-OH-DPAT infusion showing similar changes 2008). These behavioral deficiencies are a mani-
(Biello and Mrosovsky, 1996; Colecchia et al., festation of reduced numbers of rhythmic neurons
1996). The enhanced responses of tau mutant in the SCN, reduced amplitude and coherence of
hamsters to a nonphotic stimulus are consistent with the remaining rhythmic neurons, and a lack of
the responses of this strain to photic zeitgebers, synchrony between their oscillations (Brown
where “type 0” phase resetting is seen (Shimomura et al., 2005, 2007; Ciarleglio et al., 2009; Cutler
and Menaker, 1994); clearly the circadian system of et al., 2003; Hughes et al., 2008; Maywood et al.,
tau mutants is more readily manipulated by both 2006). While both Vip/ and Vipr2/ mice dis-
external stimuli and behavioral feedback than WT play altered responses to photic zeitgebers
counterparts. In clock heterozygous mice, novel (Dragich et al., 2010; Hughes and Piggins, 2008;
wheel confinement results in phase delays during Hughes et al., 2004), Vipr2/ mice were recently
the subjective day in addition to lengthening circa- shown to manifest intact food anticipatory activity
dian period in subsequent free run (Challet et al., in response to the nonphotic stimuli of restricted
2000), describing altered responses to locomotor daytime feeding (Sheward et al., 2007). The
activity for this mutant strain also. responses of these strains to regularly scheduled
voluntary exercise were recently investigated
(Power et al., 2010).
Entrainment to scheduled locomotor activity in While the entrainment of WT mice to sched-
signaling mutants uled voluntary exercise is relatively slow, taking
 20–40 cycles to achieve stable entrainment,
To date, no rhythmic core clock gene mutant depending on the relative phases of endogenous
strain has been examined for entrainment to the activity and the opportunity to run (Edgar and
feedback signals associated with scheduled loco- Dement, 1991; Marchant and Mistlberger, 1996;
motor activity. While WT mice and hamsters, Power et al., 2010), both Vip/ and Vipr2/
and to a lesser extent WT rats, readily entrain to mice synchronize within just 2–4 days (Fig. 4). In
such stimuli (as discussed previously), mice ren- stark contrast to WT mice, which tend to show
dered arrhythmic through SCN lesion fail to small aftereffects on tau, following scheduled
respond (Marchant et al., 1997). However, this activity most Vipr2/ mice show a marked
lesion model lacks a core fundamental component improvement in rhythms with a proportion of pre-
of a functioning circadian system, the SCN itself. viously arrhythmic mice subsequently sustaining
As such, if the responses of the circadian system behavioral rhythms (Power et al., 2010). Post-
to scheduled locomotor activity are assumed to scheduled exercise Vipr2/ rhythms free-run in
be mediated by the SCN and employ the core phase with the previous opportunity to run,
molecular oscillator, then its lack of effectiveness demonstrating entrainment and not simply posi-
in this model is perhaps not surprising. One alter- tive masking, and tend to do so with a near 24-h
native model of disrupted clock function is mice period which can be sustained for at least a month
lacking either VIP (Colwell et al., 2003) or its cog- (Fig. 4). A number of unknowns, however, do
nate receptor in the SCN, VPAC2 (Harmar et al., remain. Few Vip/ mice show comparable
2002). VIP–VPAC2 signaling is critical for inter- improvements in rhythmicity, indicating either a
cellular communication within the SCN and as more profound rhythm dysfunction in this strain
such, mice lacking either VIP (Vip/) or VPAC2 than in Vipr2/ mice, or that VIP acting through
324

non-VPAC2 receptors is involved in the main- on cognition and adult neurogenesis in the hippo-
tenance of improved rhythms (Power et al., campus (e.g., Kempermann et al., 2010; van
2010). Further, the relative potency of locomotor Praag, 2008), coupled with the demonstration that
activity-associated feedback to act on the SCN circadian phase regulates the effects of exercise
and/or peripheral clocks in these mice (and on neurogenesis (Holmes et al., 2004), raises the
indeed in WTs) has yet to be examined. Indeed, possibility that locomotor activity may be impor-
the possibility of potentially enhanced peripheral tant for the resetting of hippocampal clocks.
feedback to the SCN in Vipr2/ mice to contrib- Again with consideration of the multioscillator
ute to the more potent effects of scheduled activ- nature of circadian biology, investigating a role
ity on the Vipr2/ circadian system is thus far for locomotor activity in ameliorating the effects
unexplored. Changes in the activity of the NPY/ of internal desynchrony arising from circadian
IGL and serotonin/raphe systems will also need disorders, shift work and transmeridian travel will
to be assessed in relation to the enhanced actions greatly enhance our current understanding of
of locomotor activity feedback in Vipr2/ mice. nonphotic influences on the integrated circadian
system. Indeed, as internal desynchrony gains
increasing prominence as a contributory factor
Future directions in circadian disruption, further investigation of
the effects of locomotor activity on peripheral
A number of key issues surrounding the feedback and central oscillators offers an opportunity to
influences of locomotor activity on the circadian both develop our understanding of basic circadian
system have yet to be addressed. Species general- biology and may also lead to the development of
ization continues to be an issue; the bulk of early a unique noninvasive, nonpharmacological, tool
work was carried out on hamsters, with rats for the reorganization of the distributed circadian
and mice very much underrepresented. This has system.
been redressed to some extent but, as with most Related to the above points, a number of pow-
research in the postgenomic age where mice have erful modern techniques are underrepresented in
become the species of choice, there is now a risk research into locomotor activity feedback to the
of overspecializing and becoming limited to this circadian system and could be effectively
alternative model. Further, the majority of early employed to improve our understanding of this
studies were performed before the discovery of process. For example, lesion studies crucial to
core clock genes and without a detailed knowl- our current understanding carry important
edge of the multioscillator nature of mammalian caveats. Nonspecific damage may be caused along
circadian systems. As such, many are largely the path taken into the brain when electrolytic
restricted to behavioral investigations and few lesions are performed and neurochemical lesions,
have considered either the potential contributions that destroy neurons rather than simply silencing
of extra-SCN oscillators to entrainment by loco- a gene of interest, may result in unwanted collat-
motor activity or the effects of locomotor activity eral damage or pleiotropic effects related to the
specifically on these extra-SCN oscillators. We neurons in question. These concerns could be
do not yet know the relative sensitivity of SCN addressed, should appropriate targeting agents
and extra-SCN oscillators to locomotor/arousal- become available, with the use of tissue-specific
associated nonphotic signals; this may be of par- conditional knockouts of key neurochemical
ticular interest for oscillators in tissues directly elements, for instance, NPY specifically in the
involved in locomotion and metabolism including IGL or serotonin in either or both of the raphe
cardiac muscle and adipose depots. Moreover, the nuclei. Similarly, the plethora of reporter mice
well-known positive effects of locomotor activity now available (e.g., Cheng et al., 2009; Kuhlman
325

et al., 2000; Wilsbacher et al., 2002; Yoo et al., equal magnitude to younger animals, coupled with
2004) will be crucial in addressing issues related the exaggerated responses of Vipr2/ mice to
to entrainment at the cellular and molecular scheduled activity make it possible that aged
levels, in particular in elucidating the effects of animals will respond well to scheduled locomotor
locomotor activity on synchrony between oscilla- activity, raising the possibility of a potential new
tor cells within rhythmic tissues and identifying noninvasive treatment for age-related circadian
the effects of locomotor activity feedback on neu- dysfunction.
rochemically defined neuronal populations.
Another key issue yet to be resolved is which
internal physiological correlates of locomotor Zeitgeber versus zeitnehmer?
activity are critical for entrainment. In particular,
the inadvertent temporal organization of food Traditionally, chronobiologists have used the term
intake associated with scheduled activity and acti- zeitgeber, literally meaning “time giver,” to describe
vation of the HPA axis by locomotion have yet to any external stimulus that is rhythmically encoun-
be systematically assessed for their relative con- tered and acts as an entraining signal to the circa-
tributions to the entraining effects of physical dian system. This term, and its implication of a
exercise. These, along with consideration of the one-way transfer of information from the environ-
importance of ultradian rhythmicity in glucocorti- ment to the circadian system, appropriately
coid signaling (Lightman et al., 2008) may provide describes independent external stimuli such as the
fruitful avenues for future investigation. LD cycle. However, the consideration of an inter-
Our own recent data (Power et al., 2010) dem- nally controlled function, such as locomotor activity,
onstrate that locomotor activity feedback is as an entraining stimulus raises questions over the
clearly a potent entraining stimulus in a mouse appropriateness of the term “zeitgeber” in a case
model with a “weakened” SCN pacemaker; such as this. To address this issue, the introduction
Vipr2/ mice. However, a key question remains of an alternative subclassification, “zeitnehmer” (lit-
unanswered: are the entraining properties of erally meaning “time taker”—in analogy with zeit-
exercise similarly boosted in other models of a geber), has been suggested for endogenous process
weakened circadian system? Aged animals and that are themselves under circadian control but that
circadian gene mutants such as afterhours, a also feedback to actively participate in the entrain-
mutation which alters tau and appears to weaken ment process (Lakin-Thomas, 2000; Merrow et al.,
circadian function (Godinho et al., 2007), would 2003; Roenneberg and Merrow, 1999; Roenneberg
be intriguing models with which to further probe et al., 1998). To date, this “zeitnehmer” term has
the effects of scheduled activity on a weakened mostly been adopted by researchers investigating
circadian system. Indeed, given the similarities in plant circadian biology (e.g., Hicks et al., 2001;
circadian deficit between Vipr2/ mice and aged McWatters et al., 2000; Thines and Harmon, 2010),
individuals (i.e., weakened SCN function (as though need not be restricted in this way and offers
discussed); aging-like alterations in circadian a more appropriate alternative to “zeitgeber” for
function induced by VIP suppression in young the description of locomotor activity feedback in
subjects (Gerhold et al., 2005)) and that VIP the context of entrainment. The adoption of
levels in the SCN are reduced with age (Duncan “zeitnehmer” in this case allows a clear distinction
et al., 2001), Vipr2/ mice may prove to be a use- to be drawn between independent geophysical
ful model of aging in the circadian system. The stimuli such as the LD cycle and endogenous, feed-
demonstration that aging laboratory animals, when back stimuli including locomotor activity. If one is
appropriately motivated, can be induced to gener- to adopt the term “zeitnehmer,” then as in nature,
ate locomotor activity-mediated phase shifts of when a range of external zeitgebers and internal
326

zeitnehmers will always be encountered in combina- References


tion, the use of both terms will always be necessary
when considering “real-world” entrainment. Abe, H., Honma, S., & Honma, K. (2007). Daily restricted feed-
ing resets the circadian clock in the suprachiasmatic nucleus
of CS mice. American Journal of Physiology. Regulatory,
Summary and conclusions Integrative and Comparative Physiology, 292, R607–R615.
Abe, H., Kida, M., Tsuji, K., & Mano, T. (1989). Feeding
Once thought of as simply an output, locomotor cycles entrain circadian rhythms of locomotor activity in
CS mice but not in C57BL/6J mice. Physiology & Behavior,
activity has for some time been known to also pro- 45, 397–401.
vide a feedback signal to the circadian system. Dis- Abe, H., Rusak, B., & Robertson, H. A. (1991). Photic induc-
crete pulses of locomotor activity can shift tion of Fos protein in the suprachiasmatic nucleus is
circadian phase or modify reentrainment to a shifted inhibited by the NMDA receptor antagonist MK-801. Neu-
LD cycle, while scheduled locomotor activity can roscience Letters, 127, 9–12.
Abrahamson, E. E., & Moore, R. Y. (2001). Suprachiasmatic
entrain circadian behavior and even induce lasting nucleus in the mouse: Retinal innervation, intrinsic organiza-
improvements in pacemaker function in circadian tion and efferent projections. Brain Research, 916, 172–191.
mutants. The precise nature of the feedback signal Albers, H. E., & Ferris, C. F. (1984). Neuropeptide Y: Role in
is unknown but is likely to be complex, involving a light-dark cycle entrainment of hamster circadian rhythms.
variety of physiological systems. Neurochemical pat- Neuroscience Letters, 50, 163–168.
Albrecht, U., Sun, Z. S., Eichele, G., & Lee, C. C. (1997).
hways in the brain that transmit this feedback infor- A differential response of two putative mammalian circadian
mation to the SCN are somewhat better understood, regulators, mper1 and mper2, to light. Cell, 91, 1055–1064.
however. Arousal-related signals from the DR are Angeles-Castellanos, M., Amaya, J. M., Salgado-Delgado, R.,
transmitted, via a serotonergic projection, to the Buijs, R. M., & Escobar, C. (2011). Scheduled food hastens
IGL which in turn innervates the SCN via the re-entrainment more than melatonin does after a 6-h phase
advance of the light-dark cycle in rats. Journal of Biological
GHT using GABA and NPY. Together with seroto- Rhythms, 26, 324–334.
nergic pathways from the MR and orexin signaling Antle, M. C., Marchant, E. G., Niel, L., & Mistlberger, R. E.
from the LHA, these locomotor/arousal-associated (1998). Serotonin antagonists do not attenuate activity-
pathways act to suppress SCN electrical activity induced phase shifts of circadian rhythms in the Syrian
and per gene expression, resulting in phase-depen- hamster. Brain Research, 813, 139–149.
Antle, M. C., & Mistlberger, R. E. (2000). Circadian clock
dent phase shifts of the master circadian pacemaker resetting by sleep deprivation without exercise in the Syrian
and hence the wider circadian system in general. hamster. The Journal of Neuroscience, 20, 9326–9332.
Though a number of key issues relating to the timing Antle, M. C., Ogilvie, M. D., Pickard, G. E., &
and doses of exercise required for effective manipu- Mistlberger, R. E. (2003). Response of the mouse circadian
lation of the human circadian system remain to be system to serotonin 1A/2/7 agonists in vivo: Surprisingly
little. Journal of Biological Rhythms, 18, 145–158.
addressed, feedback of locomotor activity to the cir- Antle, M. C., Tse, F., Koke, S. J., Sterniczuk, R., & Hagel, K.
cadian system has the potential to form part of novel (2008). Non-photic phase shifting of the circadian clock:
noninvasive interventions to a variety of circadian Role of the extracellular signal-responsive kinases I/II/
dysfunctions, including those associated with aging, mitogen-activated protein kinase pathway. The European
shift work, transmeridian travel, and the blind. Journal of Neuroscience, 28, 2511–2518.
Aschoff, J. (1979). Circadian rhythms: General features and
endocrinological aspects. In D. T. Kreiger (Ed.), Endocrine
Acknowledgments Rhythms (pp. 1–61). New York, NY: Raven.
Aschoff, J., Figala, J., & Poppel, E. (1973). Circadian rhythms
of locomotor activity in the golden hamster (Mesocricetus
The authors would like to thank Prof. Ralph
auratus) measured with two different techniques. Journal
Mistlberger for critical comments on an earlier draft of Comparative and Physiological Psychology, 85, 20–28.
of this review and acknowledge the BBRSC and Atkinson, G., Edwards, B., Reilly, T., & Waterhouse, J. (2007).
Wellcome Trust for financial support of their work. Exercise as a synchroniser of human circadian rhythms: An
327

update and discussion of the methodological problems. Euro- Boulos, Z., & Rusak, B. (1982). Circadian phase response cur-
pean Journal of Applied Physiology, 99, 331–341. ves for dark pulses in hamsters. Journal of Comparative
Aton, S. J., Colwell, C. S., Harmar, A. J., Waschek, J., & Physiology. A, 146, 411–417.
Herzog, E. D. (2005). Vasoactive intestinal polypeptide Brown, T. M., Colwell, C. S., Waschek, J. A., & Piggins, H. D.
mediates circadian rhythmicity and synchrony in mamma- (2007). Disrupted neuronal activity rhythms in the
lian clock neurons. Nature Neuroscience, 8, 476–483. suprachiasmatic nuclei of vasoactive intestinal polypeptide-
Baehr, E. K., Fogg, L. F., & Eastman, C. I. (1999). Intermit- deficient mice. Journal of Neurophysiology, 97, 2553–2558.
tent bright light and exercise to entrain human circadian Brown, T. M., Coogan, A. N., Cutler, D. J., Hughes, A. T., &
rhythms to night work. The American Journal of Physiology, Piggins, H. D. (2008). Electrophysiological actions of
277, R1598–R1604. orexins on rat suprachiasmatic neurons in vitro. Neurosci-
Beerling, W., Koolhaas, J. M., Ahnaou, A., Bouwknecht, J. A., ence Letters, 448, 273–278.
de Boer, S. F., Meerlo, P., et al. (2011). Physiological and Brown, T. M., Hughes, A. T., & Piggins, H. D. (2005). Gastrin-
hormonal responses to novelty exposure in rats are mainly releasing peptide promotes suprachiasmatic nuclei cellular
related to ongoing behavioral activity. Physiology & Behav- rhythmicity in the absence of vasoactive intestinal polypep-
ior, 103, 412–420. tide-VPAC2 receptor signaling. The Journal of Neurosci-
Belle, M. D., Diekman, C. O., Forger, D. B., & Piggins, H. D. ence, 25, 11155–11164.
(2009). Daily electrical silencing in the mammalian circadian Brown, T. M., & Piggins, H. D. (2007). Electrophysiology of
clock. Science, 326, 281–284. the suprachiasmatic circadian clock. Progress in Neurobiol-
Benus, R. F., Koolhaas, J. M., & van Oortmerssen, G. A. ogy, 82, 229–255.
(1988). Aggression and adaptation to the light-dark cycle: Buxton, O. M., Frank, S. A., L’Hermite-Baleriaux, M.,
Role of intrinsic and extrinsic control. Physiology & Behav- Leproult, R., Turek, F. W., & Van Cauter, E. (1997a). Roles
ior, 43, 131–137. of intensity and duration of nocturnal exercise in causing
Biello, S. M. (2009). Circadian clock resetting in the mouse phase delays of human circadian rhythms. The American
changes with age. Age (Dordrecht, Netherlands), 31, Journal of Physiology, 273, E536–E542.
293–303. Buxton, O. M., L’Hermite-Baleriaux, M., Hirschfeld, U., &
Biello, S. M., Golombek, D. A., & Harrington, M. E. (1997). Cauter, E. (1997b). Acute and delayed effects of exercise
Neuropeptide Y and glutamate block each other’s phase on human melatonin secretion. Journal of Biological
shifts in the suprachiasmatic nucleus in vitro. Neuroscience, Rhythms, 12, 568–574.
77, 1049–1057. Buxton, O. M., Lee, C. W., L’Hermite-Baleriaux, M.,
Biello, S. M., Janik, D., & Mrosovsky, N. (1994). Neuropep- Turek, F. W., & Van Cauter, E. (2003). Exercise elicits phase
tide Y and behaviorally induced phase shifts. Neuroscience, shifts and acute alterations of melatonin that vary with circa-
62, 273–279. dian phase. American Journal of Physiology. Regulatory, Inte-
Biello, S. M., & Mrosovsky, N. (1995). Blocking the phase- grative and Comparative Physiology, 284, R714–R724.
shifting effect of neuropeptide Y with light. Proceedings. Canal, M. M., & Piggins, H. D. (2006). Resetting of the ham-
Biological Sciences/The Royal Society, 259, 179–187. ster circadian system by dark pulses. American Journal of
Biello, S. M., & Mrosovsky, N. (1996). Phase response curves Physiology. Regulatory, Integrative and Comparative Physi-
to neuropeptide Y in wildtype and tau mutant hamsters. ology, 290, R785–R792.
Journal of Biological Rhythms, 11, 27–34. Canbeyli, R. S., Romero, M. T., & Silver, R. (1991). Neither
Bobrzynska, K. J., Godfrey, M. H., & Mrosovsky, N. (1996a). triazolam nor activity phase advance circadian locomotor
Serotonergic stimulation and nonphotic phase-shifting in activity in SCN-lesioned hamsters bearing fetal SCN
hamsters. Physiology & Behavior, 59, 221–230. transplants. Brain Research, 566, 40–45.
Bobrzynska, K. J., & Mrosovsky, N. (1998). Phase shifting by Castillo, M. R., Hochstetler, K. J., Tavernier, R. J., Jr.,
novelty-induced running: Activity dose-response curves at Greene, D. M., & Bult-Ito, A. (2004). Entrainment of the
different circadian times. Journal of Comparative Physiol- master circadian clock by scheduled feeding. American Jour-
ogy. A, 182, 251–258. nal of Physiology. Regulatory, Integrative and Comparative
Bobrzynska, K. J., Vrang, N., & Mrosovsky, N. (1996b). Per- Physiology, 287, R551–R555.
sistence of nonphotic phase shifts in hamsters after seroto- Challet, E. (2007). Minireview: Entrainment of the
nin depletion in the suprachiasmatic nucleus. Brain suprachiasmatic clockwork in diurnal and nocturnal
Research, 741, 205–214. mammals. Endocrinology, 148, 5648–5655.
Boulos, Z., Rosenwasser, A. M., & Terman, M. (1980). Feed- Challet, E., Caldelas, I., Graff, C., & Pevet, P. (2003). Synchroni-
ing schedules and the circadian organization of behavior in zation of the molecular clockwork by light- and food-related
the rat. Behavioural Brain Research, 1, 39–65. cues in mammals. Biological Chemistry, 384, 711–719.
328

Challet, E., Scarbrough, K., Penev, P. D., & Turek, F. W. Dallmann, R., Lemm, G., & Mrosovsky, N. (2007). Toward
(1998). Roles of suprachiasmatic nuclei and intergeniculate easier methods of studying nonphotic behavioral entrain-
leaflets in mediating the phase-shifting effects of a seroto- ment in mice. Journal of Biological Rhythms, 22, 458–461.
nergic agonist and their photic modulation during subjective Dallmann, R., & Mrosovsky, N. (2007). Non-photic phase
day. Journal of Biological Rhythms, 13, 410–421. resetting of Dexras1 deficient mice: A more complicated
Challet, E., Takahashi, J. S., & Turek, F. W. (2000). Nonphotic story. Behavioural Brain Research, 180, 197–202.
phase-shifting in clock mutant mice. Brain Research, 859, Damiola, F., Le Minh, N., Preitner, N., Kornmann, B., Fleury-
398–403. Olela, F., & Schibler, U. (2000). Restricted feeding
Cheng, H. Y., Alvarez-Saavedra, M., Dziema, H., Choi, Y. S., uncouples circadian oscillators in peripheral tissues from
Li, A., & Obrietan, K. (2009). Segregation of expression of the central pacemaker in the suprachiasmatic nucleus. Genes
mPeriod gene homologs in neurons and glia: Possible diver- & Development, 14, 2950–2961.
gent roles of mPeriod1 and mPeriod2 in the brain. Human Davidson, A. J. (2006). Search for the feeding-entrainable cir-
Molecular Genetics, 18, 3110–3124. cadian oscillator: A complex proposition. American Journal
Cheng, H. Y., Obrietan, K., Cain, S. W., Lee, B. Y., of Physiology. Regulatory, Integrative and Comparative
Agostino, P. V., Joza, N. A., et al. (2004). Dexras1 Physiology, 290, R1524–R1526.
potentiates photic and suppresses nonphotic responses of DeCoursey, P. J., Walker, J. K., & Smith, S. A. (2000).
the circadian clock. Neuron, 43, 715–728. A circadian pacemaker in free-living chipmunks: Essential
Ciarleglio, C. M., Gamble, K. L., Axley, J. C., Strauss, B. R., for survival? Journal of Comparative Physiology. A, 186,
Cohen, J. Y., Colwell, C. S., et al. (2009). Population 169–180.
encoding by circadian clock neurons organizes circadian Deurveilher, S., & Semba, K. (2005). Indirect projections from
behavior. The Journal of Neuroscience, 29, 1670–1676. the suprachiasmatic nucleus to major arousal-promoting cell
Colecchia, E. F., Penev, P. D., Zee, P. C., & Turek, F. W. groups in rat: Implications for the circadian control of
(1996). Phase-shifting effects of a serotonin agonist in tau behavioural state. Neuroscience, 130, 165–183.
mutant hamsters. Brain Research, 730, 227–231. Dibner, C., Schibler, U., & Albrecht, U. (2010). The mamma-
Colwell, C. S. (2011). Linking neural activity and molecular lian circadian timing system: Organization and coordination
oscillations in the SCN. Nature Reviews. Neuroscience, 12, of central and peripheral clocks. Annual Review of Physiol-
553–569. ogy, 72, 517–549.
Colwell, C. S., Michel, S., Itri, J., Rodriguez, W., Tam, J., Doyle, S., & Menaker, M. (2007). Circadian photoreception in
Lelievre, V., et al. (2003). Disrupted circadian rhythms in vertebrates. Cold Spring Harbor Symposia on Quantitative
VIP- and PHI-deficient mice. American Journal of Physiol- Biology, 72, 499–508.
ogy—Regulatory, integrative and comparative physiology, Dragich, J. M., Loh, D. H., Wang, L. M., Vosko, A. M.,
285, R939–R949. Kudo, T., Nakamura, T. J., et al. (2010). The role of the
Coogan, A. N., & Piggins, H. D. (2005). Dark pulse suppres- neuropeptides PACAP and VIP in the photic regulation of
sion of P-ERK and c-Fos in the hamster suprachiasmatic gene expression in the suprachiasmatic nucleus. The Euro-
nuclei. The European Journal of Neuroscience, 22, 158–168. pean Journal of Neuroscience, 31(5), 864–875.
Cutler, D. J., Haraura, M., Reed, H. E., Shen, S., Sheward, W. J., Dudley, T. E., DiNardo, L. A., & Glass, J. D. (1998). Endoge-
Morrison, C. F., et al. (2003). The mouse VPAC(2) receptor nous regulation of serotonin release in the hamster
confers suprachiasmatic nuclei cellular rhythmicity and suprachiasmatic nucleus. The Journal of Neuroscience, 18,
responsiveness to vasoactive intestinal polypeptide in vitro. 5045–5052.
The European Journal of Neuroscience, 17, 197–204. Dudley, T. E., Dinardo, L. A., & Glass, J. D. (1999). In vivo
Cutler, D. J., Morris, R., Sheridhar, V., Wattam, T. A., assessment of the midbrain raphe nuclear regulation of sero-
Holmes, S., Patel, S., et al. (1999). Differential distribution tonin release in the hamster suprachiasmatic nucleus. Jour-
of orexin-A and orexin-B immunoreactivity in the rat brain nal of Neurophysiology, 81, 1469–1477.
and spinal cord. Peptides, 20, 1455–1470. Duncan, M. J., & Deveraux, A. W. (2000). Age-related
Cutler, D. J., Piggins, H. D., Selbie, L. A., & Mason, R. (1998). changes in circadian responses to dark pulses. American
Responses to neuropeptide Y in adult hamster Journal of Physiology. Regulatory, Integrative and Compara-
suprachiasmatic nucleus neurones in vitro. European Journal tive Physiology, 279, R586–R590.
of Pharmacology, 345, 155–162. Duncan, M. J., Franklin, K. M., Davis, V. A.,
Daan, S., & Pittendrigh, C. S. (1976). Functional analysis of Grossman, G. H., Knoch, M. E., & Glass, J. D. (2005).
circadian pacemakers in nocturnal rodents. 2. Variability of Short-term constant light potentiation of large-magnitude
phase response curves. Journal of Comparative Physiology. circadian phase shifts induced by 8-OH-DPAT: Effects on
A, 106, 253–266. serotonin receptors and gene expression in the hamster
329

suprachiasmatic nucleus. The European Journal of Neurosci- Gardani, M., & Biello, S. M. (2008). The effects of photic and
ence, 22, 2306–2314. nonphotic stimuli in the 5-HT7 receptor knockout mouse.
Duncan, M. J., Herron, J. M., & Hill, S. A. (2001). Aging selec- Neuroscience, 152, 245–253.
tively suppresses vasoactive intestinal peptide messenger RNA Gerhold, L. M., Rosewell, K. L., & Wise, P. M. (2005). Sup-
expression in the suprachiasmatic nucleus of the Syrian hamster. pression of vasoactive intestinal polypeptide in the
Brain Research. Molecular Brain Research, 87, 196–203. suprachiasmatic nucleus leads to aging-like alterations in
Eastman, C. I., Hoese, E. K., Youngstedt, S. D., & Liu, L. cAMP rhythms and activation of gonadotropin-releasing
(1995). Phase-shifting human circadian rhythms with exer- hormone neurons. The Journal of Neuroscience, 25, 62–67.
cise during the night shift. Physiology & Behavior, 58, Glass, J. D., Guinn, J., Kaur, G., & Francl, J. M. (2010). On the
1287–1291. intrinsic regulation of neuropeptide Y release in the mam-
Ebling, F. J., Maywood, E. S., Staley, K., Humby, T., malian suprachiasmatic nucleus circadian clock. The Euro-
Hancock, D. C., Waters, C. M., et al. (1991). The role of pean Journal of Neuroscience, 31, 1117–1126.
N-methyl-d-aspartate-type glutamatergic neurotransmission Glass, J. D., Tardif, S. D., Clements, R., & Mrosovsky, N.
in the photic induction of immediate-early gene expression (2001). Photic and nonphotic circadian phase resetting in a
in the suprachiasmatic nuclei of the Syrian hamster. Journal diurnal primate, the common marmoset. American Journal
of Neuroendocrinology, 3, 641–652. of Physiology. Regulatory, Integrative and Comparative
Edgar, D. M., & Dement, W. C. (1991). Regularly scheduled Physiology, 280, R191–R197.
voluntary exercise synchronizes the mouse circadian clock. Glossop, N. R. (2011). Circadian timekeeping in Drosophila
The American Journal of Physiology, 261, R928–R933. melanogaster and Mus musculus. Essays in Biochemistry,
Edgar, D. M., Martin, C. E., & Dement, W. C. (1991). Activity 49, 19–35.
feedback to the mammalian circadian pacemaker: Influence Godinho, S. I., Maywood, E. S., Shaw, L., Tucci, V.,
on observed measures of rhythm period length. Journal of Barnard, A. R., Busino, L., et al. (2007). The after-hours
Biological Rhythms, 6, 185–199. mutant reveals a role for Fbxl3 in determining mammalian
Edgar, D. M., Miller, J. D., Prosser, R. A., Dean, R. R., & circadian period. Science, 316, 897–900.
Dement, W. C. (1993). Serotonin and the mammalian circadian Golombek, D. A., & Ralph, M. R. (1995). Circadian responses
system: II. Phase-shifting rat behavioral rhythms with seroto- to light: The calmodulin connection. Neuroscience Letters,
nergic agonists. Journal of Biological Rhythms, 8, 17–31. 192, 101–104.
Edgar, D. M., Reid, M. S., & Dement, W. C. (1997). Seroto- Golombek, D. A., & Rosenstein, R. E. (2010). Physiology of
nergic afferents mediate activity-dependent entrainment of circadian entrainment. Physiological Reviews, 90, 1063–1102.
the mouse circadian clock. The American Journal of Physi- Goz, D., Studholme, K., Lappi, D. A., Rollag, M. D.,
ology, 273, R265–R269. Provencio, I., & Morin, L. P. (2008). Targeted destruction
Ehlen, J. C., Grossman, G. H., & Glass, J. D. (2001). In vivo of photosensitive retinal ganglion cells with a saporin conju-
resetting of the hamster circadian clock by 5-HT7 receptors gate alters the effects of light on mouse circadian rhythms.
in the suprachiasmatic nucleus. The Journal of Neuroscience, PLoS One, 3, e3153.
21, 5351–5357. Greenwood, B. N., Foley, T. E., Day, H. E., Burhans, D.,
Ellis, G. B., McKlveen, R. E., & Turek, F. W. (1982). Dark Brooks, L., Campeau, S., et al. (2005). Wheel running alters
pulses affect the circadian rhythm of activity in hamsters serotonin (5-HT) transporter, 5-HT1A, 5-HT1B, and alpha
kept in constant light. The American Journal of Physiology, 1b-adrenergic receptor mRNA in the rat raphe nuclei.
242, R44–R50. Biological Psychiatry, 57, 559–568.
Farkas, B., Vilagi, I., & Detari, L. (2002). Effect of orexin-A Gribkoff, V. K., Pieschl, R. L., Wisialowski, T. A., van den
on discharge rate of rat suprachiasmatic nucleus neurons Pol, A. N., & Yocca, F. D. (1998). Phase shifting of circadian
in vitro. Acta Biologica Hungarica, 53, 435–443. rhythms and depression of neuronal activity in the rat
Fukuhara, C., Brewer, J. M., Dirden, J. C., Bittman, E. L., suprachiasmatic nucleus by neuropeptide Y: Mediation by
Tosini, G., & Harrington, M. E. (2001). Neuropeptide Y rap- different receptor subtypes. The Journal of Neuroscience,
idly reduces Period 1 and Period 2 mRNA levels in the hamster 18, 3014–3022.
suprachiasmatic nucleus. Neuroscience Letters, 314, 119–122. Guilding, C., & Piggins, H. D. (2007). Challenging the omnip-
Fuller, P. M., Lu, J., & Saper, C. B. (2009). Standards of evi- otence of the suprachiasmatic timekeeper: Are circadian
dence in chronobiology: A response. Journal of Circadian oscillators present throughout the mammalian brain? The
Rhythms, 7, 9. European Journal of Neuroscience, 25, 3195–3216.
Gannon, R. L., & Rea, M. A. (1995). Twelve-hour phase shifts Gwinner, E. (1966). Periodicity of a circadian rhythm in birds
of hamster circadian rhythms elicited by voluntary wheel by species-specific song cycles (Aves, Fringillidae: Carduelis
running. Journal of Biological Rhythms, 10, 196–210. spinus, Serinus serinus). Experientia, 22, 765–766.
330

Hamada, T., Antle, M. C., & Silver, R. (2004). The role of the suprachiasmatic circadian pacemaker of mice lacking
Period1 in non-photic resetting of the hamster circadian the VPAC(2) receptor. The Journal of Neuroscience, 24,
pacemaker in the suprachiasmatic nucleus. Neuroscience 3522–3526.
Letters, 362, 87–90. Hughes, A. T., Guilding, C., Lennox, L., Samuels, R. E.,
Hannibal, J., & Fahrenkrug, J. (2006). Neuronal input McMahon, D. G., & Piggins, H. D. (2008). Live imaging
pathways to the brain’s biological clock and their functional of altered period1 expression in the suprachiasmatic nuclei
significance. Advances in Anatomy, Embryology, and Cell of Vipr2/ mice. Journal of Neurochemistry, 106,
Biology, 182, 1–71. 1646–1657.
Harmar, A. J., Marston, H. M., Shen, S. B., Spratt, C., Hughes, A. T., & Piggins, H. D. (2008). Behavioral responses
West, K. M., Sheward, W. J., et al. (2002). The VPAC(2) of Vipr2/ mice to light. Journal of Biological Rhythms,
receptor is essential for circadian function in the mouse 23, 211–219.
suprachiasmatic nuclei. Cell, 109, 497–508. Hut, R. A., Mrosovsky, N., & Daan, S. (1999). Nonphotic
Harrington, M. E. (1997). The ventral lateral geniculate entrainment in a diurnal mammal, the European ground
nucleus and the intergeniculate leaflet: Interrelated squirrel (Spermophilus citellus). Journal of Biological
structures in the visual and circadian systems. Neuroscience Rhythms, 14, 409–419.
and Biobehavioral Reviews, 21, 705–727. Inouye, S. T., & Kawamura, H. (1979). Persistence of circadian
Hastings, M. H., Duffield, G. E., Smith, E. J., Maywood, E. S., rhythmicity in a mammalian hypothalamic “island” con-
& Ebling, F. J. (1998). Entrainment of the circadian system taining the suprachiasmatic nucleus. Proceedings of the
of mammals by nonphotic cues. Chronobiology Interna- National Academy of Sciences of the United States of
tional, 15, 425–445. America, 76, 5962–5966.
Hastings, M. H., Maywood, E. S., & O’Neill, J. S. (2008). Cel- Jagota, A., & Kalyani, D. (2008). Daily serotonin rhythms in
lular circadian pacemaking and the role of cytosolic rat brain during postnatal development and aging. Bio-
rhythms. Current Biology, 18, R805–R815. gerontology, 9, 229–234.
Hattar, S., Lucas, R. J., Mrosovsky, N., Thompson, S., Janik, D., Mikkelsen, J. D., & Mrosovsky, N. (1995). Cellular
Douglas, R. H., Hankins, M. W., et al. (2003). Melanopsin colocalization of Fos and neuropeptide Y in the inter-
and rod-cone photoreceptive systems account for all major geniculate leaflet after nonphotic phase-shifting events.
accessory visual functions in mice. Nature, 424, 76–81. Brain Research, 698, 137–145.
Hicks, K. A., Albertson, T. M., & Wagner, D. R. (2001). EARLY Janik, D., & Mrosovsky, N. (1992). Gene expression in the
FLOWERING3 encodes a novel protein that regulates circa- geniculate induced by a nonphotic circadian phase shifting
dian clock function and flowering in Arabidopsis. The Plant stimulus. Neuroreport, 3, 575–578.
Cell, 13, 1281–1292. Janik, D., & Mrosovsky, N. (1993). Nonphotically induced
Holmes, M. M., Galea, L. A., Mistlberger, R. E., & phase shifts of circadian rhythms in the golden hamster:
Kempermann, G. (2004). Adult hippocampal neurogenesis Activity-response curves at different ambient temperatures.
and voluntary running activity: Circadian and dose- Physiology & Behavior, 53, 431–436.
dependent effects. Journal of Neuroscience Research, 76, Janik, D., & Mrosovsky, N. (1994). Intergeniculate leaflet
216–222. lesions and behaviorally-induced shifts of circadian rhythms.
Honma, K., von Goetz, C., & Aschoff, J. (1983). Effects of Brain Research, 651, 174–182.
restricted daily feeding on freerunning circadian rhythms Johnson, R. F., Smale, L., Moore, R. Y., & Morin, L. P. (1988).
in rats. Physiology & Behavior, 30, 905–913. Lateral geniculate lesions block circadian phase-shift
Honrado, G. I., & Mrosovsky, N. (1989). Arousal by sexual responses to a benzodiazepine. Proceedings of the National
stimuli accelerates the re-entrainment of hamsters to phase Academy of Sciences of the United States of America, 85,
advanced light-dark cycles. Behavioral Ecology and Sociobi- 5301–5304.
ology, 25, 57–63. Joy, J. E., & Turek, F. W. (1992). Combined effects on the cir-
Horikawa, K., & Shibata, S. (2004). Phase-resetting response cadian clock of agents with different phase response curves:
to (þ)8-OH-DPAT, a serotonin 1A/7 receptor agonist, in Phase-shifting effects of triazolam and light. Journal of
the mouse in vivo. Neuroscience Letters, 368, 130–134. Biological Rhythms, 7, 51–63.
Horikawa, K., Yokota, S., Fuji, K., Akiyama, M., Moriya, T., Kalsbeek, A., Palm, I. F., La Fleur, S. E., Scheer, F. A.,
Okamura, H., et al. (2000). Nonphotic entrainment by Perreau-Lenz, S., Ruiter, M., et al. (2006). SCN outputs
5-HT1A/7 receptor agonists accompanied by reduced Per1 and the hypothalamic balance of life. Journal of Biological
and Per2 mRNA levels in the suprachiasmatic nuclei. The Rhythms, 21, 458–469.
Journal of Neuroscience, 20, 5867–5873. Kas, M. J., & Edgar, D. M. (1999). A nonphotic stimulus
Hughes, A. T., Fahey, B., Cutler, D. J., Coogan, A. N., & inverts the diurnal-nocturnal phase preference in Octodon
Piggins, H. D. (2004). Aberrant gating of photic input to degus. The Journal of Neuroscience, 19, 328–333.
331

Kas, M. J., & Edgar, D. M. (2001). Scheduled voluntary wheel Kuhlman, S. J., Quintero, J. E., & McMahon, D. G. (2000).
running activity modulates free-running circadian body tem- GFP fluorescence reports Period 1 circadian gene regulation
perature rhythms in Octodon degus. Journal of Biological in the mammalian biological clock. Neuroreport, 11,
Rhythms, 16, 66–75. 1479–1482.
Kempermann, G., Fabel, K., Ehninger, D., Babu, H., Leal- Laemle, L. K., & Ottenweller, J. E. (1999). Nonphotic entrain-
Galicia, P., Garthe, A., et al. (2010). Why and how physical ment of activity and temperature rhythms in anophthalmic
activity promotes experience-induced brain plasticity. mice. Physiology & Behavior, 66, 461–471.
Frontiers in Neuroscience, 4, 189. Lakin-Thomas, P. L. (2000). Circadian rhythms: New functions
Klein, K. E., & Wegmann, H. M. (1974). The resynchronization for old clock genes. Trends in Genetics, 16, 135–142.
of human circadian rhythms after transmeridian flights as a Landry, G. J., & Mistlberger, R. E. (2005). Differential effects
result of flight direction and mode of activity. In L. E. of constant light on circadian clock resetting by photic and
Scheving, F. Halberg & J. E. Pauly (Eds.), Chronobiology nonphotic stimuli in Syrian hamsters. Brain Research, 1059,
(pp. 564–570). Tokyo: Igku Shoin. 52–58.
Klisch, C., Inyushkin, A., Mordel, J., Karnas, D., Pevet, P., & Li, X., Gilbert, J., & Davis, F. C. (2005). Disruption of masking
Meissl, H. (2009). Orexin A modulates neuronal activity of by hypothalamic lesions in Syrian hamsters. Journal of Com-
the rodent suprachiasmatic nucleus in vitro. The European parative Physiology A, 191, 23–30.
Journal of Neuroscience, 30, 65–75. Lightman, S. L., Wiles, C. C., Atkinson, H. C., Henley, D. E.,
Knoch, M. E., Gobes, S. M., Pavlovska, I., Su, C., Russell, G. M., Leendertz, J. A., et al. (2008). The signifi-
Mistlberger, R. E., & Glass, J. D. (2004). Short-term expo- cance of glucocorticoid pulsatility. European Journal of
sure to constant light promotes strong circadian phase- Pharmacology, 583, 255–262.
resetting responses to nonphotic stimuli in Syrian hamsters. Lovenberg, T. W., Baron, B. M., de Lecea, L., Miller, J. D.,
The European Journal of Neuroscience, 19, 2779–2790. Prosser, R. A., Rea, M. A., et al. (1993). A novel adenylyl
Knoch, M. E., Siegel, D., Duncan, M. J., & Glass, J. D. (2006). cyclase-activating serotonin receptor (5-HT7) implicated in
Serotonergic mediation of constant light-potentiated non- the regulation of mammalian circadian rhythms. Neuron,
photic phase shifting of the circadian locomotor activity 11, 449–458.
rhythm in Syrian hamsters. American Journal of Physiology. Malek, Z. S., Sage, D., Pevet, P., & Raison, S. (2007). Daily
Regulatory, Integrative and Comparative Physiology, 291, rhythm of tryptophan hydroxylase-2 messenger ribonucleic
R180–R188. acid within raphe neurons is induced by corticoid daily surge
Ko, C. H., & Takahashi, J. S. (2006). Molecular components of and modulated by enhanced locomotor activity. Endocrinol-
the mammalian circadian clock. Human Molecular Genetics, ogy, 148, 5165–5172.
15(Spec No 2), R271–R277. Marchant, E. G., & Mistlberger, R. E. (1995). Morphine
Koletar, M. M., Cheng, H. Y., Penninger, J. M., & phase-shifts circadian rhythms in mice: Role of behavioural
Ralph, M. R. (2011). Loss of dexras1 alters nonphotic circa- activation. Neuroreport, 7, 209–212.
dian phase shifts and reveals a role for the intergeniculate Marchant, E. G., & Mistlberger, R. E. (1996). Entrainment
leaflet (IGL) in gene-targeted mice. Chronobiology Interna- and phase shifting of circadian rhythms in mice by forced
tional, 28, 553–562. treadmill running. Physiology & Behavior, 60, 657–663.
Kolker, D. E., Fukuyama, H., Huang, D. S., Takahashi, J. S., Marchant, E. G., Watson, N. V., & Mistlberger, R. E. (1997).
Horton, T. H., & Turek, F. W. (2003). Aging alters circadian Both neuropeptide Y and serotonin are necessary for
and light-induced expression of clock genes in golden entrainment of circadian rhythms in mice by daily treadmill
hamsters. Journal of Biological Rhythms, 18, 159–169. running schedules. The Journal of Neuroscience, 17,
Kolker, D. E., Losee Olson, S., Dutton-Boilek, J., 7974–7987.
Bennett, K. M., Wallen, E. P., Horton, T. H., et al. (2002). Marston, O. J., Williams, R. H., Canal, M. M., Samuels, R. E.,
Feeding melatonin enhances the phase shifting response to Upton, N., & Piggins, H. D. (2008). Circadian and dark-
triazolam in both young and old golden hamsters. American pulse activation of orexin/hypocretin neurons. Molecular
Journal of Physiology. Regulatory, Integrative and Compara- Brain, 1, 19.
tive Physiology, 282, R1382–R1388. Mason, R. (1986). Circadian variation in sensitivity of
Koolhaas, J. M., Bartolomucci, A., Buwalda, B., de Boer, S. F., suprachiasmatic and lateral geniculate neurones to 5-
Flugge, G., Korte, S. M., et al. (2011). Stress revisited: A hydroxytryptamine in the rat. The Journal of Physiology,
critical evaluation of the stress concept. Neuroscience and 377, 1–13.
Biobehavioral Reviews, 35, 1291–1301. Maywood, E. S., & Mrosovsky, N. (2001). A molecular
Kronfeld-Schor, N., & Dayan, T. (2008). Activity patterns of explanation of interactions between photic and non-photic
rodents: The physiological ecology of biological rhythms. circadian clock-resetting stimuli. Brain Research. Gene
Biological Rhythm Research, 39, 193–211. Expression Patterns, 1, 27–31.
332

Maywood, E. S., Mrosovsky, N., Field, M. D., & Mileykovskiy, B. Y., Kiyashchenko, L. I., & Siegel, J. M.
Hastings, M. H. (1999). Rapid down-regulation of mamma- (2005). Behavioral correlates of activity in identified
lian period genes during behavioral resetting of the circa- hypocretin/orexin neurons. Neuron, 46, 787–798.
dian clock. Proceedings of the National Academy of Mintz, E. M., & Albers, H. E. (1997). Microinjection of
Sciences of the United States of America, 96, 15211–15216. NMDA into the SCN region mimics the phase shifting effect
Maywood, E. S., Okamura, H., & Hastings, M. H. (2002). of light in hamsters. Brain Research, 758, 245–249.
Opposing actions of neuropeptide Y and light on the Mistlberger, R. E. (1991a). Effects of daily schedules of forced
expression of circadian clock genes in the mouse suprachias- activity on free-running rhythms in the rat. Journal of
matic nuclei. The European Journal of Neuroscience, 15, Biological Rhythms, 6, 71–80.
216–220. Mistlberger, R. E. (1991b). Scheduled daily exercise or feeding
Maywood, E. S., Reddy, A. B., Wong, G. K. Y., O’Neill, J. S., alters the phase of photic entrainment in Syrian hamsters.
O’Brien, J. A., McMahon, D. G., et al. (2006). Synchroniza- Physiology & Behavior, 50, 1257–1260.
tion and maintenance of timekeeping in suprachiasmatic Mistlberger, R. E., & Antle, M. C. (2011). Entrainment of cir-
circadian clock cells by neuropeptidergic signaling. Current cadian clocks in mammals by arousal and food. Essays in
Biology, 16, 599–605. Biochemistry, 49, 119–136.
McGranaghan, P. A., & Piggins, H. D. (2001). Orexin A-like Mistlberger, R. E., Antle, M. C., Webb, I. C., Jones, M.,
immunoreactivity in the hypothalamus and thalamus of the Weinberg, J., & Pollock, M. S. (2003). Circadian clock
Syrian hamster (Mesocricetus auratus) and Siberian hamster resetting by arousal in Syrian hamsters: The role of stress
(Phodopus sungorus), with special reference to circadian and activity. American Journal of Physiology. Regulatory,
structures. Brain Research, 904, 234–244. Integrative and Comparative Physiology, 285, R917–R925.
McWatters, H. G., Bastow, R. M., Hall, A., & Millar, A. J. Mistlberger, R. E., Belcourt, J., & Antle, M. C. (2002). Circa-
(2000). The ELF3 zeitnehmer regulates light signalling to dian clock resetting by sleep deprivation without exercise in
the circadian clock. Nature, 408, 716–720. Syrian hamsters: Dark pulses revisited. Journal of Biological
Medanic, M., & Gillette, M. U. (1992). Serotonin regulates the Rhythms, 17, 227–237.
phase of the rat suprachiasmatic circadian pacemaker Mistlberger, R. E., Bossert, J. M., Holmes, M. M., &
in vitro only during the subjective day. The Journal of Phys- Marchant, E. G. (1998). Serotonin and feedback effects of
iology, 450, 629–642. behavioral activity on circadian rhythms in mice. Behav-
Medanic, M., & Gillette, M. U. (1993). Suprachiasmatic circa- ioural Brain Research, 96, 93–99.
dian pacemaker of rat shows two windows of sensitivity to Mistlberger, R. E., Buijs, R. M., Challet, E., Escobar, C.,
neuropeptide Y in vitro. Brain Research, 620, 281–286. Landry, G. J., Kalsbeek, A., et al. (2009). Standards of evi-
Meijer, J. H., Schaap, J., Watanabe, K., & Albus, H. (1997). dence in chronobiology: Critical review of a report that res-
Multiunit activity recordings in the suprachiasmatic nuclei: toration of Bmal1 expression in the dorsomedial
In vivo versus in vitro models. Brain Research, 753, 322–327. hypothalamus is sufficient to restore circadian food anticipa-
Menaker, M., & Eskin, A. (1966). Entrainment of circadian tory rhythms in Bmal1/ mice. Journal of Circadian
rhythms by sound in Passer domesticus. Science, 154, Rhythms, 7, 3.
1579–1581. Mistlberger, R. E., & Holmes, M. M. (2000). Behavioral
Merrow, M., Dragovic, Z., Tan, Y., Meyer, G., Sveric, K., feedback regulation of circadian rhythm phase angle in
Mason, M., et al. (2003). Combining theoretical and experi- light-dark entrained mice. American Journal of Physiology.
mental approaches to understand the circadian clock. Chro- Regulatory, Integrative and Comparative Physiology, 279,
nobiology International, 20, 559–575. R813–R821.
Meyer, E. L., Harrington, M. E., & Rahmani, T. (1993). Mistlberger, R. E., Marchant, E. G., & Sinclair, S. V. (1996).
A phase-response curve to the benzodiazepine chlordiaz- Nonphotic phase-shifting and the motivation to run: Cold
epoxide and the effect of geniculo-hypothalamic tract abla- exposure reexamined. Journal of Biological Rhythms, 11,
tion. Physiology & Behavior, 53, 237–243. 208–215.
Meyer-Bernstein, E. L., & Morin, L. P. (1998). Destruction of Mistlberger, R. E., & Skene, D. J. (2005). Nonphotic entrain-
serotonergic neurons in the median raphe nucleus blocks ment in humans? Journal of Biological Rhythms, 20, 339–352.
circadian rhythm phase shifts to triazolam but not to novel Morin, L. P. (2007). SCN organization reconsidered. Journal
wheel access. Journal of Biological Rhythms, 13, 494–505. of Biological Rhythms, 22, 3–13.
Meyer-Bernstein, E. L., & Morin, L. P. (1999). Electrical stim- Morin, L. P., & Allen, C. N. (2006). The circadian visual sys-
ulation of the median or dorsal raphe nuclei reduces light- tem, 2005. Brain Research Reviews, 51, 1–60.
induced FOS protein in the suprachiasmatic nucleus and Mrosovsky, N. (1991). Double-pulse experiments with non-
causes circadian activity rhythm phase shifts. Neuroscience, photic and photic phase-shifting stimuli. Journal of
92, 267–279. Biological Rhythms, 6, 167–179.
333

Mrosovsky, N. (1996). Locomotor activity and non-photic deficient neuropeptide signaling. Journal of Biological
influences on circadian clocks. Biological Reviews of the Rhythms, 25, 235–246.
Cambridge Philosophical Society, 71, 343–372. Pratt, B. L., & Goldman, B. D. (1986). Activity rhythms and
Mrosovsky, N. (1999). Masking: History, definitions, and mea- photoperiodism of Syrian hamsters in a simulated burrow
surement. Chronobiology International, 16, 415–429. system. Physiology & Behavior, 36, 83–89.
Mrosovsky, N., & Biello, S. M. (1994). Nonphotic phase Prosser, R. A., Dean, R. R., Edgar, D. M., Heller, H. C., &
shifting in the old and the cold. Chronobiology International, Miller, J. D. (1993). Serotonin and the mammalian circadian
11, 232–252. system: I. In vitro phase shifts by serotonergic agonists and
Mrosovsky, N., Reebs, S. G., Honrado, G. I., & Salmon, P. A. antagonists. Journal of Biological Rhythms, 8, 1–16.
(1989). Behavioural entrainment of circadian rhythms. Prosser, R. A., Miller, J. D., & Heller, H. C. (1990). A serotonin
Experientia, 45, 696–702. agonist phase-shifts the circadian clock in the suprachiasmatic
Mrosovsky, N., & Salmon, P. A. (1987). A behavioural nuclei in vitro. Brain Research, 534, 336–339.
method for accelerating re-entrainment of rhythms to new Ralph, M. R., & Menaker, M. (1988). A mutation of the circa-
light-dark cycles. Nature, 330, 372–373. dian system in golden hamsters. Science, 241, 1225–1227.
Mrosovsky, N., & Salmon, P. A. (1990). Triazolam and phase- Ralph, M. R., & Mrosovsky, N. (1992). Behavioral inhibition
shifting acceleration re-evaluated. Chronobiology Interna- of circadian responses to light. Journal of Biological
tional, 7, 35–41. Rhythms, 7, 353–359.
Mrosovsky, N., Salmon, P. A., Menaker, M., & Ralph, M. R. Reebs, S. G., Lavery, R. J., & Mrosovsky, N. (1989). Running
(1992). Nonphotic phase shifting in hamster clock mutants. activity mediates the phase-advancing effects of dark pulses
Journal of Biological Rhythms, 7, 41–49. on hamster circadian rhythms. Journal of Comparative Phys-
Nakamura, T. J., Nakamura, W., Yamazaki, S., Kudo, T., iology. A, 165, 811–818.
Cutler, T., Colwell, C. S., et al. (2011). Age-related decline Reebs, S. G., & Mrosovsky, N. (1989a). Effects of induced
in circadian output. The Journal of Neuroscience, 31, wheel running on the circadian activity rhythms of Syrian
10201–10205. hamsters: Entrainment and phase response curve. Journal
Nishino, H., & Koizumi, K. (1977). Responses of neurons in of Biological Rhythms, 4, 39–48.
the suprachiasmatic nuclei of the hypothalamus to putative Reebs, S. G., & Mrosovsky, N. (1989b). Large phase-shifts of
transmitters. Brain Research, 120, 167–172. circadian rhythms caused by induced running in a re-
Panda, S., Provencio, I., Tu, D. C., Pires, S. S., Rollag, M. D., entrainment paradigm: The role of pulse duration and light.
Castrucci, A. M., et al. (2003). Melanopsin is required for Journal of Comparative Physiology. A, 165, 819–825.
non-image-forming photic responses in blind mice. Science, Reebs, S. G., & St-Coeur, J. (1994). Aftereffects of scheduled
301, 525–527. daily exercise on free-running circadian period in Syrian
Pekala, D., Blasiak, T., Raastad, M., & Lewandowski, M. H. hamsters. Physiology & Behavior, 55, 1113–1117.
(2011). The influence of orexins on the firing rate and pat- Roenneberg, T., & Merrow, M. (1999). Circadian systems and
tern of rat intergeniculate leaflet neurons—Electrophysio- metabolism. Journal of Biological Rhythms, 14, 449–459.
logical and immunohistological studies. The European Roenneberg, T., Merrow, M., & Eisensamer, B. (1998). Cellu-
Journal of Neuroscience, 34, 1406–1418. lar mechanisms of circadian systems. Zoology, 100,
Penev, P. D., Turek, F. W., & Zee, P. C. (1995a). A serotonin 273–286.
neurotoxin attenuates the phase-shifting effects of triazolam Rosenwasser, A. M., & Dwyer, S. M. (2001). Circadian phase
on the circadian clock in hamsters. Brain Research, 669, shifting: Relationships between photic and nonphotic
207–216. phase-response curves. Physiology & Behavior, 73, 175–183.
Penev, P. D., Zee, P. C., Wallen, E. P., & Turek, F. W. Rusak, B., Meijer, J. H., & Harrington, M. E. (1989). Hamster
(1995b). Aging alters the phase-resetting properties of a circadian rhythms are phase-shifted by electrical stimulation
serotonin agonist on hamster circadian rhythmicity. The of the geniculo-hypothalamic tract. Brain Research, 493,
American Journal of Physiology, 268, R293–R298. 283–291.
Peyron, C., Tighe, D. K., van den Pol, A. N., de Lecea, L., Sakurai, T. (2007). The neural circuit of orexin (hypocretin):
Heller, H. C., Sutcliffe, J. G., et al. (1998). Neurons Maintaining sleep and wakefulness. Nature Reviews. Neuro-
containing hypocretin (orexin) project to multiple neuronal science, 8, 171–181.
systems. The Journal of Neuroscience, 18, 9996–10015. Salgado-Delgado, R., Angeles-Castellanos, M., Buijs, M. R., &
Piggins, H. D., & Guilding, C. (2011). Chronobiology. In M. Escobar, C. (2008). Internal desynchronization in a model of
Welham (Ed.), Essays in Biochemistry Vol. 49 (p. 156). night-work by forced activity in rats. Neuroscience, 154, 922–931.
London: Portland Press Limited. Salgado-Delgado, R., Nadia, S., Angeles-Castellanos, M.,
Power, A., Hughes, A. T., Samuels, R. E., & Piggins, H. D. Buijs, R. M., & Escobar, C. (2010). In a rat model of night
(2010). Rhythm-promoting actions of exercise in mice with work, activity during the normal resting phase produces
334

desynchrony in the hypothalamus. Journal of Biological Sinclair, S. V., & Mistlberger, R. E. (1997). Scheduled activity
Rhythms, 25, 421–431. reorganizes circadian phase of Syrian hamsters under full
Sanggaard, K. M., Hannibal, J., & Fahrenkrug, J. (2003). Sero- and skeleton photoperiods. Behavioural Brain Research,
tonin inhibits glutamate- but not PACAP-induced per gene 87, 127–137.
expression in the rat suprachiasmatic nucleus at night. The Slotten, H. A., Krekling, S., & Pevet, P. (2005). Photic and
European Journal of Neuroscience, 17, 1245–1252. nonphotic effects on the circadian activity rhythm in the
Sato, T., & Kawamura, H. (1984). Circadian rhythms in multi- diurnal rodent Arvicanthis ansorgei. Behavioural Brain
ple unit activity inside and outside the suprachiasmatic Research, 165, 91–97.
nucleus in the diurnal chipmunk (Eutamias sibiricus). Neu- Smale, L., Lee, T., & Nunez, A. A. (2003). Mammalian
roscience Research, 1, 45–52. diurnality: Some facts and gaps. Journal of Biological
Schaap, J., & Meijer, J. H. (2001). Opposing effects of behav- Rhythms, 18, 356–366.
ioural activity and light on neurons of the suprachiasmatic Smale, L., McElhinny, T., Nixon, J., Gubik, B., & Rose, S.
nucleus. The European Journal of Neuroscience, 13, (2001). Patterns of wheel running are related to Fos expres-
1955–1962. sion in neuropeptide-Y-containing neurons in the inter-
Schwartz, W. J., & Zimmerman, P. (1990). Circadian time- geniculate leaflet of Arvicanthis niloticus. Journal of
keeping in Balb/C and C57bl/6 inbred mouse strains. The Biological Rhythms, 16, 163–172.
Journal of Neuroscience, 10, 3685–3694. Smale, L., Michels, K. M., Moore, R. Y., & Morin, L. P.
Sharma, V. K., & Chandrashekaran, M. K. (1997). Rapid (1990). Destruction of the hamster serotonergic system by
phase resetting of a mammalian circadian rhythm by brief 5,7-DHT: Effects on circadian rhythm phase, entrainment
light pulses. Chronobiology International, 14, 537–548. and response to triazolam. Brain Research, 515, 9–19.
Sheward, W. J., Maywood, E. S., French, K. L., Horn, J. M., Smith, R. D., Inouye, S., & Turek, F. W. (1989). Central
Hastings, M. H., Seckl, J. R., et al. (2007). Entrainment to administration of muscimol phase-shifts the mammalian cir-
feeding but not to light: Circadian phenotype of VPAC2 cadian clock. Journal of Comparative Physiology. A, 164,
receptor-null mice. The Journal of Neuroscience, 27, 805–814.
4351–4358. Smith, V. M., Sterniczuk, R., Phillips, C. I., & Antle, M. C.
Shibata, S., & Moore, R. Y. (1993). Neuropeptide Y and optic (2008). Altered photic and non-photic phase shifts in 5-HT
chiasm stimulation affect suprachiasmatic nucleus circadian (1A) receptor knockout mice. Neuroscience, 157, 513–523.
function in vitro. Brain Research, 615, 95–100. Stephan, F. K., Swann, J. M., & Sisk, C. L. (1979a). Anticipa-
Shibata, S., Tsuneyoshi, A., Hamada, T., Tominaga, K., & tion of 24-hr feeding schedules in rats with lesions of the
Watanabe, S. (1992). Phase-resetting effect of 8-OH-DPAT, suprachiasmatic nucleus. Behavioral and Neural Biology,
a serotonin1A receptor agonist, on the circadian rhythm of 25, 346–363.
firing rate in the rat suprachiasmatic nuclei in vitro. Brain Stephan, F. K., Swann, J. M., & Sisk, C. L. (1979b). Entrain-
Research, 582, 353–356. ment of circadian rhythms by feeding schedules in rats with
Shibata, S., Watanabe, A., Hamada, T., Ono, M., & suprachiasmatic lesions. Behavioral and Neural Biology, 25,
Watanabe, S. (1994). N-methyl-D-aspartate induces phase 545–554.
shifts in circadian rhythm of neuronal activity of rat SCN Stokkan, K. A., Yamazaki, S., Tei, H., Sakaki, Y., &
in vitro. The American Journal of Physiology, 267, Menaker, M. (2001). Entrainment of the circadian clock in
R360–R364. the liver by feeding. Science, 291, 490–493.
Shigeyoshi, Y., Taguchi, K., Yamamoto, S., Takekida, S., Storch, K.-F., & Weitz, C. J. (2012). Daily rhythms of food-
Yan, L., Tei, H., et al. (1997). Light-induced resetting of a anticipatory behavioral activity do not require the known
mammalian circadian clock is associated with rapid induc- circadian clock. Proceedings of the National Academy of
tion of the mPer1 transcript. Cell, 91, 1043–1053. Sciences, 106, 6808–6813.
Shimomura, K., & Menaker, M. (1994). Light-induced phase Stranahan, A. M., Lee, K., & Mattson, M. P. (2008). Central
shifts in tau mutant hamsters. Journal of Biological mechanisms of HPA axis regulation by voluntary exercise.
Rhythms, 9, 97–110. Neuromolecular Medicine, 10, 118–127.
Shioiri, T., Takahashi, K., Yamada, N., & Takahashi, S. (1991). Subbaraj, R., & Chandrashekaran, M. K. (1978). Pulses of
Motor activity correlates negatively with free-running darkness shift the phase of a circadian rhythm in an insectiv-
period, while positively with serotonin contents in SCN in orous bat. Journal of Comparative Physiology. A, 127,
free-running rats. Physiology & Behavior, 49, 779–786. 239–243.
Shiota, M., Sudou, M., & Ohshima, M. (1996). Using out- Sumova, A., Ebling, F. J., Maywood, E. S., Herbert, J., &
door exercise to decrease jet lag in airline crewmembers. Hastings, M. H. (1994). Non-photic circadian entrainment
Aviation, Space, and Environmental Medicine, 67, in the Syrian hamster is not associated with phosphorylation
1155–1160. of the transcriptional regulator CREB within the
335

suprachiasmatic nucleus, but is associated with adrenocorti- a benzodiazepine. The American Journal of Physiology,
cal activation. Neuroendocrinology, 59, 579–589. 253, R204–R207.
Takahashi, J. S., DeCoursey, P. J., Bauman, L., & Van Reeth, O., & Turek, F. W. (1989a). Administering
Menaker, M. (1984). Spectral sensitivity of a novel photore- triazolam on a circadian basis entrains the activity rhythm
ceptive system mediating entrainment of mammalian circa- of hamsters. The American Journal of Physiology, 256,
dian rhythms. Nature, 308, 186–188. R639–R645.
Thines, B., & Harmon, F. G. (2010). Ambient temperature Van Reeth, O., & Turek, F. W. (1989b). Stimulated activity
response establishes ELF3 as a required component of mediates phase shifts in the hamster circadian clock induced
the core Arabidopsis circadian clock. Proceedings of the by dark pulses or benzodiazepines. Nature, 339, 49–51.
National Academy of Sciences of the United States of Van Reeth, O., & Turek, F. W. (1990). Daily injections of
America, 107, 3257–3262. triazolam induce long-term changes in hamster circadian
Tischkau, S. A., Gallman, E. A., Buchanan, G. F., & period. The American Journal of Physiology, 259, R514–R520.
Gillette, M. U. (2000). Differential cAMP gating of glut- Van Reeth, O., Zhang, Y., Reddy, A., Zee, P., & Turek, F. W.
amatergic signaling regulates long-term state changes in (1993). Aging alters the entraining effects of an activity-
the suprachiasmatic circadian clock. The Journal of Neuro- inducing stimulus on the circadian clock. Brain Research,
science, 20, 7830–7837. 607, 286–292.
Tominaga, K., Shibata, S., Ueki, S., & Watanabe, S. (1992). Van Reeth, O., Zhang, Y., Zee, P. C., & Turek, F. W. (1992).
Effects of 5-HT1A receptor agonists on the circadian Aging alters feedback effects of the activity-rest cycle on the
rhythm of wheel-running activity in hamsters. European circadian clock. The American Journal of Physiology, 263,
Journal of Pharmacology, 214, 79–84. R981–R986.
Turek, F. W., & Losee-Olson, S. (1986). A benzodiazepine Verwey, M., & Amir, S. (2009). Food-entrainable circadian
used in the treatment of insomnia phase-shifts the mamma- oscillators in the brain. The European Journal of Neurosci-
lian circadian clock. Nature, 321, 167–168. ence, 30, 1650–1657.
Turek, F. W., & Losee-Olson, S. H. (1987a). Dose response Vindlacheruvu, R. R., Ebling, F. J., Maywood, E. S., &
curve for the phase-shifting effect of triazolam on the mam- Hastings, M. H. (1992). Blockade of glutamatergic neuro-
malian circadian clock. Life Sciences, 40, 1033–1038. transmission in the suprachiasmatic nucleus prevents cellu-
Turek, F. W., & Losee-Olson, S. (1987b). Entrainment of the lar and behavioural responses of the circadian system to
circadian activity rhythm to the light-dark cycle can be light. The European Journal of Neuroscience, 4, 673–679.
altered by a short-acting benzodiazepine, triazolam. Journal Vitaterna, M. H., King, D. P., Chang, A. M.,
of Biological Rhythms, 2, 249–260. Kornhauser, J. M., Lowrey, P. L., McDonald, J. D., et al.
Turek, F. W., & Losee-Olson, S. (1988). The circadian rhythm (1994). Mutagenesis and mapping of a mouse gene, Clock,
of LH release can be shifted by injections of a benzodiaze- essential for circadian behavior. Science, 264, 719–725.
pine in female golden hamsters. Endocrinology, 122, von Gall, C., Duffield, G. E., Hastings, M. H., Kopp, M. D.,
756–758. Dehghani, F., Korf, H. W., et al. (1998). CREB in the
Turek, F. W., Penev, P., Zhang, Y., van Reeth, O., & Zee, P. mouse SCN: A molecular interface coding the phase-
(1995). Effects of age on the circadian system. Neuroscience adjusting stimuli light, glutamate, PACAP, and melatonin
and Biobehavioral Reviews, 19, 53–58. for clockwork access. The Journal of Neuroscience, 18,
van den Pol, A. N., Obrietan, K., Chen, G., & Belousov, A. B. 10389–10397.
(1996). Neuropeptide Y-mediated long-term depression of Watanabe, A., Shibata, S., & Watanabe, S. (1995). Circadian
excitatory activity in suprachiasmatic nucleus neurons. The rhythm of spontaneous neuronal activity in the
Journal of Neuroscience, 16, 5883–5895. suprachiasmatic nucleus of old hamster in vitro. Brain
van Geijlswijk, I. M., Korzilius, H. P., & Smits, M. G. (2010). Research, 695, 237–239.
The use of exogenous melatonin in delayed sleep phase dis- Webb, I. C., Patton, D. F., Hamson, D. K., &
order: A meta-analysis. Sleep, 33, 1605–1614. Mistlberger, R. E. (2008). Neural correlates of arousal-
van Praag, H. (2008). Neurogenesis and exercise: Past and induced circadian clock resetting: Hypocretin/orexin and
future directions. Neuromolecular Medicine, 10, 128–140. the intergeniculate leaflet. The European Journal of Neuro-
Van Reeth, O., Sturis, J., Byrne, M. M., Blackman, J. D., science, 27, 828–835.
L’Hermite-Baleriaux, M., Leproult, R., et al. (1994). Noctur- Webb, I. C., Patton, D. F., Landry, G. J., & Mistlberger, R. E.
nal exercise phase delays circadian rhythms of melatonin (2010). Circadian clock resetting by behavioral arousal:
and thyrotropin secretion in normal men. The American Neural correlates in the midbrain raphe nuclei and locus
Journal of Physiology, 266, E964–E974. coeruleus. Neuroscience, 166, 739–751.
Van Reeth, O., & Turek, F. W. (1987). Adaptation of circa- Weinert, D. (2000). Age-dependent changes of the circadian
dian rhythmicity to shift in light-dark cycle accelerated by system. Chronobiology International, 17, 261–283.
336

Weisgerber, D., Redlin, U., & Mrosovsky, N. (1997). Length- motor activity in blinded rats. Brain Research Bulletin, 25,
ening of circadian period in hamsters by novelty-induced 115–119.
wheel running. Physiology & Behavior, 62, 759–765. Yamanaka, Y., Honma, S., & Honma, K. (2008). Scheduled
Wever, R. A. (1979). The circadian system of man. results of exposures to a novel environment with a running-wheel
experiments under temporal isolation. New York, NY: differentially accelerate re-entrainment of mice peripheral
Springer-Verlag. clocks to new light-dark cycles. Genes to Cells, 13, 497–507.
Wickland, C. R., & Turek, F. W. (1991). Phase-shifting effects Yamazaki, S., Kerbeshian, M. C., Hocker, C. G., Block, G. D.,
of acute increases in activity on circadian locomotor rhythms & Menaker, M. (1998). Rhythmic properties of the hamster
in hamsters. The American Journal of Physiology, 261, suprachiasmatic nucleus in vivo. The Journal of Neurosci-
R1109–R1117. ence, 18, 10709–10723.
Wickland, C., & Turek, F. W. (1994). Lesions of the thalamic Yan, L., & Silver, R. (2002). Differential induction and locali-
intergeniculate leaflet block activity-induced phase shifts in zation of mPer1 and mPer2 during advancing and delaying
the circadian activity rhythm of the golden hamster. Brain phase shifts. The European Journal of Neuroscience, 16,
Research, 660, 293–300. 1531–1540.
Wilsbacher, L. D., Yamazaki, S., Herzog, E. D., Song, E. J., Yannielli, P., & Harrington, M. E. (2004). Let there be “more”
Radcliffe, L. A., Abe, M., et al. (2002). Photic and circadian light: Enhancement of light actions on the circadian system
expression of luciferase in mPeriod1-luc transgenic mice through non-photic pathways. Progress in Neurobiology,
invivo. Proceedings of the National Academy of Sciences of 74, 59–76.
the United States of America, 99, 489–494. Yokota, S. I., Horikawa, K., Akiyama, M., Moriya, T.,
Woelfle, M. A., Ouyang, Y., Phanvijhitsiri, K., & Ebihara, S., Komuro, G., et al. (2000). Inhibitory action of
Johnson, C. H. (2004). The adaptive value of circadian brotizolam on circadian and light-induced per1 and per2
clocks: An experimental assessment in cyanobacteria. Cur- expression in the hamster suprachiasmatic nucleus. British
rent Biology, 14, 1481–1486. Journal of Pharmacology, 131, 1739–1747.
Yamada, N., Shimoda, K., Ohi, K., Takahashi, S., & Yoo, S. H., Yamazaki, S., Lowrey, P. L., Shimomura, K.,
Takahashi, K. (1988). Free-access to a running wheel Ko, C. H., Buhr, E. D., et al. (2004). PERIOD2: LUCIFER-
shortens the period of free-running rhythm in blinded rats. ASE real-time reporting of circadian dynamics reveals per-
Physiology & Behavior, 42, 87–91. sistent circadian oscillations in mouse peripheral tissues.
Yamada, N., Shimoda, K., Takahashi, K., & Takahashi, S. Proceedings of the National Academy of Sciences of the
(1986). Change in period of free-running rhythms deter- United States of America, 101, 5339–5346.
mined by two different tools in blinded rats. Physiology & Zhang, Y., Van Reeth, O., Zee, P. C., Takahashi, J. S., &
Behavior, 36, 357–362. Turek, F. W. (1993). Fos protein expression in the circadian
Yamada, N., Shimoda, K., Takahashi, K., & Takahashi, S. clock is not associated with phase shifts induced by a nonphotic
(1990). Relationship between free-running period and stimulus, triazolam. Neuroscience Letters, 164, 203–208.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 19

The impact of the circadian timing system on


cardiovascular and metabolic function

Christopher J. Morris{,{,*, Jessica N. Yang{ and Frank A. J. L. Scheer{,{

{
Division of Sleep Medicine, Brigham and Women’s Hospital, Boston, MA, USA
{
Division of Sleep Medicine, Harvard Medical School, Boston, MA, USA

Abstract: Epidemiological studies show that adverse cardiovascular events peak in the morning (i.e.,
between 6 AM and noon) and that shift work is associated with cardiovascular disease, obesity, and
diabetes. The endogenous circadian timing system modulates certain cardiovascular risk markers to be
highest (e.g., cortisol, nonlinear dynamic heart rate control, and platelet activation) or to respond most
unfavorably to stressors such as exercise (e.g., epinephrine, norepinephrine, and vagal cardiac modulation)
at an internal body time corresponding to the time of day when adverse cardiovascular events most likely
occur. This indicates that the circadian timing system and its interaction with external cardiovascular
stressors (e.g., physical activity) could contribute to the morning peak in adverse cardiovascular events.
Moreover, circadian misalignment and simulated night work have adverse effects on cardiovascular and
metabolic function. This suggests that misalignment between the behavioral cycle and the circadian timing
system in shift workers contributes to that population’s increased risk for cardiometabolic disease.

Keywords: biological clock; circadian misalignment; glucose metabolism; heart; night work; shift work;
suprachiasmatic nucleus.

Introduction diabetes (Centers for Disease Control and Preven-


tion, 2011), and 80 million are obese (Flegal et al.,
Western societies are rife with cardiovascular dis- 2010). The influence of behaviors (e.g., poor diet
ease, diabetes, and obesity. In the United States, it and physical inactivity) as underlying causes for these
is estimated that 83 million adults have cardiovascu- diseases has been researched for decades. However,
lar disease (Roger et al., 2011), 26 million have more recently, evidence has been accumulating for
a contributing role of the endogenous circadian
timing system and its disruption in cardiovascular
*Corresponding author.
Tel.: þ1-617-525-9086; Fax: þ1-617-732-7337
and metabolic disorders. In this review, we dis-
E-mail: cjmorris@rics.bwh.harvard.edu cuss—with a focus on mammals and particularly

http://dx.doi.org/10.1016/B978-0-444-59427-3.00019-8 337
338

humans—the impact of the circadian timing system, the third ventricle. The SCN is a bilateral struc-
its interaction with behaviors (e.g., exercise), and its ture that contains approximately 50,000 neurons
disturbance on cardiovascular and metabolic in humans and 20,000 neurons in rats (Güldner,
function. 1983; Hofman et al., 1988; Swaab et al., 1985;
van den Pol, 1980). Various levels of evidence
show that the SCN regulates circadian rhythms:
The circadian timing system (1) physiological and behavioral rhythms are
abolished in SCN-lesioned animals (Moore and
Most life on earth—ranging from single cellular Eichler, 1972; Stephan and Zucker, 1972; van den
organisms, plants, flies, rats, to humans—contains Pol and Powley, 1979), (2) SCN-lesioned animals
an endogenous timing system that optimally regain circadian rhythms in locomotor activity fol-
synchronizes physiology and behavior (e.g., rest/ lowing receipt of a donor-SCN (Lehman et al.,
activity and fasting/feedings cycles) with the solar 1987), (3) SCN-lesioned animals that receive an
day. The system is known as the circadian (“circa,” SCN transplantation exhibit the same period
around; “dies,” day) timing system and has two length of the donor animal (Ralph et al., 1990),
core characteristics: (1) endogenous rhythmicity and (4) neural firing rate of the SCN exhibits a cir-
that cycles approximately every 24 h, even in the cadian rhythm in vivo and in vitro (Green and
absence of cyclic changes in external factors such Gillette, 1982; Groos and Hendriks, 1982; Meijer
as light and temperature and (2) the capability to et al., 1997). Circadian oscillators—cells that
adjust its timing in response to external factors generate circadian rhythms autonomously from
such as light and/or food intake. The circadian others—are also located in the periphery (e.g.,
timing system engineers a “biological day” and the heart, liver, and pancreas). In addition to the
“biological night” that transition in a cyclic man- SCN, they express circadian rhythms in gene
ner. Here, we define the biological night as the expression, which ultimately can produce endoge-
endogenous circadian time window corresponding nous cyclic rhythms in biology independent of
to the habitual dark episode, that is, the time nor- input from the central pacemaker (Balsalobre
mally characterized by behavioral inactivity in et al., 1998; Brown et al., 2005; Ko and Takahashi,
diurnal (day-active) species and behavioral activ- 2006; Mühlbauer et al., 2004; Storch et al., 2002;
ity in nocturnal (night-active) species. The oppo- Yamazaki et al., 2000; Yoo et al., 2004). However,
site holds true for the term biological day. In the SCN is considered to be the master pacemaker
mammals, the biological night is also the time because synchrony between peripheral clocks
when melatonin plasma concentrations are high. within organs is typically lost without input from
This is true both for diurnal and nocturnal the SCN (Guo et al., 2006).
mammals. The transition between the biological Circadian rhythms are generated and regulated
day and night is associated with relatively large by the concerted expressions of core clock genes,
changes in many physiological variables, such as which compose the primary mammalian oscillatory
circulating levels of melatonin and cortisol, and mechanism (Ko and Takahashi, 2006; Lowrey and
core body temperature (Dijk et al., 1999; Gooley Takahashi, 2004). A complex molecular network of
et al., 2011; Scheer et al., 2009; Van Cauter positive and negative feedback loops drive circadian
et al., 1994; Wehr et al., 2001). rhythms in core clock genes, such as Clock (circa-
In mammals, the circadian timing system is dian locomoter output cycles kaput), Bmal1 (brain
composed of the suprachiasmatic nucleus (SCN) and muscle arnt-like protein-1), Cryptochrome
and circadian oscillators in most peripheral (Cry1, Cry2), and Period (Per1, Per2, and Per3;
tissues. The SCN is situated in the anterior hypo- Lowrey and Takahashi, 2004; Reppert and Weaver,
thalamus on top of the optic chiasm and next to 2001; Shearman et al., 2000b). Participating in the
339

primary feedback loop are the transcription factors, time of exposure. In both diurnal and nocturnal
Clock and Bmal1, which dimerize and activate the animals, light exposure during the biological
transcription of target genes, such as Pers and Crys evening/early night phase delays circadian time
(Gekakis et al., 1998; King et al., 1997; Reppert relative to clock time; the opposite occurs (i.e., a
and Weaver, 2002). Subsequently, Clock-Bmal1- phase advance) when light exposure takes place
mediated transcription is negatively regulated by during the biological morning (Khalsa et al.,
the Per–Cry complex (Jin et al., 1999; Kume et al., 2003; Rosenberg et al., 1991).
1999; Reppert and Weaver, 2001; Zylka et al., 1998).
The importance of Bmal1 in generating circadian
rhythms is demonstrated by homozygous Bmal1- The internal clock and the timing of adverse
deficient mice (Bmal1/), which cannot entrain cardiovascular events
to light/dark cycles and are arrhythmic (Bunger
et al., 2000). Mice possessing the antimorphic (or Epidemiological data demonstrate a 24-h rhythm
dominant-negative) mutation (ClockD19/D19) dis- in the frequency of adverse cardiovascular events
play a longer period ( 27.3 h) and arrhythmicity such as angina, myocardial infarction, stroke,
in constant darkness (Vitaterna et al., 1994). These arrhythmias, cardiac arrest, and sudden cardiac
studies suggest that the Clock:Bmal1 complex is a death (Fig. 1), with the highest incidence occur-
crucial driving force in circadian clock function. ring between approximately 6 AM and noon
However, this theory is challenged by researchers (Cannon et al., 1997; Cohen et al., 1997; D’Avila
who recently demonstrated that Clock-deficient et al., 1995; Elliott, 1998; Goldberg et al., 1990;
mice (Clock/) maintain robust circadian rhyth- Levine et al., 1992; Marler et al., 1989; Muller
micity in locomotor activity in constant darkness, et al., 1985, 1987; Twidale et al., 1989; Willich
albeit with a shorter period ( 23.2 h; DeBruyne et al., 1987, 1992). The above studies may
et al., 2006). Disruptions of negative feedback under-report the occurrence of adverse cardiovas-
elements (e.g., Per1, Per2, Per3, Cry1, and Cry2) cular events during nighttime due to people
also modify circadian rhythms (Cermakian et al., normally sleeping at this time. However, this pos-
2001; Shearman et al., 2000a; Thresher et al., 1998; sible reporting bias does not apply to the morning
Van Der Horst et al., 1999; Vitaterna et al., 1999; peak in arrhythmias because the precise timing
Zheng et al., 1999, 2001). of the events was recorded by electrocardio-
Because the cycle length of the circadian timing graphic recordings (e.g., see Twidale et al., 1989).
system is not exactly 24 h (Czeisler et al., 1999; Furthermore, data obtained from implanted
Duffy et al., 2011), external photic input (i.e., cardioverter-defibrillators—which also record the
the light/dark cycle) and/or nonphotic input precise timing of the events—show that the defi-
(e.g., food intake) is needed to entrain it brillation threshold (i.e., the amount of energy
with the environment. The strongest Zeitgeber required) needed for successful defibrillation is
(German: “time giver”) to the central pacemaker greatest in the morning (Venditti et al., 1996).
is light. For brevity, we only discuss entrainment The use of plasma creatine phosphokinase MB
by light in this review. Intrinsically photosensitive activity—which is elevated approximately 4 h after
retinal ganglion cells containing the photopigment the onset of a myocardial infarction—has con-
melanopsin, together with rods and cones, ini- firmed that the incidence of myocardial infarction
tially detect light (Gooley et al., 2001; Hattar peaks between 6 AM and noon (Muller et al.,
et al., 2002). This signal is then passed along the 1985; Roberts et al., 1975; Willich et al., 1989).
retinohypothalamic tract to the SCN (Moore In patients with sleep apnea, however, the peak in
et al., 1995). The influence of light on the circa- adverse cardiovascular events occurs at night, pos-
dian timing system is dependent on the circadian sibly due to the acute hemodynamic, autonomic,
340

0.12

0.10

Relative frequency per 1-h interval 0.08

0.06

0.04

0.02

0.00
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23
Clock time (h)

Fig. 1. Relative frequency histogram for the time of onset of sustained ventricular tachycardia across the day. The figure illustrates a
broad peak in the onset of sustained ventricular tachycardia between 8 and 11 AM. Redrawn and reproduced with permission from
Twidale et al. (1989).

and oxidative stress associated with apneas (Gami 2003; Sly et al., 1999). Moreover, the above men-
et al., 2005). This further suggests that, since the tioned organs themselves rhythmically express
peak for adverse cardiovascular events for the clock genes, suggesting that the cardiovascular
population as a whole (including those with sleep system could be influenced by the circadian timing
apnea) occurs in the morning, the morning peak system at the local level in addition to input from
in those people without sleep apnea may be even the central pacemaker (Valenzuela et al., 2008;
more pronounced. Yamamoto et al., 2004). In light of the above
In order to understand the underlying mech- findings and the 24-h rhythm in adverse cardiovas-
anisms for the morning peak in adverse cardio- cular events, researchers have investigated if car-
vascular events, studies have mainly focused on diovascular variables such as heart rate, blood
the impact of behavioral changes typical of the pressure, autonomic nervous system activity, and
morning, such as the change in posture from supine platelet function exhibit circadian rhythmicity.
to upright and the start of behavioral activity at From epidemiological studies, it is not possible to
awakening. However, the circadian timing system deduct the relative role of the circadian timing sys-
may also play a role in the morning peak in adverse tem versus behavioral and environmental effects
cardiovascular events. That the SCN may modulate because influences of behaviors and the internal
cardiovascular functioning is suggested by the pres- circadian timing system occur in synchrony. Thus,
ence of multisynaptic projections from the SCN to this calls for experimental designs that can isolate
the heart and organs that regulate cardiovascular these factors.
function through hormones and blood volume Circadian rhythms in humans can be assessed
regulation (e.g., adrenal cortex, adrenal medulla, using either a constant routine protocol or a
and kidneys; Buijs et al., 1993; Scheer et al., 2001, forced desynchrony protocol. The constant routine
341

protocol, developed by Mills et al. (1978), enables and excitatory influence on resting heart rate
the assessment of circadian rhythms because the during the biological day and night, respectively.
influences of behavioral and environmental factors This is reminiscent of the regulation of melatonin
are minimized and equally distributed across the cir- and corticosteroids for which the SCN uses both
cadian cycle by maintaining constant wakefulness, inhibitory (e.g., GABA) and excitatory neuro-
constant semi-recumbent posture; limiting physical transmitters (e.g., glutamate; Kalsbeek et al.,
activity, consistent dim light conditions; and evenly 1996; Perreau-Lenz et al., 2003).
distributing isocaloric snacks or continuously infus- Despite the fact that blood pressure is often men-
ing nutrients. In the forced desynchrony protocol, tioned as a key example of a physiological variable
developed by Kleitman (1963), subjects are sched- under circadian control, two studies using constant
uled to live on a fixed sleep/wake cycle that is routine protocols found no evidence for an endoge-
adequately different from 24 h and in dim light nous circadian rhythm in blood pressure (Kerkhof
conditions, such that the sleep/wake cycle is outside et al., 1998; Van Dongen et al., 2001). However,
the range of entrainment of the master oscillator, recently, we demonstrated a robust endogenous cir-
causing the internal clock to “free run” or drift cadian rhythm in blood pressure in healthy humans,
according to its own internal period. The forced with very similar timing and amplitude across three
desynchrony protocol allows the separate assess- different protocols (a constant routine and two
ment of circadian and behavioral influence by evenly forced desynchrony protocols) that factor out envi-
spreading behavioral (e.g., sleep and wakefulness) ronmental (e.g., light) and behavioral (e.g., sleep/
factors across the circadian cycle. wake cycle and physical activity) influences
A circadian rhythm in resting heart rate has (Fig. 2; Shea et al., 2011). The circadian peak in
been reported by different research groups in blood pressure occurred during the biological even-
healthy humans, with a broad peak occurring dur- ing, corresponding to approximately 9 PM. The
ing the middle of the biological day and a trough lack of a circadian rhythm in blood pressure in pre-
during the biological night (Figs. 2 and 3; Burgess vious studies was likely related to methodological
et al., 1997; Kräuchi and Wirz-Justice, 1994; Scheer limitations, including lighting control, posture, and
et al., 2010; Shea et al., 2011). On the other hand, circadian phase assessment (see Shea et al., 2011).
the reactivity of heart rate to standardized exercise The reactivity of blood pressure to exercise, but
and postural changes is not influenced by the circa- not postural change, is also controlled by the circa-
dian timing system (Fig. 3; Hu et al., 2011; Scheer dian timing system, again with the greatest response
et al., 2010). Considering that the circadian peak occurring during the biological evening (Fig. 3; Hu
in heart rate does not occur during the biological et al., 2011; Scheer et al., 2010). Considering that
morning and that heart rate reactivity to behav- the circadian peak in blood pressure at rest and in
ioral stressors is not under circadian control, it is response to exercise does not occur during the
unlikely that the circadian timing system’s influ- biological morning, it is unlikely that the circadian
ence on heart rate contributes to the morning peak timing system’s influence on blood pressure con-
in adverse cardiovascular events. In rats, there is tributes to the morning peak in adverse cardiovas-
also a circadian rhythm in resting heart rate under cular events. However, future studies are required
constant dark conditions, thus even independent of to determine whether the timing or amplitude of
the circadian rhythm in behavioral activity (Scheer the circadian blood pressure rhythm is changed in
et al., 2001). After lesioning the SCN, the circadian individuals at risk for or with underlying cardio-
rhythm is abolished and the level of resting vascular disease.
heart rate is intermediate between that during Platelets are a key component in the develop-
the biological day and night in intact animals. ment of thromboses that cause myocardial in-
This suggests that the SCN has an inhibitory farctions and ischemic strokes. Recent work has
342

28-h forced desynchrony protocol 20-h forced desynchrony protocol


Corresponding clock time (h) Corresponding clock time (h)
5 9 13 17 21 1 5 9 13 17 21 1
120
P = 0.015 P < 0.0001
102 114 102

SBP (% mean)
118
SBP (mm Hg)

116 100 112 100

114 98 110 98

112

66 104 104
P = 0.018 64 P = 0.0002

DBP (% mean)
DBP (mm Hg)

102 102
64
100 62 100

98 98
62
60
96 96

74 68
P < 0.0001 104 P < 0.0001 104
72 66

HR (% mean)
HR (bpm)

70 100 100
64
68
96 62 96
66
60
92 92
0 60 120 180 240 300 0 60 120 180 240 300
Circadian phase (°) Circadian phase (°)

Fig. 2. Circadian rhythm in resting systolic blood pressure (SBP), diastolic blood pressure (DBP), and heart rate (HR) observed in
two (28- and 20-h) forced desynchrony protocols. Data are reported as mean  standard error of the mean and are expressed in
absolute values (left axes) and as percentages of individual participant’s averages (right axes). Data are plotted according to
circadian phase, that is, separated into six 60 -bins which all equate to 4 h. Gray bars represent the participant’s average
normal clock time for sleep under ambulatory conditions in the 2 weeks prior to admission to the laboratory. Solid lines
represent the cosinor model fits. Dashed vertical lines indicate the circadian phase at which SBP peaked in both forced
desynchrony protocols. Probability data indicate the likelihood of a circadian rhythm in blood pressure and heart rate.
Reproduced with permission from Shea et al. (2011).

demonstrated in healthy humans that platelet aggregation and adhesion to subendothelial tissue
size, count, aggregability, and platelet surface ex- (Merten et al., 2000; Phillips et al., 1988; Sadler,
pression of activated GPIIb-IIIa, P-selectin, and 1998)—are influenced by the circadian timing sys-
GP1b—factors involved in the pathway of platelet tem (Scheer et al., 2011). For platelet aggregability
343

Corresponding clock time (hh:mm) Corresponding clock time (hh:mm)


4:30 12:30 20:30 4:30 12:30 20:30 4:30 12:30 20:30 4:30 12:30 20:30
300
80 160

Epinephrine (% mean)
140

Epinephrine (pg/mL)

SBP (% mean)

SBP (mmHg)
60
200 140
120
40 Rest: P < 0.0001
Reactivity: P = 0.03
100 120
20
Rest
100
Rest: P < 0.0001
Reactivity: P = 0.04(f2) Exercise
0 0

120
Norepinephrine (% mean)

Norepinephrine (pg/mL)
500 80
200

DBP (% mean)

DBP (mmHg)
400
110 75
300 Rest: P = 0.0005
100 Reactivity: n . s .
200
70

Rest: P = 0.0009 100 100


Reactivity: P = 0.01(f2)
0 0 65

200 140
15
150
Cortisol (% mean)

Cortisol (mg/dL)

120
HR (% mean)

HR (bpm)
100 10
150 100
Rest: P < 0.0001
Reactivity: n.s.
50 5 80
100
Rest: P < 0.0001 60
0 0

150
100
Aggregabilitty (% mean)

Aggregability (W ⫻ min)
1500 120
HF (% mean)

100 90
110
HF (s2)

1000

100
50 80
500
Rest: P = 0.003
Reactivity: P = 0.002 90 Rest: P = 0.04(f2)
Reactivity: n.s.
0 0 70

150
Platelet count ( ⫻ 10 /mL)
Platelet count (% mean)

40 120
pNN50 (% mean)

2.2
pNN50 (%)

100 30 110
2.0
20 100
50 1.8
Rest: P = 0.001 10 90
Reactivity: P = 0.002 Rest: n.s.
Reactivity: n.s. 1.6
0 0 80
0 120 240 0 120 240 0 120 240 0 120 240
Cricadian phase (°) Cricadian phase (°)

Fig. 3. Continued
344

and platelet count, a subset of the data regarding the nervous activity—exhibit endogenous circadian
effect of exercise showed no significant circadian rhythmicity with a broad peak during the middle
rhythms, but a 12-h rhythm in aggregability (see of the biological day (Fig. 3; Hu et al., 2011;
Fig. 3 for platelet aggregability and count data; Scheer et al., 2009, 2010, 2011; Shea et al.,
Scheer et al., 2010). A more comprehensive dataset 2011). Therefore, if the timing would be similar
(more than four times larger) and analysis on the in vulnerable populations, the circadian peak
effect of the circadian timing system, mental stress, of catecholamines unlikely contributes to the
passive head-up tilt, and exercise demonstrated sig- morning peak in adverse cardiovascular events.
nificant circadian rhythms, not only in platelet Whether the doubling in epinephrine con-
aggregability and platelet count, but also in platelet centrations during the biological morning, possibly
size and platelet surface expression of activated coinciding with sensitized adrenergic receptors and
GPIIb-IIIa, P-selectin, and GP1b (Scheer et al., the circadian peak in cortisol, is of relevance to
2011). Platelet count and aggregability peaked at a the morning peak in adverse cardiovascular
circadian phase equivalent to 3–8 PM and thus, the events requires further study (Scheer et al.,
circadian timing system’s influence on platelet count 2010). Another cardiovascular biomarker that is
and aggregability is unlikely to contribute to the under circadian control is a measure of scale-
morning peak in myocardial infarctions and ische- invariant, or fractal, behavior of heart beat
mic strokes, unless the timing of its influence is dis- variability, which shows a peak at a circadian
turbed in people at risk for thrombotic events. On phase equivalent to approximately 9 AM (Hu
the other hand, platelet size and surface expressions et al., 2004). This fractal pattern breaks down
of activated GPIIb-IIIa, P-selectin, and GP1b were with disease and its change is predictive of mor-
all greatest during the biological morning, tality (Hu et al., 2009; Mäkikallio et al., 2004;
corresponding to 6 AM to noon (Scheer et al., Peng et al., 1995). Furthermore, lesioning of the
2011), suggesting that the circadian timing system SCN abolishes such fractal patterns in heart beat
may have a role in the greater incidence of adverse variability at time scales between 4 and 24 h, ren-
cardiovascular events in the morning through these dering the pattern similar to random noise and
thrombotic factors. The reactivity of the above demonstrating the SCN as a critical node in
platelet measures to mental, postural, and physical the neural feedback network underlying scale-
stress was not significantly influenced by the circa- invariant patterns in physiology (Hu et al., 2008).
dian timing system, and thus, the influence of the cir- The reactivity of epinephrine and norepineph-
cadian timing system and that of behavioral rine in healthy humans to exercise is dependent
stressors appear to be additive (Scheer et al., 2011). on circadian phase, with two peaks: one in the
In humans, resting plasma epinephrine and biological morning and the other in the biological
norepinephrine levels—markers of sympathetic evening (Fig. 3; Scheer et al., 2010). These

Fig. 3. Influence of the endogenous circadian timing system on cardiovascular variables at rest and in response to standardized
bicycle exercise. Participants undertook an 11-day (including 12 times 20-h “days”) forced desynchrony protocol in which they
undertook 15 min of exercise (60% of maximum heart rate) at the same time into each wake period. Data are expressed relative to
each participant’s resting value averaged across the whole forced desynchrony protocol (left axes) and in absolute values (right
axes) and are plotted according to circadian phase that is separated into six 60 -bins that each equate to  4 h. Black lines and
closed circles indicate resting values, whereas gray lines with open circles represent data obtained during standardized exercise; error
bars represent the standard error of the mean. Gray bars indicate the group average habitual clock time for sleep in the 2 weeks
prior to admission to the laboratory. The vertical dotted line represents the timing of the group average core body temperature
minimum and curly brackets signify the most vulnerable period for adverse cardiovascular events according to epidemiological
research (6 AM to noon). Probability data were obtained via cosinor analysis. A statistically significant (P < 0.05) second harmonic
of a circadian rhythm is indicated by f2 appearing after the probability value. The arrow in the upper right plot is an example of the
reactivity of systolic blood pressure to exercise at particular circadian phase. Reproduced with permission from Scheer et al. (2010).
345

morning peaks coincide with the circadian peak in gene contributes to the time of day variation in
cortisol (Scheer et al., 2009, 2010; Wehr et al., responses to coronary artery ischemia and
2001), which may further increase cardiac vulner- subsequent reperfusion.
ability since cortisol can potentiate the effects of
the sympathetic nervous system (Davies and
Lefkowitz, 1984). Also, the epinephrine response Circadian disruption and cardiovascular function
to a passive head-up tilt in healthy humans is con-
trolled by the circadian timing system, with a In the previous section, we provided recent
broad peak occurring during biological evening evidence for the impact of the circadian timing
(Hu et al., 2011). Taken all together, the circadian system on cardiovascular risk markers that—if
timing system’s influence on the reactivity of epi- confirmed in vulnerable populations—may con-
nephrine and norepinephrine to exercise, but not tribute to the morning peak in adverse cardiovas-
passive head-up tilt, may contribute to the morn- cular events. However, in healthy individuals,
ing peak in adverse cardiovascular events. the circadian timing system has been proposed
Markers of parasympathetic nervous activity dis- to optimally regulate many physiological pro-
play a circadian rhythm at rest in healthy humans, cesses to prepare for the varying demands across
with peaks in the biological night and morning, the sleep/wake cycle. If this would be the case, it
and thus may be “protective” in the morning, at would be expected that circadian disruption
least in healthy subjects at rest (Fig. 3; Burgess et al., would have adverse health effects. Circadian dis-
1997; Hu et al., 2011; Scheer et al., 2010; Shea et al., ruption can be the result of, for example, circa-
2011). The circadian timing system also influences dian misalignment (in which circadian function is
the reactivity of the sympathetic nervous system to intact, but not properly aligned with either the
behavioral stressors, including exercise and passive external environment or among different compo-
head-up tilt, with the largest reduction in para- nents of the circadian timing system), neuroana-
sympathetic markers occurring in the biological tomical damage or changes to the SCN, and
morning, corresponding to approximately 9 AM genetic mutations and variance (for more details,
(see Fig. 3 for responses to exercise; Hu et al., see Rüger and Scheer, 2009). Circadian misalign-
2011; Scheer et al., 2010). If confirmed in vulnerable ment occurs when the circadian timing system
populations, the impact of the internal circadian as a whole is desynchronized from the behavioral
timing system on reactivity of parasympathetic and/or environmental cycles, and this can be
modulation of the heart in response to behavioral caused by external factors (e.g., rapidly shifting
stressors such as exercise could thus contribute to the light/dark and sleep/wake cycle as seen in shift
the morning peak in adverse cardiovascular events workers and with jet lag) and internal factors
(Vanoli et al., 1991). (e.g., blindness preventing entrainment by light
A recent study in mice demonstrated that of the circadian timing system to the 24-h light/
45 min of experimenter-induced ischemia of the dark cycle). Circadian misalignment can also
coronary artery at the beginning of the biological occur among the different circadian oscillators in
night (start of the active phase, thus equivalent to the body, for example, between the SCN and
the human morning) as compared to the start of peripheral oscillators in the liver, as a result of
the biological day resulted in greater infarct vol- “internal desynchrony.” Circadian disruption with
ume, fibrosis, adverse remodeling, and depression underlying neuroanatomical changes in the SCN
of contractile function after 1 month of reperfu- has been demonstrated in Alzheimer’s disease
sion (Durgan et al., 2010). Moreover, deletion of and depression (Wu et al., 2006; Zhou et al.,
the cardiomyocyte circadian clock gene 2001). Animal experimental SCN lesions and
attenuated the above mentioned outcomes. This human case studies of tumors in proximity to the
indicates that the cardiomyocyte circadian clock SCN have also been shown to lead to circadian
346

disruption (Moore and Eichler, 1972; Schwartz Furthermore, mice with surgically induced heart
et al., 1986). Examples of circadian clock gene disease (cardiac hypertrophy through transverse
mutations which impact circadian function aortic constriction) that live on an abnormal
include familial advanced sleep phase syndrome light/dark (10:10 h) cycle following surgery show
in humans and the many clock gene mutations impaired compensatory cardiac and vascular remo-
in experimental animals (Toh et al., 2001; Turek deling, reduced ventricular contractile strength,
et al., 2005). These categories are not mutually and increased blood pressure as compared to
exclusive, for example, familial advanced sleep animals that live on a normal light/dark (12:12 h)
phase syndrome can lead to circadian misalign- cycle (Martino et al., 2007). Phenotypic rescue,
ment. In this section, we will provide some including reversal/attenuation of abnormal pathol-
examples of the detrimental cardiovascular health ogy and genes, only occurs when the external
consequences of circadian disruption. Adverse rhythm is allowed to correspond with the animals’
metabolic consequences of circadian disruption innate 24-h internal rhythm.
will be discussed in the next section. Further evidence for a critical role of proper
Epidemiological studies indicate that shift work is synchrony between the circadian timing system
associated with cardiovascular disease (Kawachi and the external environment comes from
et al., 1995; Knutsson et al., 1986, 1999; Suwazono follow-up studies in Tau mutant hamsters. Tau
et al., 2008b). Because differences in socioeconomic mutant hamsters have a shorter period ( 22 h)
status and life style cannot fully explain these than their wild-type ( 24 h) counterparts (Ralph
observations, researchers have begun investigating and Menaker, 1988). Tau mutant hamsters have
whether circadian misalignment itself may be one a shorter life span and have cardiomyopathy,
of the underlying mechanisms of the above fibrosis, and impaired heart contractility when
associations. Human studies investigating the effect they live on a 24-h light/dark cycle which is
of circadian misalignment or simulated night work beyond their circadian timing system’s range of
on cardiovascular function in controlled laboratory entrainment and thus causes the animal’s behav-
conditions are scarce. Scheer et al. (2009) demonstra- ioral cycle and biological timing system to be out
ted that circadian misalignment (i.e., desynchrony of phase (i.e., desynchronized; Martino et al.,
between the circadian timing system and behaviors) 2008). Remarkably, when Tau mutant hamsters
increases wake time blood pressure. Recent research are placed in 22-h light/dark cycles, similar to their
has demonstrated that an acute bout of evening exer- endogenous circadian period, these pathological
cise significantly lowers blood pressure during changes are largely reversed. Taken together, the
subsequent simulated night work, suggesting that above mentioned findings suggest that circadian
the adverse effect of circadian misalignment on disruption, such as with shift work, may worsen
blood pressure could be negated by evening exercise cardiovascular function in individuals with
(Fullick et al., 2009). Clearly, more experimental preexisting heart disease.
human research on the effects of circadian disruption It has been hypothesized that dysfunction of
on cardiovascular function is needed. the SCN is associated with hypertension. Levels
There is also evidence of adverse cardiovascular of three key neurotransmitters (vasopressin, vaso-
effects of desynchrony between the circadian active intestinal peptide, and neurotensin) in the
timing system and the imposed light/dark cycle human SCN, required for proper circadian func-
from animal experimental work. Cardiomyopathic tioning, are lower in postmortem material of
hamsters that are subjected to a 12-h shift in the patients with essential hypertension than in that
light/dark cycle every week have a shorter life of control subjects (Goncharuk et al., 2001). At
span than cardiomyopathic hamsters that maintain the same time, more corticotropin-releasing hor-
a fixed 24-h light/dark cycle (Penev et al., 1998). mone (CRH) neurons and greater CRH mRNA
347

are present in the paraventricular nucleus (PVN) Other examples of an intricate link between
of the hypothalamus of people with essential the circadian timing system and cardiovascular
hypertension than controls (Goncharuk et al., control come from recent data on the influence
2002). Because CRH raises blood pressure when of clock genes in hypertension and cardiac func-
administered intraventricularly, it has been pro- tioning. Mice without core clock components
posed that the elevated CRH levels in hyperten- Cry1 and Cry2 develop hypertension when their
sive patients may contribute to their hypertension diet contains salt, whereas wild-type mice do
(Kalin et al., 1983). The vasopressinergic neurons not (Doi et al., 2010). This finding may have been
from the SCN have an inhibitory effect on CRH due to Cry-null mice overexcreting the hormone
neurons in the PVN (Buijs et al., 1993; Hermes aldosterone which increases blood pressure
et al., 1996). Thus, the reduced levels of vasopres- through stimulation of reabsorption of sodium
sin in the SCN could result in disinhibition of ions and water. Moreover, mice lacking the
CRH production by the PVN and, in turn, contrib- Clock gene specifically within cardiomyocytes
ute to the abnormal 24-h variation in blood have brachycardia, greater myocardial oxygen
pressure observed in these patients (Goncharuk consumption, and decreased cardiac efficiency
et al., 2002). Alternatively or additionally, it (Bray et al., 2008).
has been hypothesized that the reverse may be
true and that the observed more intense CRH
projection to the SCN in hypertensive patients Circadian disruption and metabolic function
contributes to the decrease in the number of
vasopressin, vasoactive intestinal peptide, and Human endogenous circadian rhythms have been
neurotensin neurons in the SCN (Goncharuk observed in many factors related to metabolism.
et al., 2007). Based on the information above, the For example, glucose, insulin, cortisol, epineph-
evidence that the SCN is involved in cardiovascular rine, norepinephrine, and leptin display endoge-
control (Scheer et al., 2003) and the observation nous circadian variation (Morgan et al., 1998;
that melatonin administration can amplify or syn- Scheer et al., 2009, 2010; Shea et al., 2005; Van
chronize circadian rhythms (Bothorel et al., 2002; Cauter et al., 1994; Wehr et al., 2001). Recent
Koster-van Hoffen et al., 1993; Zaidan et al., research demonstrates that amino acid plasma
1994), Scheer et al. (2004) investigated if 3 weeks concentrations are under endogenous circadian
of nighttime melatonin administration can improve control, through Krüppel-like factor 15-control
blood pressure regulation in patients with essential of the rhythmic expression of multiple enzymes
hypertension. Repeated nighttime melatonin involved in mammalian nitrogen homeostasis
administration enhanced the day/night rhythm (Jeyaraj et al., 2012). SCN-intact rats, that have
amplitude of blood pressure and reduced blood a fixed 6-meals-per-day feeding schedule (to uni-
pressure during the patient’s scheduled sleep formly distribute food intake across the circadian
period. Of interest, acute nighttime melatonin cycle), also have a circadian rhythm in glucose
intake did not influence blood pressure indicating levels, showing that this is not dependent on the
that the improvement in blood pressure after 3 feeding/fasting cycle, and this rhythm is abolished
weeks of melatonin intake was not due to the in SCN-lesioned rats. Multisynaptic projections
drugs acute vasodilatory effect but possibly a result from the SCN to the liver, pancreas, adrenal
of its influence on the circadian timing system. The cortex, and adipocytes have been discovered,
blood pressure lowering effect of melatonin after and these provide neuroanatomical pathways by
repeated administration in hypertensive patients which the SCN can influence metabolic-related
has been replicated by others (Cagnacci et al., factors (Buijs et al., 1999, 2001; Kreier et al., 2006;
2005; Grossman et al., 2006). la Fleur et al., 2000). The SCN can also influence
348

metabolic-related factors (e.g., cortisol) via hor- all—of the adverse metabolic and hormonal effects
monal pathways (Buijs et al., 1993, 1999) and of night work quickly subside after short-term
peripheral clocks per se can influence metabolism exposure (Lund et al., 2001; Ribeiro et al., 1998).
(Marcheva et al., 2010). Taken together, the circa- How long it takes for adverse metabolic changes
dian timing system is an important entity that con- to normalize after chronic exposure, such as fol-
tributes to the control of many metabolic-related lowing chronic shift work, requires further study.
processes, and thus it is conceivable that circadian The simulated shift work studies discussed
disruption results in impaired metabolic function. above did not replicate the duration of wakeful-
In this section, we will provide some examples ness and behavioral activity before tests meals in
of the adverse metabolic-related consequences of both experimental conditions. Thus, it is unclear
circadian disruption. from those studies what the contribution is of cir-
Numerous epidemiological studies have demon- cadian misalignment per se on metabolic and hor-
strated that shift work is a risk factor for metabolic monal factors. Scheer et al. (2009), utilizing a
disorders such as diabetes and obesity (Morikawa forced desynchrony protocol, demonstrated that
et al., 2005; Niedhammer et al., 1996; Suwazono short-term exposure (after 3 days) to circadian
et al., 2006, 2008a). As a consequence, researchers misalignment per se increases the postprandial
have employed experimental models to investigate responses of glucose and insulin, demonstrating
the possible underlying mechanisms. A 9-h phase reduced glucose tolerance and suggesting reduced
advance of the circadian timing system—a simula- insulin sensitivity under these conditions (Fig. 4).
tion of shift work—increased the glucose and insu- It has also been demonstrated, in rats prone to
lin response to a test meal given at the same clock the development of type 2 diabetes, that phase
time when the test meal was preceded by a high- advancing the light/dark cycle by 6 h every 3 days
fat but not low-fat meal (Hampton et al., 1996; for 10 weeks accelerates the onset of type 2 diabetes
Ribeiro et al., 1998). Moreover, a 9-h phase by causing a hastened loss of beta-cell function and
advance only had a detrimental effect on the aver- mass (Gale et al., 2011). Two studies have assessed
age triglyceride response to a test meal when the the effect of more prolonged circadian misalignment
pretest meal is a low- rather than high-fat meal on human metabolism under controlled laboratory
(Hampton et al., 1996; Ribeiro et al., 1998). The conditions. Morgan et al. (1998) found no significant
influence of macronutrient content of a meal on changes in the concentrations of glucose, insulin, and
responses to subsequent meals during night triglycerides as measured in two constant routine
work requires further research. Al-Naimi et al. protocols with small hourly liquid meals, before
(2004) extended the work of the above researchers and after a 19-day forced desynchrony protocol.
by measuring glucose, insulin, and triglyceride Recently, it was demonstrated in humans that 3
responses to multiple meals across simulated day weeks of sleep restriction (5.6 h of sleep opportunity
and night work shifts. The researchers reported per 24 h) concurrent with circadian disruption
that integrated glucose (statistical trend) and tri- increases postprandial glucose levels, possibly due
glyceride levels were higher across the night shift to a reduction in post-meal insulin concentrations
than day shift. Integrated insulin levels were not (Buxton et al., 2012). These adverse alterations were
different between the night shift and day shifts. abolished following 9 days of recovery sleep and cir-
Furthermore, postprandial glucose, insulin, and tri- cadian re-entrainment to a normal sleep/wake cycle.
glyceride levels are higher during night work than Short-term circadian misalignment while
day work under field conditions (Lund et al., subjects consume isocaloric meals also decreases
2001). The glucose and insulin, but not triglyceride, circulating levels of leptin (Fig. 4; Scheer et al.,
postprandial responses normalize 2 days after 2009). Such an effect, if persisted chronically,
ceasing night work, indicating that some—but not could help explain the increased risk for the
349

Aligned development of obesity through increased food


40 Misaligned intake. In addition, rats subjected to simulated
Leptin (% from mean)

12 night work (forced activity during the normal rest


20 phase) gain significantly more body mass than

Leptin (ng/ml)
10 rats undertaking simulated day work (forced
0 activity during the normal active phase; Salgado-
8 Delgado et al., 2008). Total daily food intake did
-20 P < 0.0001
not differ between the two groups, but the “night
6 workers” were less active and this may explain
-40
B L D S the above finding. Moreover, feeding mice for 6
0 4 8 12 16 20 24 28 weeks only during their biological day (normal rest
phase) rather than biological night (normal active
P < 0.0001
phase) increases their body fat percentage (statisti-
Glucose (% from mean)

60
160 cal trend) and body mass (Arble et al., 2009).

Glucose (mg/ml)
40 These increases may have been due to lower energy
140
expended via physical activity and greater food
20 120
intake in the mice fed during the biological day,
although neither of these observations reached
0 100 statistical significance. Exposing mice to light (150
lux of continues light or 16:8 150/5 lux cycle) during
-20 80 the habitual dark phase significantly increases
B L D S
body and fat mass and impairs glucose tolerance
0 4 8 12 16 20 24 28
(Fonken et al., 2010). These results were not
300 associated with any change in total 24-h food con-
P < 0.006
sumption or locomotor activity, but the mice in the
Insulin (% from mean)

60
200 16:8 light/dim light cycle altered their feeding behav-
Insulin (mIU/ml)

ior, with more food being consumed during the


40
100 habitual light phase.
Weekly 6-h phase advances and delays—
20 experimental models applicable to shift work—
0
significantly hasten mortality in mice (Davidson
et al., 2006). In an attempt to explore the mechanisms
-100 B L D S 0 underlying these findings, Castanon-Cervantes et al.
0 4 8 12 16 20 24 28 (2010) investigated the effect of four consecutive
Time since wake (h) weekly 6-h phase advances of the light/dark schedule
on immune responses in mice. Twenty-four hours
Fig. 4. The response of circulating levels of leptin, glucose,
and insulin to circadian misalignment. Data are reported after a lipopolysaccharide (LPS) challenge, systemic
relative to each participant’s average values during circadian concentrations of the proinflammatory cytokines
alignment on the left axes and in absolute values on the right interleukine-1b (IL-1b), granulocyte macrophage
axes. Error bars represent standard error of the mean. Gray colony-stimulating factor, IL-12, and IL-13 were sig-
area indicates the participant’s sleep opportunity. The white
nificantly higher in the shifted than control mice. The
strips within the scheduled sleep opportunity represent when
participants were briefly awoken to perform pulmonary circadian phase angle, as assessed by body tempera-
function tests. B, breakfast; L, lunch; D, dinner; S, snack. ture; locomotor activity; and Per2 expression in four
Reproduced with permission from Scheer et al. (2009). tissues (SCN, liver, thymus, and spleen), at the time
350

of the LPS challenge was similar between the shifted oscillators in the liver, pancreas, kidney, and heart
and control animals, and therefore the findings were while it has no effect on the phase of the SCN
not due to the stimulus being given at a different cir- (Damiola et al., 2000; Stokkan et al., 2001). Conse-
cadian time between the two groups. The increased quently, such a feeding regime can uncouple
IL-6 release in the shifted mice was also observed peripheral oscillators from the SCN. What the
during in vitro LPS stimulation of macrophages. Fur- physiological effects of internal desynchrony are
thermore, the risk of death following the in vivo LPS on metabolism and whether such internal
challenge was significantly greater in the shifted than desynchrony occurs also in humans in a shift work
control mice. The above findings clearly indicate that setting is currently unknown.
repeatedly shifting a mouse’s light/dark cycle has an In addition to their participation in the generation
adverse effect on their immune response. Whether of circadian rhythms, molecular clock genes have
similar adverse changes in immune function are also also been extensively reported to play a role in met-
observed in humans requires further study. abolic function. Homozygous ClockD19 mutant
It is unclear how simulated night work and circa- (C57BL/6J) mice have a significantly attenuated
dian misalignment cause adverse effects such as diurnal feeding rhythm and are hyperphagic and
impaired glucose metabolism. Both night work obese compared to wild-type mice (Turek et al.,
and circadian misalignment decrease total sleep 2005). In addition, they exhibit adipocyte hypertro-
time which also decreases leptin levels and impairs phy, lipid enlargement of hepatocytes with
glucose tolerance and insulin sensitivity pronounced glycogen build-up, hypercholesterol-
(Åkerstedt, 1998; Buxton et al., 2010; Ohayon emia, hypertriglyceridemia, hypoinsulinemia, and
et al., 2002; Spiegel et al., 1999, 2004). Thus, the hyperglycemia. Marcheva et al. (2010) reported that
question rises to what extent the adverse effects whole-body Clock mutation caused impaired glu-
of night work and circadian misalignment are cose tolerance in mice. The researchers also demon-
mediated via a disturbance of sleep. A covariance strated that glucose-stimulated insulin secretion is
analysis performed by Scheer et al. (2009) suggests reduced in isolated pancreatic islets from Clock
that the effect of circadian misalignment on leptin mutants, which may explain the observed impaired
is likely to be at least in part independent of con- glucose tolerance in these animals. Knockout of
current decreases in sleep efficiency. Moreover, Bmal1 alone in C57BL/6J mice showed that isolated
slow wave sleep suppression per se has adverse mouse embryonic fibroblast (MEF) cells could not
effects on glucose metabolism (Tasali et al., successfully differentiate into adipocytes (Shimba
2008). However, during circadian misalignment, et al., 2005). Further, restoration of Bmal1 function
while sleeping during the biological day, there by adenovirus gene transfer into the initial Bmal1-
was no decrease in the amount of slow wave sleep deficient mice restored adipocyte differentiation
(Morris et al., 2012), such that also changes in slow and lipid accumulation in the MEF cells. The pres-
wave sleep could not explain these results. Thus, ence of high Bmal1 mRNA levels in adipocytes
these data suggest that the adverse effects of night increased lipogenesis activity, further suggesting that
work and circadian misalignment on endocrine Bmal1 is involved in regulating lipid metabolism.
factors are not merely mediated by a reduction in The hypoglycemic response following insulin
sleep quality or quantity. Another possible mecha- administration is exacerbated in Clock mutant and
nism underlying some adverse endocrine effects of Bmal1/ mice as compared in wild-type mice, pos-
night work and circadian misalignment relates to sibly due to an impaired counterregulatory gluco-
internal desynchrony caused by feeding schedules. neogenesis response (Rudic et al., 2004). In parallel
Feeding rodents only during the rest phase—as to the ClockD19 mutant mice (Turek et al., 2005),
similarly occurs during night work when they eat mPer2/ mice also had a defective diurnal feeding
during their night shift—can phase shift peripheral rhythm compared to their wild-type littermates.
351

These mice were also hyperphagic under a high-fat research has been undertaken in this area. Fur-
diet and this may explain why these animals became thermore, it is not known if the above strategy
obese. In mice, genetic elimination of Cry increases would entrain, in addition to the central pace-
the expression of genes responsible for hepatic maker, pertinent peripheral oscillators, for exam-
glucogenesis and also raises fasting blood glucose ple, in liver and pancreas. Recent work has
level (Zhang et al., 2010). The above studies indicate demonstrated that restricting food intake in
that core clock genes participate in the regulation of rodents to the biological night (typical active
metabolic processes and many of the phenotypes period) undertaking simulated night work prevents
induced by dysfunctional/lost clock genes are the gain in body mass and the accumulation
hallmarks of metabolic syndrome. of abdominal fat deposits induced by such a
Experimenters have also investigated the effect of work schedule; perhaps by preventing internal
organ-specific clocks on whole-body metabolism to desynchrony (Salgado-Delgado et al., 2010). This
differentiate the effects from behavioral changes. appears to be a promising countermeasure, and
Some abnormal phenotypes displayed by global investigations are needed to determine if such a
Bmal1/ mice are observed in liver-specific knock- strategy is helpful in humans. Morning exercise
outs. The loss of Bmal1 in the liver of mice leads to can suppress daytime hunger and levels of acylated
hypoglycemia and to a loss of rhythmic expression ghrelin (Broom et al., 2007). Morris et al. (2010)
of hepatic genes that regulate glucose homeostasis investigated if evening exercise suppresses hunger,
(Lamia et al., 2008). In recent work, young pan- acylated ghrelin levels, and increases leptin con-
creas-specific Bmal1 mutant mice showed increased centrations during simulated night work. Evening
ad libitum glucose levels, impaired glucose tolerance, exercise increased levels of both acylated ghrelin
lowered insulin secretion, and decreased insulin and leptin but had no influence on mean hunger
responsiveness to glucose (Marcheva et al., 2010). ratings. Participant’s activity—as measured by
These mice had normal body weight, activity, and a wrist accelerometer—was higher during the
feeding rhythms, suggesting that the above men- simulated night shift that was preceded by exercise
tioned phenotypes are attributable to pancreatic which could have beneficial implications for their
clock disruption per se rather than to secondary energy balance. Clearly, more research is needed
changes in behavior. These findings clearly reveal to determine the optimal countermeasures for the
that the abscission/disruption of the pancreatic clock adverse endocrine effects of night work.
can lead to the onset of detrimental diseases such as
diabetes. However, more evidence is still needed
to determine which organ-specific clocks contribute Summary
to global metabolic functions and what the underly-
ing mechanisms may be. The circadian timing system orchestrates cyclic
Countermeasures that prevent or attenuate the variations in numerous cardiovascular and meta-
adverse endocrine effects associated with night bolic functions independent of external influences
work are needed. Researchers have demonstrated such as darkness/light, sleep/wakefulness, rest/
in humans that appropriately timed light exposure activity, and fasting/eating. At rest, the circadian
and/or melatonin intake can entrain the central timing system causes some factors to peak during
pacemaker to an inverted (relative to the solar the biological morning (e.g., cortisol, platelet
day) sleep/wake cycle (Boivin and James, 2002; expression of activated GPIIb-IIIa, P-selectin,
Crowley et al., 2003; Czeisler et al., 1990; Smith and GP1b), which could potentiate the onset
et al., 2009). One would think that such a strategy of adverse cardiovascular events at that time.
would attenuate the adverse effects of night work This suggests that the morning peak in adverse
on metabolic and hormonal factors; however, little cardiovascular events may not be only due to
352

the acute transition from a resting (sleeping, inac- Balsalobre, A., Damiola, F., & Schibler, U. (1998). A serum
tive, supine, and fasting) to an active (awake, shock induces circadian gene expression in mammalian tis-
sue culture cells. Cell, 93, 929–937.
active, standing, and eating) state. Moreover, the Boivin, D. B., & James, F. O. (2002). Circadian adaptation to
circadian timing system causes the greatest night-shift work by judicious light and darkness exposure.
increase in epinephrine level and the largest Journal of Biological Rhythms, 17, 556–567.
decrease in parasympathetic nervous activity in Bothorel, B., Barassin, S., Saboureau, M., Perreau, S., Vivien-
response to exercise to occur in the biological Roels, B., Malan, A., et al. (2002). In the rat, exogenous
melatonin increases the amplitude of pineal melatonin
morning. This further suggests that the circadian secretion by a direct action on the circadian clock. The
timing system contributes to the morning peak European Journal of Neuroscience, 16, 1090–1098.
in adverse cardiovascular events. Previous work Bray, M. S., Shaw, C. A., Moore, M. W. S., Garcia, R. A. P.,
in the above area has been conducted using Zanquetta, M. M., Durgan, D. J., et al. (2008). Disruption
healthy people; thus, further research needs to of the circadian clock within the cardiomyocyte influences
myocardial contractile function, metabolism, and gene
be undertaken in individuals who are at risk of expression. American Journal of Physiology. Heart and Cir-
having an adverse cardiovascular event to deter- culatory Physiology, 294, H1036–H1047.
mine whether there are changes in timing and Broom, D. R., Stensel, D. J., Bishop, N. C., Burns, S. F., &
amplitude of the endogenous circadian rhythms Miyashita, M. (2007). Exercise-induced suppression of
in cardiovascular risk markers. Research has also acylated ghrelin in humans. Journal of Applied Physiology,
102, 2165–2171.
demonstrated that circadian disruption has a pro- Brown, S. A., Fleury-Olela, F., Nagoshi, E., Hauser, C.,
found effect of cardiovascular and metabolic Juge, C., Meier, C. A., et al. (2005). The period length of
function. More research is needed to fully under- fibroblast circadian gene expression varies widely among
stand the underlying mechanisms. Such research human individuals. PLoS Biology, 3, e338.
may help in the development of novel therapies Buijs, R. M., Chun, S. J., Niijima, A., Romijn, H. J., &
Nagai, K. (2001). Parasympathetic and sympathetic control
in the treatment of circadian related disorders. of the pancreas: A role for the suprachiasmatic nucleus
and other hypothalamic centers that are involved in the reg-
ulation of food intake. The Journal of Comparative Neurol-
Acknowledgments ogy, 431, 405–423.
Buijs, R. M., Markman, M., Nunes-Cardoso, B., Hou, Y.-X., &
Shinn, S. (1993). Projections of the suprachiasmatic nucleus
C. J. M. was supported by the National Space to stress-related areas in the rat hypothalamus: A light and
Biomedical Research Institute through NASA electron microscopic study. The Journal of Comparative
NCC 9-58. F. A. J. L. S. was supported by National Neurology, 335, 42–54.
Institute of Health Grants P30-HL101299 and R01 Buijs, R. M., Wortel, J., Van Heerikhuize, J. J., Feenstra, M. G. P.,
Ter Horst, G. J., Romijn, H. J., et al. (1999). Anatomical and
HL094806. functional demonstration of a multisynaptic suprachiasmatic
nucleus adrenal (cortex) pathway. The European Journal of
Neuroscience, 11, 1535–1544.
Bunger, M. K., Wilsbacher, L. D., Moran, S. M., Clendenin, C.,
References Redcliffe, L. A., Hogenesch, J. B., et al. (2000). Mop3 is an
essential component of the master circadian pacemaker in
Åkerstedt, T. (1998). Shift work and disturbed sleep/wakeful- mammals. Cell, 103, 1009–1017.
ness. Sleep Medicine Reviews, 2, 117–128. Burgess, H. J., Trinder, J., Kim, Y., & Luke, D. (1997). Sleep
Al-Naimi, S., Hampton, S. M., Richard, P., Tzung, C., & and circadian influences on cardiac autonomic nervous
Morgan, L. M. (2004). Postprandial metabolic profiles fol- system activity. American Journal of Physiology. Heart and
lowing meals and snacks eaten during simulated night and Circulatory Physiology, 273, H1761–H1768.
day shift work. Chronobiology International, 21, 937–947. Buxton, O. M., Pavlova, M., Reid, E. W., Wang, W.,
Arble, D. M., Bass, J., Laposky, A. D., Vitaterna, M. H., & Simonson, D. C., & Adler, G. K. (2010). Sleep restriction
Turek, F. W. (2009). Circadian timing of food intake con- for 1 week reduces insulin sensitivity in healthy men. Diabe-
tributes to weight gain. Obesity, 17, 2100–2102. tes, 59, 2126–2133.
353

Buxton, O. M., Cain, S. W., O’Connor, S. P., Porter, J. H., the central pacemaker in the suprachiasmatic nucleus. Genes
Duffy, J. F., Wang, W., Czeisler, C.A., Shea, S. A. (2012). & Development, 14, 2950–2961.
Adverse Metabolic Consequences in Humans of Prolonged Davidson, A. J., Sellix, M. T., Daniel, J., Yamazaki, S.,
Sleep Restriction Combined with Circadian Disruption. Sci- Menaker, M., & Block, G. D. (2006). Chronic jet-lag increases
ence Translational Medicine, 4(129), 129ra143. http://dx.doi. mortality in aged mice. Current Biology, 16, R914–R916.
org/10.1126/scitranslmed.3003200. Davies, A. O., & Lefkowitz, R. J. (1984). Regulation of
Cagnacci, A., Cannoletta, M., Renzi, A., Baldassari, F., b-adrenergic receptors by steroid hormones. Annual Review
Arangino, S., & Volpe, A. (2005). Prolonged melatonin of Physiology, 46, 119–130.
administration decreases nocturnal blood pressure in DeBruyne, J. P., Noton, E., Lambert, C. M., Maywood, E. S.,
women. American Journal of Hypertension, 18, 1614–1618. Weaver, D. R., & Reppert, S. M. (2006). A clock shock:
Cannon, M. D. C. P., McCabe, B. S. C. H., Stone, M. D. P. H., Mouse CLOCK is not required for circadian oscillator func-
Schactman, M. S. M., Thompson, P. B., Theroux, M. D. P., tion. Neuron, 50, 465–477.
et al. (1997). Circadian variation in the onset of unstable Dijk, D. J., Duffy, J. F., Riel, E., Shanahan, T. L., &
angina and non-Q-wave acute myocardial infarction (the Czeisler, C. A. (1999). Ageing and the circadian and homeo-
TIMI III Registry and TIMI IIIB). The American Journal static regulation of human sleep during forced desynchrony
of Cardiology, 79, 253–258. of rest, melatonin and temperature rhythms. Journal of
Castanon-Cervantes, O., Wu, M., Ehlen, J. C., Paul, K., Physiology (London), 516(2), 611–627.
Gamble, K. L., Johnson, R. L., et al. (2010). Dysregulation Doi, M., Takahashi, Y., Komatsu, R., Yamazaki, F.,
of inflammatory responses by chronic circadian disruption. Yamada, H., Haraguchi, S., et al. (2010). Salt-sensitive hyper-
The Journal of Immunology, 185, 5796–5805. tension in circadian clock-deficient Cry-null mice involves
Centers for Disease Control and Prevention, (2011). National dysregulated adrenal Hsd3b6. Nature Medicine, 16, 67–74.
diabetes fact sheet: National estimates and general informa- Duffy, J. F., Cain, S. W., Chang, A.-M., Phillips, A. J. K.,
tion on diabetes and prediabetes in the United States, 2011. Münch, M. Y., Gronfier, C., et al. (2011). Sex difference in
Atlanta, GA: U.S. Department of Health and Human the near-24-hour intrinsic period of the human circadian
Services, Centers for Disease Control and Prevention. timing system. Proceedings of the National Academy of
Cermakian, N., Monaco, L., Pando, M. P., Dierich, A., & Sciences of the United States of America, 108, 15602–15608.
Sassone-Corsi, P. (2001). Altered behavioral rhythms and Durgan, D. J., Pulinilkunnil, T., Villegas-Montoya, C.,
clock gene expression in mice with a targeted mutation in Garvey, M. E., Frangogiannis, N. G., Michael, L. H., et al.
the Period1 gene. The EMBO Journal, 20, 3967–3974. (2010). Short communication: Ischemia/reperfusion toler-
Cohen, M. C., Rohtla, K. M., Lavery, C. E., Muller, J. E., & ance is time-of-day–dependent. Circulation Research, 106,
Mittleman, M. A. (1997). Meta-analysis of the morning 546–550.
excess of acute myocardial infarction and sudden cardiac Elliott, W. J. (1998). Circadian variation in the timing of stroke
death. The American Journal of Cardiology, 79, 1512–1516. onset: A meta-analysis. Stroke, 29, 992–996.
Crowley, S. J., Lee, C., Tseng, C. Y., Fogg, L. F., & Eastman, C. I. Flegal, K. M., Carroll, M. D., Ogden, C. L., & Curtin, L. R.
(2003). Combinations of bright light, scheduled dark, (2010). Prevalence and trends in obesity among US adults,
sunglasses, and melatonin to facilitate circadian entrainment 1999-2008. Journal of the American Medical Association,
to night shift work. Journal of Biological Rhythms, 18, 303, 235–241.
513–523. Fonken, L. K., Workman, J. L., Walton, J. C., Weil, Z. M.,
Czeisler, C. A., Duffy, J. F., Shanahan, T. L., Brown, E. N., Morris, J. S., Haim, A., et al. (2010). Light at night increases
Mitchell, J. F., Rimmer, D. W., et al. (1999). Stability, body mass by shifting the time of food intake. Proceedings
precision, and near-24-hour period of the human circadian of the National Academy of Sciences of the United States of
pacemaker. Science, 284, 2177–2181. America, 107, 18664–18669.
Czeisler, C. A., Johnson, M. P., Duffy, J. F., Brown, E. N., Fullick, S., Morris, C., Jones, H., & Atkinson, G. (2009). Prior
Ronda, J. M., & Kronauer, R. E. (1990). Exposure to bright exercise lowers blood pressure during simulated night-work
light and darkness to treat physiologic maladaptation to with different meal schedules. American Journal of Hyper-
night work. The New England Journal of Medicine, 322, tension, 22, 835–841.
1253–1259. Gale, J. E., Cox, H. I., Qian, J., Block, G. D., Colwell, C. S., &
D’Avila, A., Wellens, F., Andries, E., & Brugada, P. (1995). Matveyenko, A. V. (2011). Disruption of circadian rhythms
At what time are implantable defibrillator shocks delivered? accelerates development of diabetes through pancreatic
European Heart Journal, 16, 1231–1233. beta-cell loss and dysfunction. Journal of Biological
Damiola, F., Le Minh, N., Preitner, N., Kornmann, B., Fleury- Rhythms, 26, 423–433.
Olela, F., & Schibler, U. (2000). Restricted feeding Gami, A. S., Howard, D. E., Olson, E. J., & Somers, V. K.
uncouples circadian oscillators in peripheral tissues from (2005). Day–night pattern of sudden death in obstructive
354

sleep apnea. The New England Journal of Medicine, 352, Hampton, S. M., Morgan, L. M., Lawrence, N.,
1206–1214. Anastasiadou, T., Norris, F., Deacon, S., et al. (1996). Post-
Gekakis, N., Staknis, D., Nguyen, H. B., Davis, F. C., prandial hormone and metabolic responses in simulated
Wilsbacher, L. D., King, D. P., et al. (1998). Role of shift work. The Journal of Endocrinology, 151, 259–267.
CLOCK protein in the mammalian circadian mechanism. Hattar, S., Liao, H. W., Takao, M., Berson, D. M., &
Science, 280, 1564–1569. Yau, K. W. (2002). Melanopsin-containing retinal ganglion
Goldberg, R. J., Brady, P., Muller, J. E., Chen, Z., de cells: Architecture, projections, and intrinsic photosensitiv-
Groot, M., Zonneveld, P., et al. (1990). Time of onset of ity. Science, 295, 1065–1070.
symptoms of acute myocardial infarction. The American Hermes, M. L., Coderre, E. M., Buijs, R. M., & Renaud, L. P.
Journal of Cardiology, 66, 140–144. (1996). GABA and glutamate mediate rapid neurotransmis-
Goncharuk, V. D., Buijs, R. M., & Swaab, D. F. (2007). sion from suprachiasmatic nucleus to hypothalamic para-
Corticotropin-releasing hormone neurons in hypertensive ventricular nucleus in rat. The Journal of Physiology, 496
patients are activated in the hypothalamus but not in the (Pt. 3), 749–757.
brainstem. The Journal of Comparative Neurology, 503, Hofman, M. A., Fliers, E., Goudsmit, E., & Swaab, D. F. (1988).
148–168. Morphometric analysis of the suprachiasmatic and para-
Goncharuk, V. D., van Heerikhuize, J., Dai, J. P., Swaab, D. F., ventricular nuclei in the human brain: Sex differences and
& Buijs, R. M. (2001). Neuropeptide changes in the age-dependent changes. Journal of Anatomy, 160, 127–143.
suprachiasmatic nucleus in primary hypertension indicate Hu, K., Ivanov, P. C., Hilton, M. F., Chen, Z., Ayers, R. T.,
functional impairment of the biological clock. The Journal Stanley, H. E., et al. (2004). Endogenous circadian rhythm
of Comparative Neurology, 431, 320–330. in an index of cardiac vulnerability independent of changes
Goncharuk, V. D., van Heerikhuize, J., Swaab, D. F., & in behavior. Proceedings of the National Academy of
Buijs, R. M. (2002). Paraventricular nucleus of the Sciences of the United States of America, 101, 18223–18227.
human hypothalamus in primary hypertension: Activation Hu, K., Scheer, F. A., Buijs, R. M., & Shea, S. A. (2008). The
of corticotropin-releasing hormone neurons. The Journal of endogenous circadian pacemaker imparts a scale-invariant
Comparative Neurology, 443, 321–331. pattern of heart rate fluctuations across time scales spanning
Gooley, J. J., Chamberlain, K., Smith, K. A., Khalsa, S. B. S., minutes to 24 hours. Journal of Biological Rhythms, 23,
Rajaratnam, S. M. W., Van Reen, E., et al. (2011). 265–273.
Exposure to room light before bedtime suppresses melato- Hu, K., Scheer, F. A., Laker, M., Smales, C., & Shea, S. A.
nin onset and shortens melatonin duration in humans. The (2011). Endogenous circadian rhythm in vasovagal response
Journal of Clinical Endocrinology and Metabolism, 96, to head-up tilt. Circulation, 123, 961–970.
E463–E472. Hu, K., Van Someren, E. J. W., Shea, S. A., & Scheer, F. A.
Gooley, J. J., Lu, J., Chou, T. C., Scammell, T. E., & (2009). Reduction of scale invariance of activity fluctuations
Saper, C. B. (2001). Melanopsin in cells of origin of the ret- with aging and Alzheimer’s disease: Involvement of the cir-
inohypothalamic tract. Nature Neuroscience, 4, 1165. cadian pacemaker. Proceedings of the National Academy of
Green, D. J., & Gillette, R. (1982). Circadian rhythm of firing Sciences of the United States of America, 106, 2490–2494.
rate recorded from single cells in the rat suprachiasmatic Jeyaraj, D., Scheer, F. A., Ripperger, J. A., Haldar, S. M.,
brain slice. Brain Research, 245, 198–200. Lu, Y., Prosdocimo, D. A., et al. (2012). Klf15 orchestrates
Groos, G., & Hendriks, J. (1982). Circadian rhythms in electri- circadian nitrogen homeostasis. Cell Metabolism, 15,
cal discharge of rat suprachiasmatic neurones recorded 311–323.
in vitro. Neuroscience Letters, 34, 283–288. Jin, X., Shearman, L. P., Weaver, D. R., Zylka, M. J., De
Grossman, E., Laudon, M., Yalcin, R., Zengil, H., Peleg, E., Vries, G. J., & Reppert, S. M. (1999). A molecular mecha-
Sharabi, Y., et al. (2006). Melatonin reduces night nism regulating rhythmic output from the suprachiasmatic
blood pressure in patients with nocturnal hypertension. circadian clock. Cell, 96, 57–68.
The American Journal of Medicine, 119, 898–902. Kalin, N. H., Shelton, S. E., Kraemer, G. W., &
Güldner, F. H. (1983). Numbers of neurons and astroglial cells McKinney, W. T. (1983). Corticotropin-releasing factor
in the suprachiasmatic nucleus of male and female rats. administered intraventricularly to rhesus monkeys. Peptides,
Experimental Brain Research, 50, 373–376. 4, 217–220.
Guo, H., Brewer, J. M., Lehman, M. N., & Bittman, E. L. Kalsbeek, A., van Heerikhuize, J. J., Wortel, J., & Buijs, R. M.
(2006). Suprachiasmatic regulation of circadian rhythms of (1996). A diurnal rhythm of stimulatory input to the
gene expression in hamster peripheral organs: Effects of hypothalamo-pituitary-adrenal system as revealed by timed
transplanting the pacemaker. The Journal of Neuroscience, intrahypothalamic administration of the vasopressin V1
26, 6406–6412. antagonist. The Journal of Neuroscience, 16, 5555–5565.
355

Kawachi, I., Colditz, G. A., Stampfer, M. J., Willett, W. C., restored by neural transplant. Immunocytochemical charac-
Manson, J. E., Speizer, F. E., et al. (1995). Prospective study terization of the graft and its integration with the host brain.
of shift work and risk of coronary heart disease in women. The Journal of Neuroscience, 7, 1626–1638.
Circulation, 92, 3178–3182. Levine, R. L., Pepe, P. E., Fromm, R. E., Curka, P. A.,
Kerkhof, G. A., Van Dongen, H. P. A., & Bobbert, A. C. (1998). Clark, P. A., Chernow, B., et al. (1992). Prospective evidence
Absence of endogenous circadian rhythmicity in blood pres- of a circadian rhythm for out-of-hospital cardiac arrests. Jour-
sure? American Journal of Hypertension, 11, 373–377. nal of the American Medical Association, 267, 2935–2937.
Khalsa, S. B. S., Jewett, M. E., Cajochen, C., & Czeisler, C. A. Lowrey, P. L., & Takahashi, J. S. (2004). Mammalian circadian
(2003). A phase response curve to single bright light pulses in biology: Elucidating genome-wide levels of temporal organi-
human subjects. Journal of Physiology (London), 549, 945–952. zation. Annual Review of Genomics and Human Genetics, 5,
King, D. P., Zhao, Y., Sangoram, A. M., Wilsbacher, L. D., 407–441.
Tanaka, M., Antoch, M. P., et al. (1997). Positional cloning Lund, J., Arendt, J., Hampton, S. M., English, J., &
of the mouse circadian clock gene. Cell, 89, 641–653. Morgan, L. M. (2001). Postprandial hormone and metabolic
Kleitman, N. (1963). Sleep and wakefulness. Chicago: Univer- responses amongst shift workers in Antarctica. The Journal
sity of Chicago Press. of Endocrinology, 171, 557–564.
Knutsson, A., Åkerstedt, T., Jonsson, B. G., & Orth- Mäkikallio, A. M., Mäkikallio, T. H., Korpelainen, J. T.,
Gomer, K. (1986). Increased risk of ischaemic heart disease Sotaniemi, K. A., Huikuri, H. V., & Myllylä, V. V. (2004).
in shift workers. The Lancet, 2, 89–92. Heart rate dynamics predict poststroke mortality. Neurol-
Knutsson, A., Hallquist, J., Reuterwall, C., Theorell, T., & ogy, 62, 1822–1826.
Åkerstedt, T. (1999). Shiftwork and myocardial infarction: Marcheva, B., Ramsey, K. M., Buhr, E. D., Kobayashi, Y.,
A case-control study. Occupational and Environmental Su, H., Ko, C. H., et al. (2010). Disruption of the clock com-
Medicine, 56, 46–50. ponents CLOCK and BMAL1 leads to hypoinsulinaemia
Ko, C. H., & Takahashi, J. S. (2006). Molecular components of and diabetes. Nature, 466, 627–631.
the mammalian circadian clock. Human Molecular Genetics, Marler, J. R., Price, T. R., Clark, G. L., Muller, J. E.,
15, R271–R277. Robertson, T., Mohr, J. P., et al. (1989). Morning increase
Koster-van Hoffen, G. C., Mirmiran, M., Bos, N. P., in onset of ischemic stroke. Stroke, 20, 473–476.
Witting, W., Delagrange, P., & Guardiola-Lemaitre, B. Martino, T. A., Oudit, G. Y., Herzenberg, A. M., Tata, N.,
(1993). Effects of a novel melatonin analog on circadian Koletar, M. M., Kabir, G. M., et al. (2008). Circadian
rhythms of body temperature and activity in young, mid- rhythm disorganization produces profound cardiovascular
dle-aged, and old rats. Neurobiology of Aging, 14, 565–569. and renal disease in hamsters. American Journal of Physiol-
Kräuchi, K., & Wirz-Justice, A. (1994). Circadian rhythm of ogy. Regulatory, Integrative and Comparative Physiology,
heat production, heart rate, and skin and core temperature 294, R1675–R1683.
under unmasking conditions in men. The American Journal Martino, T. A., Tata, N., Belsham, D. D., Chalmers, J.,
of Physiology, 267, R819–R829. Straume, M., Lee, P., et al. (2007). Disturbed diurnal rhythm
Kreier, F., Kap, Y. S., Mettenleiter, T. C., van Heijningen, C., alters gene expression and exacerbates cardiovascular disease
van der Vliet, J., Kalsbeek, A., et al. (2006). Tracing from with rescue by resynchronization. Hypertension, 49, 1104–1113.
fat tissue, liver, and pancreas: A neuroanatomical frame- Meijer, J. H., Schaap, J., Watanabe, K., & Albus, H. (1997). Mul-
work for the role of the brain in type 2 diabetes. Endocrinol- tiunit activity recordings in the suprachiasmatic nuclei: In vivo
ogy, 147, 1140–1147. versus in vitro models. Brain Research, 753, 322–327.
Kume, K., Zylka, M. J., Sriram, S., Shearman, L. P., Merten, M., Chow, T., Hellums, J. D., & Thiagarajan, P.
Weaver, D. R., Jin, X., et al. (1999). mCRY1 and mCRY2 (2000). A new role for P-selectin in shear-induced platelet
are essential components of the negative limb of the circa- aggregation. Circulation, 102, 2045–2050.
dian clock feedback loop. Cell, 98, 193–205. Mills, J. N., Minors, D. S., & Waterhouse, J. M. (1978). Adap-
la Fleur, S. E., Kalsbeek, A., Wortel, J., & Buijs, R. M. (2000). tation to abrupt time shifts of the oscillator(s) controlling
Polysynaptic neural pathways between the hypothalamus, human circadian rhythms. The Journal of Physiology, 285,
including the suprachiasmatic nucleus, and the liver. Brain 455–470.
Research, 871, 50–56. Moore, R. Y., & Eichler, V. B. (1972). Loss of a circadian
Lamia, K. A., Storch, K. F., & Weitz, C. J. (2008). Physiologi- adrenal corticosterone rhythm following suprachiasmatic
cal significance of a peripheral tissue circadian clock. Pro- lesions in the rat. Brain Research, 42, 201–206.
ceedings of the National Academy of Sciences of the United Moore, R. Y., Speh, J. C., & Card, J. P. (1995). The ret-
States of America, 105, 15172–15177. inohypothalamic tract originates from a distinct subset of
Lehman, M. N., Silver, R., Gladstone, W. R., Kahn, R. M., retinal ganglion cells. The Journal of Comparative Neurol-
Gibson, M., & Bittman, E. L. (1987). Circadian rhythmicity ogy, 352, 351–366.
356

Morgan, L., Arendt, J., Owens, D., Folkard, S., Hampton, S., Ralph, M. R., & Menaker, M. (1988). A mutation of the circa-
Deacon, S., et al. (1998). Effects of the endogenous clock dian system in golden hamsters. Science, 241, 1225–1227.
and sleep time on melatonin, insulin, glucose and lipid Reppert, S. M., & Weaver, D. R. (2001). Molecular analysis of
metabolism. The Journal of Endocrinology, 157, 443–451. mammalian circadian rhythms. Annual Review of Physiol-
Morikawa, Y., Nakagawa, H., Miura, K., Soyama, Y., ogy, 63, 647–676.
Ishizaki, M., Kido, T., et al. (2005). Shift work and the risk of Reppert, S. M., & Weaver, D. R. (2002). Coordination of cir-
diabetes mellitus among Japanese male factory workers. Scan- cadian timing in mammals. Nature, 418, 935–941.
dinavian Journal of Work, Environment & Health, 31, 179–183. Ribeiro, D., Hampton, S. M., Morgan, L., Deacon, S., &
Morris, C. J., Aeschbach, D., & Scheer, F. A. (2012). Circadian Arendt, J. (1998). Altered postprandial hormone and meta-
system, sleep and endocrinology. Molecular and Cellular bolic responses in a simulated shift work environment. The
Endocrinology, 349, 91–104. Journal of Endocrinology, 158, 305–310.
Morris, C. J., Fullick, S., Gregson, W., Clarke, N., Doran, D., Roberts, R., Gowda, K. S., Ludbrook, P. A., & Sobel, B. E.
MacLaren, D., et al. (2010). Paradoxical post-exercise (1975). Specificity of elevated serum MB creatine phospho-
responses of acylated ghrelin and leptin during a simulated kinase activity in the diagnosis of acute myocardial infarc-
night shift. Chronobiology International, 27, 590–605. tion. The American Journal of Cardiology, 36, 433–437.
Mühlbauer, E., Wolgast, S., Finckh, U., Peschke, D., & Roger, V. L., Go, A. S., Lloyd-Jones, D. M., Adams, R. J.,
Peschke, E. (2004). Indication of circadian oscillations in Berry, J. D., Brown, T. M., et al. (2011). Heart disease and
the rat pancreas. FEBS Letters, 564, 91–96. stroke statistics—2011 update: A report from the american
Muller, J. E., Ludmer, P. L., Willich, S. N., Tofler, G. H., heart association. Circulation, 123, e18–e209.
Aylmer, G., Klangos, I., et al. (1987). Circadian variation in Rosenberg, R. S., Zee, P. C., & Turek, F. W. (1991). Phase
the frequency of sudden cardiac death. Circulation, 75, 131–138. response curves to light in young and old hamsters. Ameri-
Muller, J. E., Stone, P. H., Turi, Z. G., Rutherford, J. D., can Journal of Physiology. Regulatory, Integrative and Com-
Czeisler, C. A., Parker, C., et al. (1985). Circadian variation parative Physiology, 261, R491–R495.
in the frequency of onset of acute myocardial infarction. The Rudic, R. D., McNamara, P., Curtis, A.-M., Boston, R. C.,
New England Journal of Medicine, 313, 1315–1322. Panda, S., Hogenesch, J. B., et al. (2004). BMAL1 and
Niedhammer, I., Lert, F., & Marne, M. J. (1996). Prevalence of CLOCK, two essential components of the circadian clock,
overweight and weight gain in relation to night work in a are involved in glucose homeostasis. PLoS Biology, 2, e377.
nurses’ cohort. International Journal of Obesity and Related Rüger, M., & Scheer, F. A. (2009). Effects of circadian disrup-
Metabolic Disorders, 20, 625–633. tion on the cardiometabolic system. Reviews in Endocrine &
Ohayon, M. M., Lemoine, P., Arnaud-Briant, V., & Metabolic Disorders, 10, 245–260.
Dreyfus, M. (2002). Prevalence and consequences of sleep Sadler, J. E. (1998). Biochemistry and genetics of von Willebrand
disorders in a shift worker population. Journal of Psychoso- factor. Annual Review of Biochemistry, 67, 395–424.
matic Research, 53, 577–583. Salgado-Delgado, R., Ángeles-Castellanos, M., Buijs, M. R., &
Penev, P. D., Kolker, D. E., Zee, P. C., & Turek, F. W. (1998). Escobar, C. (2008). Internal desynchronization in a model of
Chronic circadian desynchronization decreases the survival of night-work by forced activity in rats. Neuroscience, 154,
animals with cardiomyopathic heart disease. The American 922–931.
Journal of Physiology, 275, H2334–H2337. Salgado-Delgado, R., Angeles-Castellanos, M., Saderi, N.,
Peng, C. K., Havlin, S., Hausdorff, J. M., Mietus, J. E., Buijs, R. M., & Escobar, C. (2010). Food intake during the
Stanley, H. E., & Goldberger, A. L. (1995). Fractal normal activity phase prevents obesity and circadian
mechanisms and heart rate dynamics: Long-range cor- desynchrony in a rat model of night work. Endocrinology,
relations and their breakdown with disease. Journal of Elec- 151, 1019–1029.
trocardiology, 28(Suppl. 1), 59–65. Scheer, F. A., Hilton, M. F., Mantzoros, C. S., & Shea, S. A.
Perreau-Lenz, S., Kalsbeek, A., Garidou, M. L., Wortel, J., (2009). Adverse metabolic and cardiovascular consequences
van der Vliet, J., van Heijningen, C., et al. (2003). of circadian misalignment. Proceedings of the National
Suprachiasmatic control of melatonin synthesis in rats: Academy of Sciences of the United States of America, 106,
Inhibitory and stimulatory mechanisms. The European Jour- 4453–4458.
nal of Neuroscience, 17, 221–228. Scheer, F. A., Hu, K., Evoniuk, H., Kelly, E. E., Malhotra, A.,
Phillips, D., Charo, I., Parise, L., & Fitzgerald, L. (1988). The Hilton, M. F., et al. (2010). Impact of the human circadian
platelet membrane glycoprotein IIb-IIIa complex. Blood, system, exercise, and their interaction on cardiovascular
71, 831–843. function. Proceedings of the National Academy of Sciences
Ralph, M. R., Foster, R. G., Davis, F. C., & Menaker, M. of the United States of America, 107, 20541–20546.
(1990). Transplanted suprachiasmatic nucleus determines Scheer, F. A., Kalsbeek, A., & Buijs, R. M. (2003). Cardiovas-
circadian period. Science, 247, 975–978. cular control by the suprachiasmatic nucleus: Neural and
357

neuroendocrine mechanisms in human and rat. Biological leptin levels, elevated ghrelin levels and increased hunger
Chemistry, 384, 697–709. and appetite. Annals of Internal Medicine, 141, 846–850.
Scheer, F. A., Michelson, A. D., Frelinger, A. L., III, Spiegel, K., Leproult, R., & Van Cauter, E. (1999). Impact of
Evoniuk, H., Kelly, E. E., McCarthy, M., et al. (2011). The sleep debt on metabolic and endocrine function. The Lancet,
human endogenous circadian system causes greatest platelet 354, 1435–1439.
activation during the biological morning independent of Stephan, F. K., & Zucker, I. (1972). Circadian rhythms in
behaviors. PLoS One, 6, e24549. drinking behavior and locomotor activity of rats are
Scheer, F. A., Ter Horst, G. J., van der Vliet, J., & Buijs, R. M. eliminated by hypothalamic lesions. Proceedings of the
(2001). Physiological and anatomic evidence for regulation National Academy of Sciences of the United States of
of the heart by suprachiasmatic nucleus in rats. American America, 69, 1583–1586.
Journal of Physiology. Heart and Circulatory Physiology, Stokkan, K. A., Yamazaki, S., Tei, H., Sakaki, Y., &
280, H1391–H1399. Menaker, M. (2001). Entrainment of the circadian clock in
Scheer, F. A., Van Montfrans, G. A., Van Someren, E. J., the liver by feeding. Science, 291, 490–493.
Mairuhu, G., & Buijs, R. M. (2004). Daily nighttime melato- Storch, K. F., Lipan, O., Leykin, I., Viswanathan, N.,
nin reduces blood pressure in male patients with essential Davis, F. C., Wong, W. H., et al. (2002). Extensive and
hypertension. Hypertension, 43, 192–197. divergent circadian gene expression in liver and heart.
Schwartz, W. J., Busis, N. A., & Tessa Hedley-Whyte, E. Nature, 417, 78–83.
(1986). A discrete lesion of ventral hypothalamus and optic Suwazono, Y., Dochi, M., Sakata, K., Okubo, Y., Oishi, M.,
chiasm that disturbed the daily temperature rhythm. Journal Tanaka, K., et al. (2008a). A longitudinal study on the effect
of Neurology, 233, 1–4. of shift work on weight gain in male Japanese workers. Obe-
Shea, S. A., Hilton, M. F., Hu, K., & Scheer, F. A. (2011). sity, 16, 1887–1893.
Existence of an endogenous circadian blood pressure Suwazono, Y., Dochi, M., Sakata, K., Okubo, Y., Oishi, M.,
rhythm in humans that peaks in the evening. Circulation Tanaka, K., et al. (2008b). Shift work is a risk factor for
Research, 108, 980–984. increased blood pressure in Japanese men. Hypertension,
Shea, S. A., Hilton, M. F., Orlova, C., Ayers, R. T., & 52, 581–586.
Mantzoros, C. S. (2005). Independent circadian and sleep/ Suwazono, Y., Sakata, K., Okubo, Y., Harada, H., Oishi, M.,
wake regulation of adipokines and glucose in humans. The Kobayashi, E., et al. (2006). Long-term longitudinal study
Journal of Clinical Endocrinology and Metabolism, 90, on the relationship between alternating shift work and the
2537–2544. onset of diabetes mellitus in male Japanese workers. Journal
Shearman, L. P., Jin, X., Lee, C., Reppert, S. M., & of Occupational and Environmental Medicine, 48, 455–461.
Weaver, D. R. (2000). Targeted disruption of the mPer3 http://dx.doi.org/10.1097/01.jom.0000214355.69182.fa.
gene: Subtle effects on circadian clock function. Molecular Swaab, D. F., Fliers, E., & Partiman, T. S. (1985). The
and Cellular Biology, 20, 6269–6275. suprachiasmatic nucleus of the human brain in relation to
Shearman, L. P., Sriram, S., Weaver, D. R., Maywood, E. S., sex, age and senile dementia. Brain Research, 342, 37–44.
Chaves, I., Zheng, B., et al. (2000). Interacting molecular Tasali, E., Leproult, R., Ehrmann, D. A., & Van Cauter, E.
loops in the mammalian circadian clock. Science, 288, (2008). Slow-wave sleep and the risk of type 2 diabetes in
1013–1019. humans. Proceedings of the National Academy of Sciences
Shimba, S., Ishii, N., Ohta, Y., Ohno, T., Watabe, Y., of the United States of America, 105, 1044–1049.
Hayashi, M., et al. (2005). Brain and muscle Arnt-like Thresher, R. J., Vitaterna, M. H., Miyamoto, Y.,
protein-1 (BMAL1), a component of the molecular clock, Kazantsev, A., Hsu, D. S., Petit, C., et al. (1998). Role of
regulates adipogenesis. Proceedings of the National mouse cryptochrome blue-light photoreceptor in circadian
Academy of Sciences of the United States of America, 102, photoresponses. Science, 282, 1490–1494.
12071–12076. Toh, K. L., Jones, C. R., He, Y., Eide, E. J., Hinz, W. A.,
Sly, D. J., Colvill, L., McKinley, M. J., & Oldfield, B. J. (1999). Virshup, D. M., et al. (2001). An hPer2 phosphorylation site
Identification of neural projections from the forebrain to the mutation in familial advanced sleep phase syndrome. Sci-
kidney, using the virus pseudorabies. Journal of the Auto- ence, 291, 1040–1043.
nomic Nervous System, 77, 73–82. Turek, F. W., Joshu, C., Kohsaka, A., Lin, E., Ivanova, G.,
Smith, M. R., Fogg, L. F., & Eastman, C. I. (2009). Practical McDearmon, E., et al. (2005). Obesity and metabolic syndrome
interventions to promote circadian adaptation to permanent in circadian Clock mutant mice. Science, 308, 1043–1045.
night shift work: Study 4. Journal of Biological Rhythms, 24, Twidale, N., Taylor, S., Heddle, W. F., Ayres, B. F., &
161–172. Tonkin, A. M. (1989). Morning increase in the time of onset
Spiegel, K., Leproult, R., Tasali, E., Penev, P., & Van of sustained ventricular tachycardia. The American Journal of
Cauter, E. (2004). Sleep curtailment results in decreased Cardiology, 64, 1204–1206.
358

Valenzuela, F. J., Torres-Farfan, C., Richter, H. G., the incidence of sudden cardiac death in the framingham
Mendez, N., Campino, C., Torrealba, F., et al. (2008). Clock heart study population. The American Journal of Cardiol-
gene expression in adult primate suprachiasmatic nuclei and ogy, 60, 801–806.
adrenal: Is the adrenal a peripheral clock responsive to mel- Willich, S. N., Linderer, T., Wegscheider, K., Leizorovicz, A.,
atonin? Endocrinology, 149, 1454–1461. Alamercery, I., & Schroder, R. (1989). Increased morning
Van Cauter, E., Sturis, J., Byrne, M. M., Blackman, J. D., incidence of myocardial infarction in the ISAM Study:
Leproult, R., Ofek, G., et al. (1994). Demonstration of rapid Absence with prior beta-adrenergic blockade. ISAM Study
light-induced advances and delays of the human circadian Group. Circulation, 80, 853–858.
clock using hormonal phase markers. American Journal of Wu, Y.-H., Fischer, D. F., Kalsbeek, A., Garidou-Boof, M.-L.,
Physiology. Endocrinology And Metabolism, 266, E953–E963. van der Vliet, J., van Heijningen, C., et al. (2006). Pineal
van den Pol, A. N. (1980). The hypothalamic suprachiasmatic clock gene oscillation is disturbed in Alzheimer’s disease,
nucleus of rat: Intrinsic anatomy. The Journal of Compara- due to functional disconnection from the “master clock”
tive Neurology, 191, 661–702. The FASEB Journal, 20, 1874–1876.
van den Pol, A. N., & Powley, T. (1979). A fine-grained Yamamoto, T., Nakahata, Y., Soma, H., Akashi, M.,
anatomical analysis of the role of the rat suprachiasmatic Mamine, T., & Takumi, T. (2004). Transcriptional oscilla-
nucleus in circadian rhythms of feeding and drinking. Brain tion of canonical clock genes in mouse peripheral tissues.
Research, 160, 307–326. BMC Molecular Biology, 5, 18.
Van Der Horst, G. T. J., Muijtjens, M., Kobayashi, K., Yamazaki, S., Numano, R., Abe, M., Hida, A., Takahashi, R.,
Takano, R., Kanno, S. I., Takao, M., et al. (1999). Mamma- Ueda, M., et al. (2000). Resetting central and peripheral cir-
lian cry1 and cry2 are essential for maintenance of circadian cadian oscillators in transgenic rats. Science, 288, 682685.
rhythms. Nature, 398, 627–630. Yoo, S. H., Yamazaki, S., Lowrey, P. L., Shimomura, K., Ko, C. H.,
Van Dongen, H. P. A., Maislin, G., & Kerkhof, G. A. (2001). Buhr, E. D., et al. (2004). PERIOD2::LUCIFERASE real-time
Repeated assessment of the endogenous 24-hour profile of reporting of circadian dynamics reveals persistent circadian
blood pressure under constant routine. Chronobiology Inter- oscillations in mouse peripheral tissues. Proceedings of the
national, 18, 85–98. National Academy of Sciences of the United States of America,
Vanoli, E., De Ferrari, G., Stramba-Badiale, M., Hull, S., 101, 5339–5346.
Foreman, R., & Schwartz, P. (1991). Vagal stimulation and Zaidan, R., Geoffriau, M., Brun, J., Taillard, J., Bureau, C.,
prevention of sudden death in conscious dogs with a healed Chazot, G., et al. (1994). Melatonin is able to influence its
myocardial infarction. Circulation Research, 68, 1471–1481. secretion in humans: Description of a phase-response curve.
Venditti, F. J., John, R. M., Hull, M., Tofler, G. H., Neuroendocrinology, 60, 105–112.
Shahian, D. M., & Martin, D. T. (1996). Circadian variation Zhang, E. E., Liu, Y., Dentin, R., Pongsawakul, P. Y.,
in defibrillation energy requirements. Circulation, 94, Liu, A. C., Hirota, T., et al. (2010). Cryptochrome mediates
1607–1612. circadian regulation of cAMP signaling and hepatic gluco-
Vitaterna, M. H., King, D. P., Chang, A. M., Kornhauser, J. M., neogenesis. Nature Medicine, 16, 1152–1156.
Lowrey, P. L., McDonald, J. D., et al. (1994). Mutagenesis Zheng, B., Albrecht, U., Kaasik, K., Sage, M., Lu, W.,
and mapping of a mouse gene, Clock, essential for circadian Vaishnav, S., et al. (2001). Nonredundant roles of the mPer1
behavior. Science, 264, 719–725. and mPer2 genes in the mammalian circadian clock. Cell,
Vitaterna, M. H., Selby, C. P., Todo, T., Niwa, H., 105, 683–694.
Thompson, C., Fruechte, E. M., et al. (1999). Differential Zheng, B., Larkin, D. W., Albrecht, U., Sun, Z. S., Sage, M.,
regulation of mammalian period genes and circadian Eichele, G., et al. (1999). The mPer2 gene encodes a func-
rhythmicity by cryptochromes 1 and 2. Proceedings of the tional component of the mammalian circadian clock. Nature,
National Academy of Science of the United States of 400, 169–173.
America, 96, 12114–12119. Zhou, J.-N., Riemersma, R. F., Unmehopa, U. A.,
Wehr, T. A., Aeschbach, D., & Duncan, W. C. J. (2001). Evi- Hoogendijk, W. J. G., van Heerikhuize, J. J., Hofman, M. A.,
dence for a biological dawn and dusk in the human circadian et al. (2001). Alterations in arginine vasopressin neurons in
timing system. The Journal of Physiology, 535, 937–951. the suprachiasmatic nucleus in depression. Archives of General
Willich, S. N., Goldberg, R. J., Maclure, M., Perriello, L., & Psychiatry, 58, 655–662.
Muller, J. E. (1992). Increased onset of sudden cardiac Zylka, M. J., Shearman, L. P., Weaver, D. R., &
death in the first three hours after awakening. The American Reppert, S. M. (1998). Three period homologs in mammals:
Journal of Cardiology, 70, 65–68. Differential light responses in the suprachiasmatic circadian
Willich, S. N., Levy, D., Rocco, M. B., Tofler, G. H., clock and oscillating transcripts outside the brain. Neuron,
Stone, P. H., & Muller, J. E. (1987). Circadian variation in 20, 1103–1110.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 20

Nutrition and the circadian timing system

Dirk Jan Stenvers{,*, Cora F. Jonkers{, Eric Fliers{, Peter H. L. T. Bisschop{ and
Andries Kalsbeek{,}

{
Department of Endocrinology and Metabolism, Academic Medical Center (AMC), University of Amsterdam,
Amsterdam, The Netherlands
{
Department of Nutrition, Academic Medical Center (AMC), University of Amsterdam, Amsterdam, The Netherlands
}
Hypothalamic Integration Mechanisms, Netherlands Institute for Neuroscience (NIN), An Institute of the
Royal Netherlands Academy of Arts and Sciences (KNAW), Amsterdam, The Netherlands

Abstract: Life on earth has evolved under the daily rhythm of light and dark. Consequently, most creatures
experience a daily rhythm in food availability. In this review, we first introduce the mammalian circadian timing
system, consisting of a central clock in the suprachiasmatic nucleus (SCN) and peripheral clocks in various
metabolic tissues including liver, pancreas, and intestine. We describe how peripheral clocks are
synchronized by the SCN and metabolic signals. Second, we review the influence of the circadian timing
system on food intake behavior, activity of the gastrointestinal system, and several aspects of glucose and
lipid metabolism. Third, the circadian control of digestion and metabolism may have important implications
for several aspects of food intake in humans. Therefore, we review the human literature on health aspects of
meal timing, meal frequency, and breakfast consumption, and we describe the potential implications of the
clock system for the timing of enteral tube feeding and parenteral nutrition. Finally, we explore the
connection between type 2 diabetes and the circadian timing system. Although the past decade has provided
exciting knowledge about the reciprocal relation between biological clocks and feeding/energy metabolism,
future research is necessary to further elucidate this fascinating relationship in order to improve human health.

Keywords: circadian clocks; circadian rhythm; biological clocks; suprachiasmatic nucleus; glucose; glucose
tolerance test; lipids; nutritional sciences; eating; appetite regulation; digestive system; gastrointestinal
system; parenteral nutrition; enteral nutrition; diabetes mellitus type 2.

Introduction

Ever since evolution started, life on earth is sub-


*Corresponding author.
Tel.: þ31 20 5663507; Fax: þ31 20 6917682 ject to the daily rhythm of light and dark. As a
E-mail: d.j.stenvers@amc.uva.nl consequence, many organisms experience a daily

http://dx.doi.org/10.1016/B978-0-444-59427-3.00020-4 359
360

rhythm in food availability. Most creatures, rang- Approximately 20 years ago, it was discovered
ing from bacteria to humans, developed a circa- that the endogenous SCN rhythm is based on a
dian timing system to prepare for the alternating molecular clock mechanism. The core clock genes
daily periods of food intake and fasting. Recent (Clock, BMAL1, Period, and Cryptochrome) are
scientific work has generated a lot of knowledge expressed in a transcriptional–translational feed-
about the circadian control of digestion and back loop with a duration of approximately 24 h
metabolism. There are several indications that (reviewed in Bass and Takahashi, 2010). In turn,
disturbances in daily rhythms may lead to obesity the clock genes regulate the expression of a pleth-
and diabetes. Consequently, at present, scientists ora of other genes in SCN neurons and thereby
aim to use the increasing knowledge about the force SCN electrical activity into a 24-h rhythm
interaction between daily food intake rhythms (Herzog et al., 1998; Nakamura et al., 2002).
and the circadian timing system to improve
health.
Peripheral clocks

The circadian timing system Soon after the discovery of the clock genes, it
became clear that virtually every mammalian cell
Central clock contains a molecular clock (Balsalobre et al., 1998;
Zylka et al., 1998). Animal studies have shown
In mammals, including humans, the central bio- peripheral clock gene oscillations in digestive
logical clock resides in the bilateral hypothalamic organs including liver (Balsalobre et al., 2000;
suprachiasmatic nucleus (SCN). The SCN Damiola et al., 2000; Oishi et al., 2000; Sladek
generates an autonomic rhythm of electrical activ- et al., 2007; Stokkan et al., 2001; Zylka et al.,
ity with a period of approximately 24 h. This 1998), pancreas (Damiola et al., 2000; Muhlbauer
rhythm continues to oscillate even when SCN cells et al., 2004; Sadacca et al., 2011), stomach
are removed from a living organism and brought (Bostwick et al., 2010; Hoogerwerf et al., 2007;
into culture (Bos and Mirmiran, 1990; Green and LeSauter et al., 2009), and intestine (Bostwick
Gillette, 1982; Groos and Hendriks, 1982; Welsh et al., 2010; Hoogerwerf et al., 2007; Sladek et al.,
et al., 1995). The SCN is located superior to the 2007). In human adipose tissue, approximately
optic chiasm and receives direct light information 25% of the genes show a circadian rhythm of
from the retina via the retinohypothalamic tract. expression (Loboda et al., 2009).
Through this light input, the endogenous approxi- Elegant experiments targeting the molecular
mate 24-h rhythm of the SCN is synchronized with clock revealed important aspects of the physiologi-
the environmental rhythm with its period duration cal function of peripheral clocks. For instance, the
of exactly 24 h (Dibner et al., 2010; Kohsaka and molecular liver clock regulates the expression of
Bass, 2007). The central clock signals to several important metabolic genes such as phosphoenol-
hypothalamic (e.g., the paraventricular nucleus pyruvate carboxykinase (Kornmann et al., 2007;
(PVN) and the subparaventricular zone (sSPZ) Lamia et al., 2008; Panda et al., 2002; Roesler
and thalamic (e.g., the intergeniculate leaflet and et al., 1992; Zhang et al., 2010) and glucose-6-
paraventricular nucleus of the thalamus) regions. phosphatase (Kornmann et al., 2007; Lamia et al.,
The central clock is of utmost importance for the 2008; Panda et al., 2002; Zhang et al., 2010), two
regulation of daily sleep/wake rhythms, daily key enzymes in gluconeogenesis. Furthermore, the
rhythms in the secretion of hormones, and daily liver clock controls the expression of major
rhythms in feeding behavior (reviewed in Morin enzymes in oxidative phosphorylation and lipid
and Allen, 2006; Saper et al., 2005). metabolism (Kornmann et al., 2007; Panda et al.,
361

2002). On the other hand, in the pancreatic beta peripheral clocks are strongly influenced by
cells, clock genes play an important role in insulin energy metabolism. Animal studies show that clock
secretion (Allaman-Pillet et al., 2004; Marcheva rhythms in liver and pancreas can be inverted by
et al., 2010; Sadacca et al., 2011). To date, the phys- feeding animals when they normally fast (Damiola
iological function of the molecular clocks in stom- et al., 2000; Stokkan et al., 2001). Presumably, this
ach and intestine remains to be determined. The is due to the responsiveness of the molecular clock
function of the adipose tissue clock is reviewed else- mechanism to metabolites such as NADH and
where in this volume (van der Spek et al., 2012). AMP (Lamia et al., 2009; Ramsey et al., 2009;
Peripheral clocks also show an autonomic cycle Rutter et al., 2001) (see Fig. 1). Second, mammalian
of approximately (but not exactly) 24 h (Brown peripheral clocks are synchronized by the daily
et al., 2008; Yamazaki et al., 2000) that needs syn- rhythm of body temperature (Brown et al., 2002;
chronization with the external light–dark cycle. Buhr et al., 2010). Third, peripheral clocks are
Since most peripheral clocks do not receive light also influenced by hormones that are released upon
information, the SCN communicates its rhythm feeding, such as insulin (Kuriyama et al., 2004;
to peripheral clocks. As indicated, the SCN signal Tahara et al., 2011; Zhang et al., 2009).
is passed on via neural projections to several Taken together, accumulating evidence indicates
thalamic and hypothalamic areas. Subsequently, that a major function of peripheral clocks is to pre-
the SCN signal is forwarded to peripheral clocks pare the human body for the alternating periods of
via hormonal signals and the autonomic nervous food intake (day) and fasting (night) (Bass and Taka-
system. For example, the SCN controls the circa- hashi, 2010; Green et al., 2008; Lamia et al., 2008).
dian rhythm in the release of glucocorticoids
by the adrenal gland (Moore and Eichler, 1972;
Reppert et al., 1981) as well as the circadian Circadian aspects of digestion and metabolism
rhythm in the release of melatonin by the pineal
gland (Reppert et al., 1981). Glucocorticoids are Food intake
potent synchronizers of peripheral clock rhythms
(Balsalobre et al., 2000; Le Minh et al., 2001), During the day, most people grow hungry at reg-
and melatonin may have a role in synchronizing ular intervals. At night, however, most people
peripheral clocks as well (Torres-Farfan et al., sleep without the arousing effect of appetite
2006; Valenzuela et al., 2008). In addition to despite the much longer period of fasting. Does
these hormonal pathways, the SCN has neural the SCN modulate appetite and/or the timing of
connections with a variety of tissues including food intake?
liver (Buijs et al., 2003; Kalsbeek et al., 2004; La Since the discovery of leptin, knowledge about
Fleur et al., 2000; Niijima et al., 1992), pancreas the hypothalamic control of food intake has grown
(Buijs et al., 2001; Niijima et al., 1992), and adi- exponentially, and the key role of the hypothalamic
pose tissue (Kreier et al., 2006) via polysynaptic arcuate nucleus has been revealed by numerous ani-
autonomic projections. Evidence from in vivo mal experiments. First, the arcuate nucleus contains
experiments supports an important functional receptors for various homeostatic messages from
role in transmitting the circadian signal to periph- the body and is therefore considered the “window
eral clocks via these autonomic projections as of the brain.” These messages include ghrelin pro-
well (Cailotto et al., 2009; Guo et al., 2005; duction by the empty stomach, pancreatic insulin
Terazono et al., 2003; Vujovic et al., 2008). production upon feeding, and adipose tissue leptin
In addition to the direct hormonal and neuro- production. Second, the arcuate nucleus contains
nal signals described above, peripheral clocks two sets of neurons that have a major role in the
are synchronized by several indirect signals. First, control of food intake. One subset of neurons
362

Fig. 1. The central biological clock resides in the suprachiasmatic nucleus (SCN). The SCN communicates with peripheral clocks in
liver, pancreas, adipose tissue, and intestine via both indirect signals (sleep–wake rhythm, appetite) and direct signals (autonomic
nervous system, hormones). Figure adapted from Dibner et al. (2010) (with permission).

produces Agouti-related peptide (Ollmann et al., of so-called second order neurons in (1) the
1997; Rossi et al., 1998) and Neuropeptide-Y (Clark PVN that releases neuropeptides including cortico-
et al., 1984; Stanley and Leibowitz, 1985) that trophin-releasing hormone (Arase et al., 1988) and
serve as orexigenic signals. Another subset of thyrotropin-releasing hormone (Ahima et al., 1996;
neurons in the arcuate nucleus produces a-melano- Legradi and Lechan, 1999; Vella et al., 2011) which
cyte-stimulating hormone (Fan et al., 1997) that modulate food intake and (2) the lateral hypothala-
serves as an anorexigenic signal. With their output mus that releases orexin (Edwards et al., 1999;
signals, both subsets of neurons regulate the activity Sakurai et al., 1998; Sweet et al., 1999) and thus
363

increases food intake (for an extensive review on the limit ourselves to data concerning the motor
hypothalamic control of food intake, see Woods and activity of the human gastrointestinal system.
D’Alessio, 2008). An extensive review covering animal experiments
The involvement of the SCN in the timing of on circadian gastrointestinal motor activity can be
food intake was demonstrated using a number of found elsewhere (Hoogerwerf, 2010).
approaches. Lesioning the SCN or isolating it Second, the exocrine secretion of several digestive
from its surrounding tissue results in a loss of fluids is influenced by the time of day. For example,
diurnal rhythms in food intake (Nagai et al., the production of saliva increases over the day in
1978; Nishio et al., 1979), and changes of the light- humans (Dawes, 1975). During fasting, human gas-
dark regimen strongly affect the timing of food tric acid secretion shows a daily rhythm with
intake in rodents (Plata-Salaman and Oomura, increased production in the evening. Presumably,
1987; Zucker, 1971). Furthermore, several genetic this rhythm in gastric acid secretion is regulated by
knockout models show that whole-body disruptions the parasympathetic branch of the autonomic ner-
of the molecular clock also abolish diurnal rhythms vous system as vagotomy (performed as a treatment
of food intake (e.g., Turek et al., 2005; Yang et al., for gastric ulceration) abolishes this rhythm (Moore,
2009). However, most knowledge on the hypotha- 1991). Furthermore, animal research shows that the
lamic regulation of food intake is derived from exocrine production of pancreatic juice is increased
rodent research, and food intake in rodents is tightly during the period of habitual food intake (Maouyo
coupled to wakefulness. Thus, the question whether et al., 1993; Thaela et al., 1998).
the SCN directly regulates food intake behavior and Third, diurnal rhythms exist in the activity of
appetite itself (apart from the SCN effects on wake- digestive enzymes in the brush border of the small
fulness and activity) remains to be elucidated. Nev- intestine. Oligosaccharidases including sucrase
ertheless, a direct effect of the SCN on appetite or show a circadian rhythm of increased activity dur-
food intake behavior is suggested by the SCN output ing the period of food intake, and this rhythm
projections to both the CRF-producing neurons in proceeds during several days of fasting (Saito
the PVN (Vrang et al., 1995) and orexin-producing et al., 1976, 1980). Furthermore, transporters that
neurons in the lateral hypothalamus (Abrahamson absorb glucose and protein from the intestinal
et al., 2001). lumen are more active during the period of food
intake in rats fed ad libitum, but this rhythm does
not seem to persist under constant conditions
Gastrointestinal system (Pan et al., 2002, 2004).

Since most people eat during the day, increased


daytime activity of the digestive system would Carbohydrates
be convenient. Indeed, clear daily rhythms exist
in gastrointestinal motility, exocrine secretion, Considering the daily rhythms in many gastrointes-
and the activity of digestive enzymes. tinal processes, it makes sense that metabolic meal
First, human stomach emptying after identical responses show daily rhythms as well. Indeed, in
meals occurs faster in the morning than in the healthy humans, plasma glucose excursions after
evening (Goo et al., 1987). Furthermore, small identical meals are higher at night than in the morn-
bowel propagation velocity of the migrating ing (Biston et al., 1996; Van Cauter et al., 1992).
motor complex (Kumar et al., 1986; Sanders and However, intestinal processes are not the only
Moore, 1992) as well as colonic motor activity determinant of plasma glucose values. Plasma glu-
(Narducci et al., 1987; Rao et al., 2001) is higher cose levels depend on glucose influx from intestinal
during daytime than during nighttime. Here, we absorption and (during fasting) hepatic glucose
364

production, and glucose efflux to muscle, brain, adi- the net result of rhythms in insulin production,
pose tissue, and (after feeding) the liver. insulin sensitivity, and insulin-independent glu-
Daily rhythms in human glucose metabolism cose uptake.
have been clearly demonstrated by studies in In healthy humans, a daily rhythm in pancreatic
which glucose is administered intravenously. insulin production is observed in response to the
A single intravenous glucose bolus results in administration of tolbutamide, with increased
lower plasma glucose levels in the morning than insulin production in the morning compared to
in the evening (Verrillo et al., 1989; Whichelow the evening (Baker and Jarrett, 1972; Carroll
et al., 1974) (for an excellent review on human and Nestel, 1973; Sensi et al., 1973). The same
circadian glucose metabolism, see Van Cauter rhythm has been described both in response to
et al., 1997). This phenomenon is probably caused intravenous glucose (Carroll and Nestel, 1973;
by daily rhythms in the separate glucose fluxes. Lee et al., 1992; Melani et al., 1976) and intrave-
Human fasting studies suggest the presence of a nous glucagon (Melani et al., 1976). Cultured rat
day–night rhythm in hepatic glucose production pancreatic islets show an autonomous circadian
(Clore et al., 1989). Animal studies from our rhythm of insulin production with enhanced
department have shown a circadian rhythm in nighttime insulin production, which persists for 7
hepatic glucose production as well, with increased days in culture (Peschke and Peschke, 1998).
glucose production at the onset of the active In contrast, rats show a tendency toward increased
period, which is the night in nocturnal animals. daytime insulin production after intravenous glu-
The SCN determines this rhythm in hepatic glu- cose administration (La Fleur et al., 2001). This
cose production (La Fleur et al., 2001) via a diur- paradox may be explained by the observation
nal rhythm of GABAergic inhibitory output to that in vivo insulin-independent mechanisms
preautonomic neurons in the PVN (Kalsbeek also determine the circadian rhythm in glucose
et al., 2004, 2008). From the PVN, the SCN signal tolerance (see above and below), and insulin
is transferred to the liver via a balance between production is in itself dependent on plasma glucose
sympathetic and parasympathetic activity (Cailotto levels.
et al., 2008; Kalsbeek et al., 2004). Interestingly, Furthermore, mammals show a daily rhythm in
one of our studies not only shows that hepatic insulin sensitivity. After an intravenous insulin
clock gene expression is independent from these bolus, blood glucose levels show a more rapid
autonomic signals to the liver, but moreover that decline at the onset of the activity period (day-
rhythmic hepatic clock gene expression is insuffi- time in humans, nighttime in rats), both in
cient to maintain the circadian rhythm in hepatic humans (Gibson and Jarrett, 1972; Gibson et al.,
glucose production after complete hepatic dener- 1975) and in rats (Kalsbeek et al., 2003; La Fleur
vation (Cailotto et al., 2005). This observation is et al., 2001). Finally, some studies indicate a cir-
in contrast with the proposed role of the liver clock cadian rhythm in insulin-independent glucose
in the regulation of gluconeogenesis (Kornmann uptake: After an intravenous glucose tolerance
et al., 2007; Lamia et al., 2008; Panda et al., 2002; test, the calculated evening insulin-independent
Roesler et al., 1992; Zhang et al., 2010), and most glucose uptake is lower than the calculated morn-
likely for an optimal glucose homeostasis, both ing insulin-independent uptake (Lee et al., 1992),
components are necessary. and cultured adipocytes show a rhythm in glucose
In addition to glucose production, glucose uptake in the absence of insulin (Feneberg and
uptake by various tissues shows daily rhythmicity Lemmer, 2004).
as well. At rest, human skeletal muscle consumes Taken together, it is obvious that the circadian
more glucose in the morning than in the evening timing system strongly influences virtually all aspects
(Verrillo et al., 1989). This rhythm is probably of glucose metabolism. A likely physiological
365

explanation for the daily rhythms in glucose metab- Furthermore, hepatic cholesterol production
olism is that an organism needs to be prepared for shows a clear day–night rhythm. In humans,
the active period, that is, the period of food intake cholesterol synthesis shows a strong increase
and increased metabolic needs (La Fleur, 2003; during the night (Cella et al., 1995). Interestingly,
Van Cauter et al., 1997). in rats, hepatic cholesterol synthesis also shows
an increase during the night, due to increased
nighttime synthesis of the rate-limiting enzyme
Lipids 3-hydroxy-3-methylglutaryl-coenzyme A reduc-
tase (Edwards et al., 1972). Thus, unlike many
In addition to carbohydrate metabolism, lipid other physiological processes, the day–night pat-
metabolism seems adjusted to the daily rhythm tern of cholesterol synthesis is not inverted in
of fasting and feeding as well. In humans who nocturnal animals.
eat during the day, plasma free fatty acids (FFAs) The circadian rhythm in hepatic cholesterol
show a strong increase at night due to increased synthesis has clinical implications. Several clinical
lipolysis during the nighttime fast (Schlierf and trials show that the cholesterol-lowering effect of
Dorow, 1973). Evidence for an underlying circa- statins is increased when patients take the statin
dian regulation of lipid metabolism is provided in the evening compared to the morning (Lund
by studies that compare morning and evening et al., 2002; Saito et al., 1991; Wallace et al.,
responses to identical stimuli. Thus, in the 2003). Therefore, it is now common practice to
morning, fasting FFA levels are lower than in instruct patients to take the statin in the evening.
the evening (Carroll and Nestel, 1973; Gibson However, it may be interesting to note that these
et al., 1975; Zimmet et al., 1974). Moreover, trials all show an effect on morning (fasting) cho-
plasma FFAs show a smaller decrease in the lesterol levels, which does not necessarily impli-
morning than in the evening, not only after an cate an effect on average (24 h) cholesterol levels.
oral glucose tolerance test (Zimmet et al., 1974)
but also after identical meals (Pontiroli
et al., 1992; Romon et al., 1997) and after an Meal patterns and human health
intravenous insulin bolus (Carroll and Nestel,
1973; Gibson et al., 1975; Morgan et al., 1999). The circadian control of digestion and metabo-
Finally, plasma levels of triglycerides and FFAs lism may have important implications for the con-
show a slight increase during the night in humans cept of healthy food intake in humans. Next, we
under continuous conditions of 3-h sucrose inges- review the literature on the effects of meal timing,
tion (Barter et al., 1971). The nighttime increase meal frequency, and breakfast consumption on
of lipolysis is possibly regulated by the nightly human health. Furthermore, we discuss potential
surge of growth hormone (GH) secretion, since implications of the circadian timing system for
patients with GH deficiency do not show the the timing of enteral tube feeding and parenteral
nightly increase of FFA levels, and GH replace- nutrition.
ment restores this normal nocturnal increase of
lipolysis, albeit in a supraphysiological dosage
(Boyle et al., 1992). However, the autonomic ner- Day and night
vous system may have an additional role in
regulating the circadian rhythm of lipid metabo- The biological clock has evolved under a pro-
lism, considering the strong effects of the auto- nounced daily alteration of light and dark, cou-
nomic nervous system on the metabolic activity pled to clear alterations of activity and sleep,
of adipose tissue (Fliers et al., 2003). and feeding and fasting. In humans, activity and
366

feeding generally occur during the light period ghrelin levels and reduced leptin levels (Spiegel
(Roenneberg et al., 2012). However, in the pres- et al., 2004).
ent 24-h society, the synchrony between these Third, people with night eating syndrome (NES)
rhythms is easily disturbed. Due to artificial light, often show obesity and an inability to lose weight.
television and internet, average sleep duration NES is characterized by (1) excessive food intake
has decreased over the years (Spiegel et al., late at night, (2) difficulties falling asleep, and (3)
2009). Furthermore, many people become less frequent nightly awakenings. NES occurs in
exposed to bright sunlight since they are living 1–5% of the general population and in 6–14%
and working inside. Finally, people perform less of patients with obesity (Sim et al., 2010). Fourth,
physical exercise during the daytime (Brownson in an interesting experiment, humans showed
et al., 2005). Several lines of evidence suggest that decreased glucose tolerance when they ingested
mismatches between the daily rhythms of activity food at times that conflicted with their internal
and food intake (day), and rest and fasting (night) clock rhythm (Scheer et al., 2009).
lead to health complaints and disease in humans Thus, mismatches between day–night rhythms
(see Fig. 2). have severe pathophysiological implications. Only
First, shift workers are more prone to gastroin- recently, clinical trials investigating the modula-
testinal problems such as constipation, diarrhea, tion of day–night rhythms to prevent or treat dis-
or gastric ulcers. In addition, they are at increased eases have started. For example, one clinical trial
risk of obesity, diabetes (Pan et al., 2011), cardio- is investigating the treatment of obesity by
vascular disease, and cancer (reviewed in Wang increasing sleep duration with 1.5 h per night
et al., 2011). (Cizza et al., 2010).
Second, not only shift workers but also short
sleepers are at increased risk of obesity and dia-
betes (reviewed in Spiegel et al., 2009). One sin- Meal frequency
gle night of sleep disruption already strongly
reduces glucose tolerance (Donga et al., 2010) It is generally believed that a healthy meal pat-
while two nights of short sleep (4 h) causes a tern consists of three meals per day, with a limited
24% increase of appetite and an increased intake number of snacks. The Netherlands Nutrition
of high calorie food in association with increased Centre Foundation also advises to follow this

Light Dark

Wake Sleep

Food intake Fasting

Energy storage Energy use

Fig. 2. In the physiological situation, wakefulness, food intake, and anabolic metabolic processes coincide with the light period.
Sleep, fasting, and catabolic metabolic processes are coupled to the dark period. Disturbance of this circadian synchrony may
lead to obesity and diabetes.
367

meal pattern (http://www.voedingscentrum.nl/nl/ irregular meal pattern (Farshchi et al., 2005a).


schijf-van-vijf/eet-gevarieerd/regelmatig-eten. Thus, although the advice to eat regular meals
aspx, 2012). However, evidence supporting this makes sense from a conceptual point of view,
advice is scarce. sound scientific support still needs to be gathered.
A comprehensive review on the relation
between meal frequency and energy balance
shows obvious discrepancies between a number Breakfast
of epidemiological studies investigating the corre-
lation between meal frequency and obesity Much literature exists on the pros and cons of a
(Bellisle et al., 1997). Furthermore, these epide- daily breakfast, mostly based on epidemiological
miological studies are difficult to interpret, since studies. People who report to consume a regular
obese people tend to record food intake breakfast also report a higher intake of car-
unreliably (Bellisle et al., 1997; Lichtman et al., bohydrates, fat, and protein than people who
1992; Salle et al., 2006). report to omit breakfast. In addition, people who
Thus far, intervention studies have not been report a regular breakfast are more prone to ingest
able to detect significant effects of meal frequency sufficient micronutrients. Many studies indicate
on body weight, body composition, or 24-h energy that children who consume a regular breakfast
expenditure in people on a weight loss diet have a lower body mass index (BMI), but there
(Bellisle et al., 1997; Palmer et al., 2009). Several are also studies showing no relation between
reviews into the effects of meal frequency inter- breakfast consumption and BMI (for systematic
ventions on cardiovascular risk markers found a reviews, see Rampersaud et al., 2005; Ruxton and
trend toward decreasing plasma total cholesterol Kirk, 1997; Szajewska and Ruszczynski, 2010).
and low-density lipoprotein (LDL) levels with Only few intervention studies into the effects
higher meal frequencies. However, plasma high- of breakfast consumption have been performed.
density lipoprotein levels may also decrease with One study suggested a beneficial effect of break-
increasing meal frequencies, which make the fast consumption on insulin sensitivity and
effect on the overall cardiovascular risk profile plasma levels of total cholesterol and LDL
uncertain (Bhutani and Varady, 2009; Mann, (Farshchi et al., 2005b). Furthermore, in devel-
1997; Palmer et al., 2009). The most firmly oping countries, school performance improves
established causal relation is that a high meal fre- in children who consume breakfast regularly
quency leads to an increase of dental caries (Jacoby et al., 1998; Powell et al., 1998). Thus,
(Kandelman, 1997). In summary, a meal fre- several clues indicate positive effects of break-
quency of three meals per day mainly protects fast consumption, but there is a need for more
our teeth, but the evidence for other beneficial controlled intervention trials.
health effects is very limited.
Another frequent advice is to eat regularly.
Considering the strong integration between the Clinical nutrition
circadian timing system and metabolism, this
advice seems to make sense. However, surpris- In the preceding paragraphs, the influence of the
ingly, little research has been performed into the clock system on physiological food intake was
effects of regular food intake on health. One discussed. However, many patients with gastroin-
interesting paper shows that in obese women, a testinal disease need nutritional supplements via
2-week regular meal pattern leads to reduced the enteral or parenteral route of administration.
energy intake, lower fasting cholesterol levels, Presently, the time schedule of administration
and increased insulin sensitivity, compared to an is mainly determined by practical arguments.
368

At home, both parenteral and enteral tube feed- et al., 2006; Metheny, 2006; Metheny et al.,
ing are usually administered during the night. In 2006). Since most people outside the intensive
contrast, patients who are submitted to the hospi- care maintain an upright position during the day,
tal usually receive enteral nutrition during the day the time schedule of enteral tube feeding may
and parenteral nutrition continuously over 24 h. influence risk of aspiration outside the intensive
Below, studies investigating different time care, but this question remains to be assessed in
schedules for (par)enteral nutrition will be clinical trials.
discussed. Only few studies have addressed the Thus, several clues indicate metabolic effects of
question whether patients receiving (par)enteral the time schedule of enteral tube feeding on
nutrition therapy benefit from a schedule that res- nitrogen balance, body temperature, and auto-
embles the physiological situation of feeding dur- nomic activity. Theoretically, the time schedule
ing the day and fasting during the night. Most may affect aspiration risk or general well-being,
studies investigated the differences between con- but evidence is too limited to make an evidence-
tinuous and cyclic administration. Importantly, based choice for a certain schedule. For the time
long-term intervention trials investigating differ- being, the practical advantage of being free of
ent time schedules are still lacking. enteral tube feeding during the day will probably
guide the choice.

Enteral tube feeding


Parenteral nutrition
Two studies showed no difference in appetite sup-
pression between daytime and continuous admin- In a recent systematic review, the metabolic
istration of enteral nutrition (reviewed in Stratton effects of cyclic parenteral nutrition were com-
and Elia, 1999). Healthy volunteers provided with pared to those of continuous parenteral nutrition
enteral tube feedings for 3 days (in addition to (Stout and Cober, 2010). No differences were
regular food intake) showed no difference in detected in nitrogen balance, total energy expen-
appetite, satiation, or resting energy expenditure diture, or the rates of carbohydrate and lipid oxi-
between different time schedules of administra- dation, although cyclic administration did cause
tion (daytime, nighttime, or continuously over rhythms in energy expenditure and substrate oxi-
24 h) (Stratton et al., 2003). Cyclic nighttime dation (increased energy expenditure and carbo-
administration of enteral nutrition caused lower hydrate oxidation during administration). Cyclic
oxygen consumption, lower nitrogen balance, administration also reduced liver enzyme elevation
and lower 24-h catecholamine excretion com- compared to continuous administration, especially
pared to continuous administration in 18 patients in patients with low baseline bilirubin levels (Stout
who had undergone surgery for malignancy in the and Cober, 2010). In support, animal experiments
buccopharynx or larynx (Campbell et al., 1990). show that cyclic administration leads to less
Patients in a vegetative state showed a shift of hepatic steatosis than continuous administration
the daily peak of body temperature from 8 PM (Lanza-Jacoby, 1986). Another, less favorable,
to 4 AM after nighttime compared to daytime effect of nightly cyclic parenteral nutrition is
administration (Nishimura et al., 1992) and a sim- increased urine production at night (Boncompain-
ilar shift in their daily cortisol peak (Saito et al., Gerard et al., 2000).
1989). In the intensive care unit setting, the Thus, cyclic administration of parenteral nutri-
head-of-bed elevation influences microaspiration tion mainly serves to prevent hepatic steatosis
risk, with the lowest risk of aspiration with a and plasma liver enzyme elevations. In humans,
head-of-bed elevation of 30–45 (Kattelmann no studies have compared daytime administration
369

with nighttime administration. Again, practical (Rana et al., 2003). The first two factors are most
arguments will probably influence the choice for likely to cause the altered rhythm in the occur-
a schedule of administration. Many patients pre- rence of myocardial infarctions, as one study in
fer administration at night because of the large patients with type 2 diabetes did find a diurnal
infusion system involved, although the increased rhythm of platelet activation with an early morn-
nighttime urine production is a disadvantage. ing peak (Spano et al., 1993), and the same
rhythm has been found in healthy individuals
(Dalby et al., 2000; Feuring et al., 2009).
Disturbed rhythms and type 2 diabetes Third, rhythms of peripheral clock gene expres-
sion may be affected in type 2 diabetes as patients
Type 2 diabetes is considered a multifactorial dis- with type 2 diabetes show an altered daily rhythm
ease caused by genetic factors and obesity due to of clock gene expression in circulating leukocytes
excess caloric intake and reduced physical activ- (Ando et al., 2009a). Altered adipose clock gene
ity. However, several lines of evidence suggest a expression rhythms were detected in a murine
role of the circadian system in the pathophysiol- model of type 2 diabetes (Ando et al., 2005, 2011),
ogy of type 2 diabetes. We previously discussed but not in two different rat models (Ando et al.,
how disturbed day–night rhythms can lead to obe- 2009b; Motosugi et al., 2011). In the only human
sity and type 2 diabetes. Interestingly, a number study thus far, no differences in the rhythms of
of alterations in the 24-h rhythms of patients with white adipose tissue clock gene expression were
type 2 diabetes have been observed. detected between lean, obese, and diabetic
First, the daily rhythm in glucose metabolism individuals (Otway et al., 2011).
seems to be different in patients with type 2 diabe- In summary, mismatches in day–night rhythms
tes. Whereas healthy individuals show a decrease can lead to obesity and type 2 diabetes, while altered
of insulin sensitivity toward the night (Jarrett circadian rhythms are observed in type 2 diabetes.
et al., 1972; Whichelow et al., 1974), insulin sensi- Thus, the circadian system is likely to be involved
tivity (as measured with a 72-h hyperglycemic in the pathophysiology of type 2 diabetes. But if
clamp) increases toward the night in patients with so, what could be the pathophysiological mecha-
type 2 diabetes (Boden et al., 1996). Also, when nism? Does the molecular clock get damaged in
patients with type 2 diabetes are fed three identi- SCN, liver, pancreas, or adipose tissue, or do pro-
cal meals, morning glucose excursions are higher blems occur in the communication between the dif-
than evening glucose excursions (Peter et al., ferent clocks? Is the circadian system a potential
2010), again pointing to an increased insulin sen- target in the treatment of type 2 diabetes? Future
sitivity (or at least increased glucose tolerance) research will have to answer these questions.
toward the evening. On the other hand, daily
rhythms in glucose tolerance have never been
compared between patients with type 2 diabetes Conclusion
and controls in a single design.
Second, a 24-h rhythm exists in the occurrence In order to adjust to the daily light–dark rhythm,
of myocardial infarctions in the population at humans possess a central brain clock in the
large, with an increased incidence of myocardial SCN. Furthermore, virtually all cells in the body
infarctions in the early morning. Patients with contain molecular clocks. These peripheral clocks
type 2 diabetes do not show this rhythm, which are synchronized both by the SCN and through
could be due to altered rhythms in the activity metabolic signals. Daily rhythms are present in
of the autonomic nervous system, circulating appetite regulation, digestion and absorption,
hemostatic factors, and/or platelet aggregation and carbohydrate and lipid metabolism. Thus,
370

meal patterns may strongly influence human 2 diabetic Goto-Kakizaki rats. Clinical and Experimental
health. Strong clues indicate that it is better to Hypertension, 31(3), 201–207.
Ando, H., Yanagihara, H., Hayashi, Y., Obi, Y., Tsuruoka, S.,
maintain a rhythm of daytime food intake and Takamura, T., et al. (2005). Rhythmic messenger ribonucleic
nighttime sleep than to eat at night. A limited acid expression of clock genes and adipocytokines in mouse
meal frequency helps to prevent caries, but evi- visceral adipose tissue. Endocrinology, 146(12), 5631–5636.
dence for additional health effects remains scarce. Arase, K., York, D. A., Shimizu, H., Shargill, N., &
Eating regular meals including a daily breakfast Bray, G. A. (1988). Effects of corticotropin-releasing factor
on food intake and brown adipose tissue thermogenesis in
appears to benefit health, but more intervention rats. The American Journal of Physiology, 255(3 Pt. 1),
trials are needed. E255–E259.
In our modern society, the synchrony between Baker, I. A., & Jarrett, R. J. (1972). Diurnal variation in the
clock gene rhythms, metabolic rhythms, and blood-sugar and plasma-insulin response to tolbutamide.
behavioral rhythms is easily disturbed due to the Lancet, 2(7784), 945–947.
Balsalobre, A., Brown, S. A., Marcacci, L., Tronche, F.,
around-the-clock availability of artificial light- and Kellendonk, C., Reichardt, H. M., et al. (2000). Resetting
high-caloric food. This may contribute to the of circadian time in peripheral tissues by glucocorticoid sig-
increasing incidence of obesity and type 2 diabetes. naling. Science, 289(5488), 2344–2347.
The question whether in patients with gastrointes- Balsalobre, A., Damiola, F., & Schibler, U. (1998). A serum
tinal disease, the desynchrony due to nighttime shock induces circadian gene expression in mammalian tis-
sue culture cells. Cell, 93(6), 929–937.
administration of (par)enteral nutrition disturbs Barter, P. J., Carroll, K. F., & Nestel, P. J. (1971). Diurnal
metabolism needs further investigation. Undoubt- fluctuations in triglyceride, free fatty acids, and insulin
edly, future research will shed more light on the during sucrose consumption and insulin infusion in man.
fascinating interaction between the circadian The Journal of Clinical Investigation, 50(3), 583–591.
timing system, metabolism, and human health. Bass, J., & Takahashi, J. S. (2010). Circadian integration of
metabolism and energetics. Science, 330(6009), 1349–1354.
Bellisle, F., McDevitt, R., & Prentice, A. M. (1997). Meal fre-
quency and energy balance. The British Journal of Nutrition,
References 77(Suppl. 1), S57–S70.
Bhutani, S., & Varady, K. A. (2009). Nibbling versus feasting:
Which meal pattern is better for heart disease prevention?
Abrahamson, E. E., Leak, R. K., & Moore, R. Y. (2001). The
Nutrition Reviews, 67(10), 591–598.
suprachiasmatic nucleus projects to posterior hypothalamic
Biston, P., Van Cauter, E., Ofek, G., Linkowski, P.,
arousal systems. Neuroreport, 12(2), 435–440.
Polonsky, K. S., & Degaute, J. P. (1996). Diurnal variations
Ahima, R. S., Prabakaran, D., Mantzoros, C., Qu, D.,
in cardiovascular function and glucose regulation in normo-
Lowell, B., Maratos-Flier, E., et al. (1996). Role of leptin
tensive humans. Hypertension, 28(5), 863–871.
in the neuroendocrine response to fasting. Nature, 382
Boden, G., Chen, X., & Urbain, J. L. (1996). Evidence for a
(6588), 250–252.
circadian rhythm of insulin sensitivity in patients with
Allaman-Pillet, N., Roduit, R., Oberson, A., Abdelli, S.,
NIDDM caused by cyclic changes in hepatic glucose produc-
Ruiz, J., Beckmann, J. S., et al. (2004). Circadian regulation
tion. Diabetes, 45(8), 1044–1050.
of islet genes involved in insulin production and secretion.
Boncompain-Gerard, M., Robert, D., Fouque, D., & Hadj-
Molecular and Cellular Endocrinology, 226(1–2), 59–66.
Aissa, A. (2000). Renal function and urinary excretion of
Ando, H., Kumazaki, M., Motosugi, Y., Ushijima, K.,
electrolytes in patients receiving cyclic parenteral nutrition.
Maekawa, T., Ishikawa, E., et al. (2011). Impairment of
Journal of Parenteral and Enteral Nutrition, 24(4), 234–239.
peripheral circadian clocks precedes metabolic abnormalities
Bos, N. P., & Mirmiran, M. (1990). Circadian rhythms in spon-
in ob/ob mice. Endocrinology, 152(4), 1347–1354.
taneous neuronal discharges of the cultured suprachiasmatic
Ando, H., Takamura, T., Matsuzawa-Nagata, N., Shima, K. R.,
nucleus. Brain Research, 511(1), 158–162.
Eto, T., Misu, H., et al. (2009). Clock gene expression in
Bostwick, J., Nguyen, D., Cornelissen, G., Halberg, F., &
peripheral leucocytes of patients with type 2 diabetes.
Hoogerwerf, W. A. (2010). Effects of acute and chronic
Diabetologia, 52(2), 329–335.
STZ-induced diabetes on clock gene expression and feeding
Ando, H., Ushijima, K., Yanagihara, H., Hayashi, Y.,
in the gastrointestinal tract. Molecular and Cellular Bio-
Takamura, T., Kaneko, S., et al. (2009). Clock gene expres-
chemistry, 338(1–2), 203–213.
sion in the liver and adipose tissues of non-obese type
371

Boyle, P. J., Avogaro, A., Smith, L., Bier, D. M., Pappu, A. S., Carroll, K. F., & Nestel, P. J. (1973). Diurnal variation in glu-
Illingworth, D. R., et al. (1992). Role of GH in regulating cose tolerance and in insulin secretion in man. Diabetes, 22
nocturnal rates of lipolysis and plasma mevalonate levels (5), 333–348.
in normal and diabetic humans. The American Journal of Cella, L. K., Van, C. E., & Schoeller, D. A. (1995). Diurnal
Physiology, 263(1 Pt. 1), E168–E172. rhythmicity of human cholesterol synthesis: Normal pattern
Brown, S. A., Kunz, D., Dumas, A., Westermark, P. O., and adaptation to simulated “jet lag” The American Journal
Vanselow, K., Tilmann-Wahnschaffe, A., et al. (2008). of Physiology, 269(3 Pt. 1), E489–E498.
Molecular insights into human daily behavior. Proceedings Cizza, G., Marincola, P., Mattingly, M., Williams, L.,
of the National Academy of Sciences of the United States of Mitler, M., Skarulis, M., et al. (2010). Treatment of obesity
America, 105(5), 1602–1607. with extension of sleep duration: A randomized, prospec-
Brown, S. A., Zumbrunn, G., Fleury-Olela, F., Preitner, N., & tive, controlled trial. Clinical Trials, 7(3), 274–285.
Schibler, U. (2002). Rhythms of mammalian body tempera- Clark, J. T., Kalra, P. S., Crowley, W. R., & Kalra, S. P. (1984).
ture can sustain peripheral circadian clocks. Current Biol- Neuropeptide Y and human pancreatic polypeptide stimulate
ogy, 12(18), 1574–1583. feeding behavior in rats. Endocrinology, 115(1), 427–429.
Brownson, R. C., Boehmer, T. K., & Luke, D. A. (2005). Clore, J. N., Nestler, J. E., & Blackard, W. G. (1989). Sleep-
Declining rates of physical activity in the United States: associated fall in glucose disposal and hepatic glucose output
What are the contributors? Annual Review of Public Health, in normal humans. Putative signaling mechanism linking
26, 421–443. peripheral and hepatic events. Diabetes, 38(3), 285–290.
Buhr, E. D., Yoo, S. H., & Takahashi, J. S. (2010). Tempera- Dalby, M. C., Davidson, S. J., Burman, J. F., & Davies, S. W.
ture as a universal resetting cue for mammalian circadian (2000). Diurnal variation in platelet aggregation iwth the
oscillators. Science, 330(6002), 379–385. PFA-100 platelet function analyser. Platelets, 11(6), 320–324.
Buijs, R. M., Chun, S. J., Niijima, A., Romijn, H. J., & Damiola, F., Le Minh, N., Preitner, N., Kornmann, B., Fleury-
Nagai, K. (2001). Parasympathetic and sympathetic control Olela, F., & Schibler, U. (2000). Restricted feeding
of the pancreas: A role for the suprachiasmatic nucleus uncouples circadian oscillators in peripheral tissues from
and other hypothalamic centers that are involved in the reg- the central pacemaker in the suprachiasmatic nucleus. Genes
ulation of food intake. The Journal of Comparative Neurol- & Development, 14(23), 2950–2961.
ogy, 431(4), 405–423. Dawes, C. (1975). The effect of food deprivation on the circa-
Buijs, R. M., La Fleur, S. E., Wortel, J., Van Heyningen, C., dian rhythms in flow rate and protein concentration in
Zuiddam, L., Mettenleiter, T. C., et al. (2003). The human parotid saliva. Archives of Oral Biology, 20(8),
suprachiasmatic nucleus balances sympathetic and parasympa- 547–549.
thetic output to peripheral organs through separate pre- Dibner, C., Schibler, U., & Albrecht, U. (2010). The mamma-
autonomic neurons. The Journal of Comparative Neurology, lian circadian timing system: Organization and coordination
464(1), 36–48. of central and peripheral clocks. Annual Review of Physiol-
Cailotto, C., La Fleur, S. E., van Heijningen, C., Wortel, J., ogy, 72, 517–549.
Kalsbeek, A., Feenstra, M., et al. (2005). The Donga, E., van Dijk, M., van Dijk, J. G., Biermasz, N. R.,
suprachiasmatic nucleus controls the daily variation of Lammers, G. J., van Kralingen, K. W., et al. (2010). A single
plasma glucose via the autonomic output to the liver: Are night of partial sleep deprivation induces insulin resistance
the clock genes involved? The European Journal of Neuro- in multiple metabolic pathways in healthy subjects. The
science, 22(10), 2531–2540. Journal of Clinical Endocrinology and Metabolism, 95(6),
Cailotto, C., Lei, J., van der Vliet, J., van Heijningen, C., van 2963–2968.
Eden, C. G., Kalsbeek, A., et al. (2009). Effects of nocturnal light Edwards, C. M., Abusnana, S., Sunter, D., Murphy, K. G.,
on (clock) gene expression in peripheral organs: A role for the Ghatei, M. A., & Bloom, S. R. (1999). The effect of the
autonomic innervation of the liver. PLoS One, 4(5), e5650. orexins on food intake: Comparison with neuropeptide Y,
Cailotto, C., van Heijningen, C., van der Vliet, J., van der melanin-concentrating hormone and galanin. The Journal
Plasse, G., Habold, C., Kalsbeek, A., et al. (2008). Daily of Endocrinology, 160(3), R7–R12.
rhythms in metabolic liver enzymes and plasma glucose Edwards, P. A., Muroya, H., & Gould, R. G. (1972). In vivo
require a balance in the autonomic output to the liver. demonstration of the circadian thythm of cholesterol biosyn-
Endocrinology, 149(4), 1914–1925. thesis in the liver and intestine of the rat. Journal of Lipid
Campbell, I. T., Morton, R. P., Macdonald, I. A., Judd, S., Research, 13(3), 396–401.
Shapiro, L., & Stell, P. M. (1990). Comparison of the meta- Fan, W., Boston, B. A., Kesterson, R. A., Hruby, V. J., &
bolic effects of continuous postoperative enteral feeding and Cone, R. D. (1997). Role of melanocortinergic neurons in
feeding at night only. The American Journal of Clinical feeding and the agouti obesity syndrome. Nature, 385
Nutrition, 52(6), 1107–1112. (6612), 165–168.
372

Farshchi, H. R., Taylor, M. A., & Macdonald, I. A. (2005a). Jacoby, E. R., Cueto, S., & Pollitt, E. (1998). When science
Beneficial metabolic effects of regular meal frequency on and politics listen to each other: Good prospects from a
dietary thermogenesis, insulin sensitivity, and fasting lipid new school breakfast program in Peru. The American Jour-
profiles in healthy obese women. The American Journal of nal of Clinical Nutrition, 67(4), 795S–797S.
Clinical Nutrition, 81(1), 16–24. Jarrett, R. J., Baker, I. A., Keen, H., & Oakley, N. W. (1972).
Farshchi, H. R., Taylor, M. A., & Macdonald, I. A. (2005b). Diurnal variation in oral glucose tolerance: Blood sugar
Deleterious effects of omitting breakfast on insulin sensitiv- and plasma insulin levels morning, afternoon, and evening.
ity and fasting lipid profiles in healthy lean women. The British Medical Journal, 1(5794), 199–201.
American Journal of Clinical Nutrition, 81(2), 388–396. Kalsbeek, A., Foppen, E., Schalij, I., van Heijningen, C., van
Feneberg, R., & Lemmer, B. (2004). Circadian rhythm of glu- der Vliet, J., Fliers, E., et al. (2008). Circadian control of
cose uptake in cultures of skeletal muscle cells and the daily plasma glucose rhythm: An interplay of GABA
adipocytes in Wistar-Kyoto, Wistar, Goto-Kakizaki, and and glutamate. PLoS One, 3(9), e3194.
spontaneously hypertensive rats. Chronobiology Interna- Kalsbeek, A., La Fleur, S., van Heijningen, C., & Buijs, R. M.
tional, 21(4–5), 521–538. (2004). Suprachiasmatic GABAergic inputs to the para-
Feuring, M., Wehling, M., Ruf, A., & Schultz, A. (2009). Circa- ventricular nucleus control plasma glucose concentrations
dian variation of platelet function measured with the PFA- in the rat via sympathetic innervation of the liver. The Jour-
100. Platelets, 20(7), 466–470. nal of Neuroscience, 24(35), 7604–7613.
Fliers, E., Kreier, F., Voshol, P. J., Havekes, L. M., Kalsbeek, A., Ruiter, M., La Fleur, S. E., van Heijningen, C.,
Sauerwein, H. P., Kalsbeek, A., et al. (2003). White adipose & Buijs, R. M. (2003). The diurnal modulation of hormonal
tissue: Getting nervous. Journal of Neuroendocrinology, 15 responses in the rat varies with different stimuli. Journal of
(11), 1005–1010. Neuroendocrinology, 15(12), 1144–1155.
Gibson, T., & Jarrett, R. J. (1972). Diurnal variation in insulin Kandelman, D. (1997). Sugar, alternative sweeteners and meal
sensitivity. Lancet, 2(7784), 947–948. frequency in relation to caries prevention: New perspectives.
Gibson, T., Stimmler, L., Jarrett, R. J., Rutland, P., & Shiu, M. The British Journal of Nutrition, 77(Suppl. 1), S121–S128.
(1975). Diurnal variation in the effects of insulin on blood Kattelmann, K. K., Hise, M., Russell, M., Charney, P.,
glucose, plasma non-esterified fatty acids and growth hor- Stokes, M., & Compher, C. (2006). Preliminary evidence
mone. Diabetologia, 11(1), 83–88. for a medical nutrition therapy protocol: Enteral feedings
Goo, R. H., Moore, J. G., Greenberg, E., & Alazraki, N. P. for critically ill patients. Journal of the American Dietetic
(1987). Circadian variation in gastric emptying of meals in Association, 106(8), 1226–1241.
humans. Gastroenterology, 93(3), 515–518. Kohsaka, A., & Bass, J. (2007). A sense of time: How molecu-
Green, D. J., & Gillette, R. (1982). Circadian rhythm of firing lar clocks organize metabolism. Trends in Endocrinology
rate recorded from single cells in the rat suprachiasmatic and Metabolism, 18(1), 4–11.
brain slice. Brain Research, 245(1), 198–200. Kornmann, B., Schaad, O., Bujard, H., Takahashi, J. S., &
Green, C. B., Takahashi, J. S., & Bass, J. (2008). The meter of Schibler, U. (2007). System-driven and oscillator-dependent
metabolism. Cell, 134(5), 728–742. circadian transcription in mice with a conditionally active
Groos, G., & Hendriks, J. (1982). Circadian rhythms in electri- liver clock. PLoS Biology, 5(2), e34.
cal discharge of rat suprachiasmatic neurones recorded Kreier, F., Kap, Y. S., Mettenleiter, T. C., van, H. C., van der
in vitro. Neuroscience Letters, 34(3), 283–288. Vliet, J., Kalsbeek, A., et al. (2006). Tracing from fat tissue, liver,
Guo, H., Brewer, J. M., Champhekar, A., Harris, R. B., & and pancreas: A neuroanatomical framework for the role of the
Bittman, E. L. (2005). Differential control of peripheral circa- brain in type 2 diabetes. Endocrinology, 147(3), 1140–1147.
dian rhythms by suprachiasmatic-dependent neural signals. Kumar, D., Wingate, D., & Ruckebusch, Y. (1986). Circadian
Proceedings of the National Academy of Sciences of the United variation in the propagation velocity of the migrating motor
States of America, 102(8), 3111–3116. complex. Gastroenterology, 91(4), 926–930.
Herzog, E. D., Takahashi, J. S., & Block, G. D. (1998). Clock Kuriyama, K., Sasahara, K., Kudo, T., & Shibata, S. (2004).
controls circadian period in isolated suprachiasmatic nucleus Daily injection of insulin attenuated impairment of liver cir-
neurons. Nature Neuroscience, 1(8), 708–713. cadian clock oscillation in the streptozotocin-treated dia-
Hoogerwerf, W. A. (2010). Role of clock genes in gastrointes- betic mouse. FEBS Letters, 572(1–3), 206–210.
tinal motility. American Journal of Physiology. Gastrointesti- La Fleur, S. E. (2003). Daily rhythms in glucose metabolism:
nal and Liver Physiology, 299(3), G549–G555. Suprachiasmatic nucleus output to peripheral tissue. Journal
Hoogerwerf, W. A., Hellmich, H. L., Cornelissen, G., of Neuroendocrinology, 15(3), 315–322.
Halberg, F., Shahinian, V. B., Bostwick, J., et al. (2007). La Fleur, S. E., Kalsbeek, A., Wortel, J., & Buijs, R. M.
Clock gene expression in the murine gastrointestinal (2000). Polysynaptic neural pathways between the hypothal-
tract: Endogenous rhythmicity and effects of a feeding regi- amus, including the suprachiasmatic nucleus, and the liver.
men. Gastroenterology, 133(4), 1250–1260. Brain Research, 871(1), 50–56.
373

La Fleur, S. E., Kalsbeek, A., Wortel, J., Fekkes, M. L., & Marcheva, B., Ramsey, K. M., Buhr, E. D., Kobayashi, Y.,
Buijs, R. M. (2001). A daily rhythm in glucose tolerance: A role Su, H., Ko, C. H., et al. (2010). Disruption of the clock com-
for the suprachiasmatic nucleus. Diabetes, 50(6), 1237–1243. ponents CLOCK and BMAL1 leads to hypoinsulinaemia
Lamia, K. A., Sachdeva, U. M., DiTacchio, L., Williams, E. C., and diabetes. Nature, 466(7306), 627–631.
Alvarez, J. G., Egan, D. F., et al. (2009). AMPK regulates Melani, F., Verrillo, A., Marasco, M., Rivellese, A., Osorio, J.,
the circadian clock by cryptochrome phosphorylation and & Bertolini, M. G. (1976). Diurnal variation in blood sugar
degradation. Science, 326(5951), 437–440. and serum insulin in response to glucose and/or glucagon
Lamia, K. A., Storch, K. F., & Weitz, C. J. (2008). Physiologi- in healthy subjects. Hormone and Metabolic Research, 8
cal significance of a peripheral tissue circadian clock. Pro- (2), 85–88.
ceedings of the National Academy of Sciences of the United Metheny, N. A. (2006). Preventing respiratory complications
States of America, 105(39), 15172–15177. of tube feedings: Evidence-based practice. American Journal
Lanza-Jacoby, S. (1986). Effect of continuous and discontinu- of Critical Care, 15(4), 360–369.
ous intravenous or intragastric total parenteral nutrition in Metheny, N. A., Clouse, R. E., Chang, Y. H., Stewart, B. J.,
rats on serum lipids, liver lipids and liver lipogenic rates. Oliver, D. A., & Kollef, M. H. (2006). Tracheobronchial
The Journal of Nutrition, 116(5), 733–741. aspiration of gastric contents in critically ill tube-fed
Le Minh, N., Damiola, F., Tronche, F., Schutz, G., & patients: Frequency, outcomes, and risk factors. Critical
Schibler, U. (2001). Glucocorticoid hormones inhibit food- Care Medicine, 34(4), 1007–1015.
induced phase-shifting of peripheral circadian oscillators. Moore, J. G. (1991). Circadian dynamics of gastric acid secre-
The EMBO Journal, 20(24), 7128–7136. tion and pharmacodynamics of H2 receptor blockade.
Lee, A., Ader, M., Bray, G. A., & Bergman, R. N. (1992). Annals of the New York Academy of Sciences, 618, 150–158.
Diurnal variation in glucose tolerance. Cyclic suppression Moore, R. Y., & Eichler, V. B. (1972). Loss of a circadian
of insulin action and insulin secretion in normal-weight, adrenal corticosterone rhythm following suprachiasmatic
but not obese, subjects. Diabetes, 41(6), 750–759. lesions in the rat. Brain Research, 42(1), 201–206.
Legradi, G., & Lechan, R. M. (1999). Agouti-related protein Morgan, L. M., Aspostolakou, F., Wright, J., & Gama, R.
containing nerve terminals innervate thyrotropin-releasing (1999). Diurnal variations in peripheral insulin resistance
hormone neurons in the hypothalamic paraventricular and plasma non-esterified fatty acid concentrations: A possi-
nucleus. Endocrinology, 140(8), 3643–3652. ble link? Annals of Clinical Biochemistry, 36(Pt. 4), 447–450.
LeSauter, J., Hoque, N., Weintraub, M., Pfaff, D. W., & Morin, L. P., & Allen, C. N. (2006). The circadian visual sys-
Silver, R. (2009). Stomach ghrelin-secreting cells as food- tem, 2005. Brain Research Reviews, 51(1), 1–60.
entrainable circadian clocks. Proceedings of the National Motosugi, Y., Ando, H., Ushijima, K., Maekawa, T.,
Academy of Sciences of the United States of America, 106 Ishikawa, E., Kumazaki, M., et al. (2011). Tissue-dependent
(32), 13582–13587. alterations of the clock gene expression rhythms in leptin-
Lichtman, S. W., Pisarska, K., Berman, E. R., Pestone, M., resistant Zucker diabetic fatty rats. Chronobiology Interna-
Dowling, H., Offenbacher, E., et al. (1992). Discrepancy tional, 28(10), 968–972.
between self-reported and actual caloric intake and exercise Muhlbauer, E., Wolgast, S., Finckh, U., Peschke, D., &
in obese subjects. The New England Journal of Medicine, Peschke, E. (2004). Indication of circadian oscillations in
327(27), 1893–1898. the rat pancreas. FEBS Letters, 564(1–2), 91–96.
Loboda, A., Kraft, W. K., Fine, B., Joseph, J., Nebozhyn, M., Nagai, K., Nishio, T., Nakagawa, H., Nakamura, S., & Fukuda, Y.
Zhang, C., et al. (2009). Diurnal variation of the human adi- (1978). Effect of bilateral lesions of the suprachiasmatic nuclei
pose transcriptome and the link to metabolic disease. BMC on the circadian rhythm of food-intake. Brain Research, 142
Medical Genomics, 2, 7. (2), 384–389.
Lund, T. M., Torsvik, H., Falch, D., Christophersen, B., Nakamura, W., Honma, S., Shirakawa, T., & Honma, K.
Skardal, R., & Gullestad, L. (2002). Effect of morning ver- (2002). Clock mutation lengthens the circadian period with-
sus evening intake of simvastatin on the serum cholesterol out damping rhythms in individual SCN neurons. Nature
level in patients with coronary artery disease. The American Neuroscience, 5(5), 399–400.
Journal of Cardiology, 90(7), 784–786. Narducci, F., Bassotti, G., Gaburri, M., & Morelli, A. (1987).
Mann, J. (1997). Meal frequency and plasma lipids and Twenty four hour manometric recording of colonic motor
lipoproteins. The British Journal of Nutrition, 77(Suppl. 1), activity in healthy man. Gut, 28(1), 17–25.
S83–S90. Niijima, A., Nagai, K., Nagai, N., & Nakagawa, H. (1992).
Maouyo, D., Sarfati, P., Guan, D., Morisset, J., & Adelson, J. W. Light enhances sympathetic and suppresses vagal outflows
(1993). Circadian rhythm of exocrine pancreatic secretion in and lesions including the suprachiasmatic nucleus eliminate
rats: Major and minor cycles. The American Journal of these changes in rats. Journal of the Autonomic Nervous
Physiology, 264(4 Pt. 1), G792–G800. System, 40(2), 155–160.
374

Nishimura, K., Kato, H., & Saito, M. (1992). Effects of cyclic effects of methionyl growth hormone and of acipimox, an
and continuous total enteral nutrition on 24-h rhythms of inhibitor of lipolysis. Journal of Endocrinological Investiga-
body temperature and urinary excretions. Journal of tion, 15(2), 85–91.
Nutritional Science and Vitaminology, 38(2), 117–125. Powell, C. A., Walker, S. P., Chang, S. M., & Grantham-
Nishio, T., Shiosaka, S., Nakagawa, H., Sakumoto, T., & McGregor, S. M. (1998). Nutrition and education:
Satoh, K. (1979). Circadian feeding rhythm after hypotha- A randomized trial of the effects of breakfast in rural pri-
lamic knife-cut isolating suprachiasmatic nucleus. Physiol- mary school children. The American Journal of Clinical
ogy & Behavior, 23(4), 763–769. Nutrition, 68(4), 873–879.
Oishi, K., Fukui, H., & Ishida, N. (2000). Rhythmic expression Rampersaud, G. C., Pereira, M. A., Girard, B. L., Adams, J.,
of BMAL1 mRNA is altered in Clock mutant mice: Differ- & Metzl, J. D. (2005). Breakfast habits, nutritional status,
ential regulation in the suprachiasmatic nucleus and periph- body weight, and academic performance in children and
eral tissues. Biochemical and Biophysical Research adolescents. Journal of the American Dietetic Association,
Communications, 268(1), 164–171. 105(5), 743–760.
Ollmann, M. M., Wilson, B. D., Yang, Y. K., Kerns, J. A., Ramsey, K. M., Yoshino, J., Brace, C. S., Abrassart, D.,
Chen, Y., Gantz, I., et al. (1997). Antagonism of central mel- Kobayashi, Y., Marcheva, B., et al. (2009). Circadian clock
anocortin receptors in vitro and in vivo by agouti-related feedback cycle through NAMPT-mediated NADþ biosyn-
protein. Science, 278(5335), 135–138. thesis. Science, 324(5927), 651–654.
Otway, D. T., Mantele, S., Bretschneider, S., Wright, J., Rana, J. S., Mukamal, K. J., Morgan, J. P., Muller, J. E., &
Trayhurn, P., Skene, D. J., et al. (2011). Rhythmic diurnal Mittleman, M. A. (2003). Circadian variation in the onset
gene expression in human adipose tissue from individuals of myocardial infarction: Effect of duration of diabetes. Dia-
who are lean, overweight, and type 2 diabetic. Diabetes, 60 betes, 52(6), 1464–1468.
(5), 1577–1581. Rao, S. S., Sadeghi, P., Beaty, J., Kavlock, R., & Ackerson, K.
Palmer, M. A., Capra, S., & Baines, S. K. (2009). Association (2001). Ambulatory 24-h colonic manometry in healthy
between eating frequency, weight, and health. Nutrition humans. American Journal of Physiology. Gastrointestinal
Reviews, 67(7), 379–390. and Liver Physiology, 280(4), G629–G639.
Pan, A., Schernhammer, E. S., Sun, Q., & Hu, F. B. (2011). Reppert, S. M., Perlow, M. J., Ungerleider, L. G., Mishkin, M.,
Rotating night shift work and risk of type 2 diabetes: Two Tamarkin, L., Orloff, D. G., et al. (1981). Effects of damage
prospective cohort studies in women. PLoS Medicine, 8 to the suprachiasmatic area of the anterior hypothalamus on
(12), e1001141. the daily melatonin and cortisol rhythms in the rhesus mon-
Pan, X., Terada, T., Irie, M., Saito, H., & Inui, K. (2002). Diur- key. The Journal of Neuroscience, 1(12), 1414–1425.
nal rhythm of Hþ-peptide cotransporter in rat small intes- Roenneberg, T., Allebrandt, K. V., Merrow, M., & Vetter, C.
tine. American Journal of Physiology. Gastrointestinal and (2012). Social jetlag and obesity. Current Biology, 22(10),
Liver Physiology, 283(1), G57–G64. 939–943.
Pan, X., Terada, T., Okuda, M., & Inui, K. (2004). The diurnal Roesler, W. J., McFie, P. J., & Dauvin, C. (1992). The liver-
rhythm of the intestinal transporters SGLT1 and PEPT1 is enriched transcription factor D-site-binding protein
regulated by the feeding conditions in rats. The Journal of activates the promoter of the phosphoenolpyruvate car-
Nutrition, 134(9), 2211–2215. boxykinase gene in hepatoma cells. The Journal of
Panda, S., Antoch, M. P., Miller, B. H., Su, A. I., Biological Chemistry, 267(29), 21235–21243.
Schook, A. B., Straume, M., et al. (2002). Coordinated tran- Romon, M., Le, F. C., Lebel, P., Edme, J. L., Fruchart, J. C., &
scription of key pathways in the mouse by the circadian Dallongeville, J. (1997). Circadian variation of postprandial
clock. Cell, 109(3), 307–320. lipemia. The American Journal of Clinical Nutrition, 65(4),
Peschke, E., & Peschke, D. (1998). Evidence for a circadian 934–940.
rhythm of insulin release from perifused rat pancreatic islets. Rossi, M., Kim, M. S., Morgan, D. G., Small, C. J.,
Diabetologia, 41(9), 1085–1092. Edwards, C. M., Sunter, D., et al. (1998). A C-terminal frag-
Peter, R., Dunseath, G., Luzio, S. D., Chudleigh, R., ment of Agouti-related protein increases feeding and
Roy, C. S., & Owens, D. R. (2010). Daytime variability of antagonizes the effect of alpha-melanocyte stimulating hor-
postprandial glucose tolerance and pancreatic B-cell func- mone in vivo. Endocrinology, 139(10), 4428–4431.
tion using 12-h profiles in persons with Type 2 diabetes. Dia- Rutter, J., Reick, M., Wu, L. C., & McKnight, S. L. (2001).
betic Medicine, 27(3), 266–273. Regulation of clock and NPAS2 DNA binding by the redox
Plata-Salaman, C. R., & Oomura, Y. (1987). Food intake state of NAD cofactors. Science, 293(5529), 510–514.
dependence on acute changes in light schedule. Physiology Ruxton, C. H., & Kirk, T. R. (1997). Breakfast: A review of
& Behavior, 41(2), 135–140. associations with measures of dietary intake, physiology and
Pontiroli, A. E., Lanzi, R., Monzani, M., Musatti, L., biochemistry. The British Journal of Nutrition, 78(2), 199–213.
Guglielmone, C., & Pozza, G. (1992). Plasma free fatty acids Sadacca, L. A., Lamia, K. A., deLemos, A. S., Blum, B., &
and serum insulin in subjects feeding at 12-hour intervals; Weitz, C. J. (2011). An intrinsic circadian clock of the
375

pancreas is required for normal insulin release and glucose variations in platelet aggregability in non insulin dependent
homeostasis in mice. Diabetologia, 54(1), 120–124. diabetes patients (NIDDM). La Clinica Terapeutica, 142(1),
Saito, M., Kato, H., & Suda, M. (1980). Circadian rhythm of 19–22.
intestinal disaccharidases of rats fed with a diurnal periodic- Spiegel, K., Tasali, E., Leproult, R., & Van Cauter, E. (2009).
ity. The American Journal of Physiology, 238(2), G97–G101. Effects of poor and short sleep on glucose metabolism and
Saito, M., Murakami, E., & Suda, M. (1976). Circadian obesity risk. Nature Reviews. Endocrinology, 5(5), 253–261.
rhythms in disaccharidases of rat small intestine and its rela- Spiegel, K., Tasali, E., Penev, P., & Van Cauter, E. (2004).
tion to food intake. Biochimica et Biophysica Acta, 421(1), Brief communication: Sleep curtailment in healthy young
177–179. men is associated with decreased leptin levels, elevated
Saito, M., Nishimura, K., & Kato, H. (1989). Modifications of ghrelin levels, and increased hunger and appetite. Annals
circadian cortisol rhythm by cyclic and continuous total of Internal Medicine, 141(11), 846–850.
enteral nutrition. Journal of Nutritional Science and Stanley, B. G., & Leibowitz, S. F. (1985). Neuropeptide Y
Vitaminology, 35(6), 639–647. injected in the paraventricular hypothalamus: A powerful
Saito, Y., Yoshida, S., Nakaya, N., Hata, Y., & Goto, Y. stimulant of feeding behavior. Proceedings of the National
(1991). Comparison between morning and evening doses Academy of Sciences of the United States of America, 82
of simvastatin in hyperlipidemic subjects. A double-blind (11), 3940–3943.
comparative study. Arteriosclerosis and Thrombosis, 11(4), Stokkan, K. A., Yamazaki, S., Tei, H., Sakaki, Y., &
816–826. Menaker, M. (2001). Entrainment of the circadian clock in
Sakurai, T., Amemiya, A., Ishii, M., Matsuzaki, I., the liver by feeding. Science, 291(5503), 490–493.
Chemelli, R. M., Tanaka, H., et al. (1998). Orexins and Stout, S. M., & Cober, M. P. (2010). Metabolic effects of cyclic
orexin receptors: A family of hypothalamic neuropeptides parenteral nutrition infusion in adults and children. Nutri-
and G protein-coupled receptors that regulate feeding tion in Clinical Practice, 25(3), 277–281.
behavior. Cell, 92(4), 573–585. Stratton, R. J., & Elia, M. (1999). The effects of enteral tube
Salle, A., Ryan, M., & Ritz, P. (2006). Underreporting of food feeding and parenteral nutrition on appetite sensations and
intake in obese diabetic and nondiabetic patients. Diabetes food intake in health and disease. Clinical Nutrition, 18(2),
Care, 29(12), 2726–2727. 63–70.
Sanders, S. W., & Moore, J. G. (1992). Gastrointestinal Stratton, R. J., Stubbs, R. J., & Elia, M. (2003). Short-term
chronopharmacology: Physiology, pharmacology and thera- continuous enteral tube feeding schedules did not suppress
peutic implications. Pharmacology & Therapeutics, 54(1), appetite and food intake in healthy men in a placebo-
1–15. controlled trial. The Journal of Nutrition, 133(8), 2570–2576.
Saper, C. B., Scammell, T. E., & Lu, J. (2005). Hypothalamic Sweet, D. C., Levine, A. S., Billington, C. J., & Kotz, C. M.
regulation of sleep and circadian rhythms. Nature, 437 (1999). Feeding response to central orexins. Brain Research,
(7063), 1257–1263. 821(2), 535–538.
Scheer, F. A., Hilton, M. F., Mantzoros, C. S., & Shea, S. A. Szajewska, H., & Ruszczynski, M. (2010). Systematic review
(2009). Adverse metabolic and cardiovascular consequences demonstrating that breakfast consumption influences body
of circadian misalignment. Proceedings of the National weight outcomes in children and adolescents in Europe.
Academy of Sciences of the United States of America, 106 Critical Reviews in Food Science and Nutrition, 50(2),
(11), 4453–4458. 113–119.
Schlierf, G., & Dorow, E. (1973). Diurnal patterns of Tahara, Y., Otsuka, M., Fuse, Y., Hirao, A., & Shibata, S.
triglycerides, free fatty acids, blood sugar, and insulin during (2011). Refeeding after fasting elicits insulin-dependent reg-
carbohydrate-induction in man and their modification by ulation of Per2 and Rev-erbalpha with shifts in the liver
nocturnal suppression of lipolysis. The Journal of Clinical clock. Journal of Biological Rhythms, 26(3), 230–240.
Investigation, 52(3), 732–740. Terazono, H., Mutoh, T., Yamaguchi, S., Kobayashi, M.,
Sensi, S., Capani, F., & Caradonna, P. (1973). Diurnal varia- Akiyama, M., Udo, R., et al. (2003). Adrenergic regulation
tion in response to tolbutamide. Lancet, 1(7807), 830. of clock gene expression in mouse liver. Proceedings of the
Sim, L. A., McAlpine, D. E., Grothe, K. B., Himes, S. M., National Academy of Sciences of the United States of
Cockerill, R. G., & Clark, M. M. (2010). Identification and America, 100(11), 6795–6800.
treatment of eating disorders in the primary care setting. Thaela, M. J., Jensen, M. S., Cornelissen, G., Halberg, F.,
Mayo Clinic Proceedings, 85(8), 746–751. Noddegaard, F., Jakobsen, K., et al. (1998). Circadian and
Sladek, M., Rybova, M., Jindrakova, Z., Zemanova, Z., ultradian variation in pancreatic secretion of meal-fed pigs
Polidarova, L., Mrnka, L., et al. (2007). Insight into the cir- after weaning. Journal of Animal Science, 76(4), 1131–1139.
cadian clock within rat colonic epithelial cells. Gastroenterol- Torres-Farfan, C., Seron-Ferre, M., Dinet, V., & Korf, H. W.
ogy, 133(4), 1240–1249. (2006). Immunocytochemical demonstration of day/night
Spano, G. M., La, M. R., Pettirossi, G., Pulcinelli, F. M., changes of clock gene protein levels in the murine adrenal
Gazzaninga, P. P., & Cordova, C. (1993). Circadian gland: Differences between melatonin-proficient (C3H)
376

and melatonin-deficient (C57BL) mice. Journal of Pineal Wang, X. S., Armstrong, M. E., Cairns, B. J., Key, T. J., &
Research, 40(1), 64–70. Travis, R. C. (2011). Shift work and chronic disease: The
Turek, F. W., Joshu, C., Kohsaka, A., Lin, E., Ivanova, G., epidemiological evidence. Occupational Medicine (Oxford,
McDearmon, E., et al. (2005). Obesity and metabolic syndrome England), 61(2), 78–89.
in circadian Clock mutant mice. Science, 308(5724), 1043–1045. Welsh, D. K., Logothetis, D. E., Meister, M., & Reppert, S. M.
Valenzuela, F. J., Torres-Farfan, C., Richter, H. G., (1995). Individual neurons dissociated from rat
Mendez, N., Campino, C., Torrealba, F., et al. (2008). Clock suprachiasmatic nucleus express independently phased cir-
gene expression in adult primate suprachiasmatic nuclei and cadian firing rhythms. Neuron, 14(4), 697–706.
adrenal: Is the adrenal a peripheral clock responsive to mel- Whichelow, M. J., Sturge, R. A., Keen, H., Jarrett, R. J.,
atonin? Endocrinology, 149(4), 1454–1461. Stimmler, L., & Grainger, S. (1974). Diurnal variation in
Van Cauter, E., Polonsky, K. S., & Scheen, A. J. (1997). Roles response to intravenous glucose. British Medical Journal, 1
of circadian rhythmicity and sleep in human glucose regula- (5906), 488–491.
tion. Endocrine Reviews, 18(5), 716–738. Woods, S. C., & D’Alessio, D. A. (2008). Central control of
Van Cauter, E., Shapiro, E. T., Tillil, H., & Polonsky, K. S. body weight and appetite. The Journal of Clinical Endocri-
(1992). Circadian modulation of glucose and insulin nology and Metabolism, 93(11 Suppl. 1), S37–S50.
responses to meals: Relationship to cortisol rhythm. The Yamazaki, S., Numano, R., Abe, M., Hida, A., Takahashi, R.,
American Journal of Physiology, 262(4 Pt. 1), E467–E475. Ueda, M., et al. (2000). Resetting central and peripheral cir-
van der Spek, R., Kreier, F., Fliers, E., & Kalsbeek, A. (2012). cadian oscillators in transgenic rats. Science, 288(5466),
Circadian rhythms in white adipose tissue. Progress in Brain 682–685.
Research, 199, 183–201. Yang, S., Liu, A., Weidenhammer, A., Cooksey, R. C.,
Vella, K. R., Ramadoss, P., Lam, F. S., Harris, J. C., Ye, F. D., McClain, D., Kim, M. K., et al. (2009). The role of mPer2
Same, P. D., et al. (2011). NPY and MC4R signaling regulate clock gene in glucocorticoid and feeding rhythms. Endocri-
thyroid hormone levels during fasting through both central nology, 150(5), 2153–2160.
and peripheral pathways. Cell Metabolism, 14(6), 780–790. Zhang, E. E., Liu, Y., Dentin, R., Pongsawakul, P. Y.,
Verrillo, A., De Teresa, A., Martino, C., Di Chiara, G., Liu, A. C., Hirota, T., et al. (2010). Cryptochrome mediates
Pinto, M., Verrillo, L., et al. (1989). Differential roles of circadian regulation of cAMP signaling and hepatic gluco-
splanchnic and peripheral tissues in determining diurnal neogenesis. Nature Medicine, 16(10), 1152–1156.
fluctuation of glucose tolerance. The American Journal of Zhang, E. E., Liu, A. C., Hirota, T., Miraglia, L. J., Welch, G.,
Physiology, 257(4 Pt. 1), E459–E465. Pongsawakul, P. Y., et al. (2009). A genome-wide RNAi
Vrang, N., Larsen, P. J., & Mikkelsen, J. D. (1995). Direct projec- screen for modifiers of the circadian clock in human cells.
tion from the suprachiasmatic nucleus to hypophysiotrophic Cell, 139(1), 199–210.
corticotropin-releasing factor immunoreactive cells in the par- Zimmet, P. Z., Wall, J. R., Rome, R., Stimmler, L., &
aventricular nucleus of the hypothalamus demonstrated by Jarrett, R. J. (1974). Diurnal variation in glucose tolerance:
means of Phaseolus vulgaris-leucoagglutinin tract tracing. Associated changes in plasma insulin, growth hormone,
Brain Research, 684(1), 61–69. and non-esterified fatty acids. British Medical Journal, 1
Vujovic, N., Davidson, A. J., & Menaker, M. (2008). Sympa- (5906), 485–488.
thetic input modulates, but does not determine, phase of Zucker, I. (1971). Light-dark rhythms in rat eating and drink-
peripheral circadian oscillators. American Journal of Physi- ing behavior. Physiology & Behavior, 6(2), 115–126.
ology. Regulatory, Integrative and Comparative Physiology, Zylka, M. J., Shearman, L. P., Weaver, D. R., &
295(1), R355–R360. Reppert, S. M. (1998). Three period homologs in mammals:
Wallace, A., Chinn, D., & Rubin, G. (2003). Taking simva- Differential light responses in the suprachiasmatic circadian
statin in the morning compared with in the evening: clock and oscillating transcripts outside of brain. Neuron, 20
Randomised controlled trial. BMJ, 327(7418), 788. (6), 1103–1110.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 21

Managing neurobehavioral capability when social


expediency trumps biological imperatives

Andrea M. Spaeth{, Namni Goel{ and David F. Dinges{,*

{
Department of Psychology, School of Arts and Sciences, University of Pennsylvania, Philadelphia, PA, USA
{
Division of Sleep and Chronobiology, Department of Psychiatry, Perelman School of Medicine,
University of Pennsylvania, Philadelphia, PA, USA

Abstract: Sleep, which is evolutionarily conserved across species, is a biological imperative that cannot
be ignored or replaced. However, the percentage of habitually sleep-restricted adults has increased in
recent decades. Extended work hours and commutes, shift work schedules, and television viewing are
particularly potent social factors that influence sleep duration. Chronic partial sleep restriction, a
product of these social expediencies, leads to the accumulation of sleep debt over time and
consequently increases sleep propensity, decreases alertness, and impairs critical aspects of cognitive
functioning. Significant interindividual variability in the neurobehavioral responses to sleep restriction
exists—this variability is stable and phenotypic—suggesting a genetic basis. Identifying vulnerability to
sleep loss is essential as many adults cannot accurately judge their level of impairment in response
to sleep restriction. Indeed, the consequences of impaired performance and the lack of insight due to
sleep loss can be catastrophic. In order to cope with the effects of social expediencies on biological
imperatives, identification of biological (including genetic) and behavioral markers of sleep loss
vulnerability as well as development of technological approaches for fatigue management are critical.

Keywords: sleep deprivation; sleep duration; neurobehavioral functions; fatigue management; individual
differences; genetics; biomarkers.

Introduction

“Sleep persists in predators and prey, carnivores


and vegetarians, on the land and in the water, in
*Corresponding author. most mammals as they lie down relaxed, in rum-
Tel.: þ1-215-898-9665; Fax: þ1-215-573-6410 inants while they stand, in birds while they perch,
E-mail: dinges@mail.med.upenn.edu

http://dx.doi.org/10.1016/B978-0-444-59427-3.00021-6 377
378

and in dolphins which constantly swim . . . in sleep, if one exists, remains elusive (Rechtschaffen,
the smartest and in the dumbest of all mammalian 1998; Siegel, 2005; Tobler, 1995). Our lack of knowl-
species” (Rechtschaffen, 1998). Although an organ- edge regarding sleep’s fundamental function has
ism cannot procreate, gather food, or protect itself led a select few to assume that humans need
or offspring during sleep, sleep has been evolution- only a small “core” amount of sleep to function nor-
arily conserved from the sloth to the human. mally (Horne, 1985, 1988); however, a large body of
Indeed, some evidence suggests that sleep is vital research indicates that most humans who undergo
to life. A number of studies from 1894 to 1962 chronic partial sleep deprivation exhibit profound
reported lethal outcomes from total sleep depriva- neurobiological and physiological changes that can
tion in dogs, rabbits, and rats (Kleitman, 1963). impede performance and negatively impact health
However, these studies lacked proper controls for (Banks et al., 2010; Dement and Greenber, 1966;
the procedures implemented to keep the animals Dinges, 2004; Goel et al., 2009a; Leproult and Van
awake. This methodological failure was corrected Cauter, 2010).
in 1983 when Rechtschaffen and colleagues devel-
oped a novel disk-over-water methodology with
yoked control procedures (Rechtschaffen et al., Healthy human sleep
1983). Using this paradigm, mortality was more
likely in the sleep-deprived rats (death occurred Normal human sleep comprises two states—rapid
after 5–33 days) compared with yoked control rats eye movement (REM) and non-rapid eye move-
that experienced the same forced ambulation but ment (NREM)—that alternate cyclically during
were allowed to rest. Research using Drosophila a sleep episode. Using electroencephalography
also showed that prolonged sleep deprivation could (EEG), characteristics of these sleep states have
be deadly (Shaw et al., 2002). The results obtained been well defined (Rechtschaffen and Kales, 1968;
from these more recent studies as well as earlier Silber et al., 2007). NREM sleep shows synchro-
studies strongly suggest prolonged sleep depriva- nous cortical EEG (sleep spindles, K-complexes,
tion can be lethal and can cause death even more and slow waves) and is associated with low muscle
rapidly than food deprivation (Everson et al., 1989). tone and minimal psychological activity whereas
Sleeping, like eating, is a biological imperative REM sleep shows desynchronized EEG and is
that consumes approximately one-third of human associated with muscle atonia and dreaming
life. Just as missing a day’s worth of meals pro- (Kryger et al., 2011). In healthy humans, the timing,
duces strong feelings of hunger, being deprived of intensity, and duration of sleep are primarily
a night’s sleep leads to overwhelming feelings of regulated by two processes: homeostatic regulation
fatigue. Sleep resists being deprived; if exhausted and circadian timing (Borbély, 1982, 1998).
enough, humans can fall asleep under even dan-
gerous circumstances (Nansen, 1999). When sleep
is denied, there is an increase in the frequency Sleep homeostasis
and duration of sleep episodes and an elevation
in the intensity of sleep (Banks et al., 2010; Sleep homeostasis describes the drive for sleep
Borbély and Tobler, 1996; Brunner et al., 1993; that increases progressively during wakefulness
Goel et al., 2009a, 2010; Van Dongen et al., and decreases progressively during NREM sleep
2003). These properties of sleep and its ubiquitous (Borbély, 1994). Organisms strive to maintain
manifestation across species indicate that sleep sleep at a constant level; therefore, the propensity
must serve an adaptive function. Despite our under- for sleep is increased after prolonged wakefulness
standing that sleep plays a necessary and adaptive and decreased after prolonged sleep. Similarly,
evolutionary purpose, the universal function of sleep restriction leads to an increase in the
379

intensity and duration of subsequent sleep, dark cycle which coincides with a period of inactiv-
whereas excess sleep leads to decreased sleep inten- ity and sleepiness and decreases during the light
sity and duration (Borbély, 1982). Researchers have cycle which coincides with a period of activity
found that slow waves, low-frequency EEG, likely and wakefulness (Aschoff and Wever, 1981;
represent a measure of sleep intensity—extended Shanahan and Czeisler, 1991). This circadian pro-
waking leads to increases in slow wave energy cess interacts with but is independent from sleep
(SWE) during the subsequent recovery night and homeostasis (Åkerstedt and Froberg, 1978).
the extent of this SWE increase is a function of prior Scientific evidence shows that human sleep is
wake duration (Åkerstedt et al., 2009; Brunner naturally regulated by these two processes, with
et al., 1990). In addition to SWE, other biological sunrise and sunset providing the photic signals
markers of sleep homeostasis have been identified necessary to entrain the sleep–wake cycle (Wehr
including extracellular adenosine, central nitrous et al., 2001). However, with the introduction of
oxide levels, and salivary amylase; levels of these artificial light and other technologies, television
markers increase with prolonged sleep–wakefulness and alarm clocks have replaced these natural
and thus may reflect an increased sleep drive signals and thereby may not allow optimal sleep
(Kalinchuk et al., 2006; Porkka-Heiskanen and duration. Additionally, industries that operate
Kalinchuk, 2011; Scharf et al., 2008; Seugnet et al., 24 h a day create light and noise that can interfere
2006). The homeostatic process of sleep–wake regu- with sleep (Basner et al., 2011) and require
lation interacts with but is independent from circa- employees to work during hours usually devoted
dian control (Dijk et al., 1989). to sleep. Finally, with the increase in international
business and air travel, many adults are fre-
quently traveling across time zones which affects
Sleep as a circadian process the circadian timing system. Thus, many humans
are challenging basic biological pressures in order
The circadian process that controls sleep is to accommodate social norms and obligations.
described as the 24-h oscillatory variation in the
propensity for sleep. This 24-h period is the time
it takes the Earth to rotate about its own axis Human sleep duration
which generates daily environmental cycles of
ambient temperature and illumination. The alter- For most healthy adults, physiological sleep dura-
nation of light and darkness directly entrains an tion appears to range between 7.0 and 8.5 h; how-
organism’s circadian rhythms and thus influences ever, habitual sleep duration among adults is
its life patterns, creating species that are active determined by a variety of factors and shows con-
primarily during the light (diurnal), the dark (noc- siderable variance within and between individuals
turnal), twilight periods (crepuscular), or during (Van Dongen et al., 2005). Data from the
the light and the dark (cathemeral) (Pittendrigh, 2005–2008 National Health and Nutrition Exami-
1981). Environmental light is transmitted from nation Survey suggest that self-reported sleep
the retina to the suprachiasmatic nuclei in the ante- duration is distributed approximately normally
rior hypothalamus which then transmit this infor- and 2004–2007 National Health Interview
mation to, among others, the pineal gland via a Survey-Sample Adult data reveal that 7.8% of
multisynaptic pathway; the pineal gland secretes adults report sleeping less than 5 h per night,
melatonin, a hormone that regulates various 20.5% report sleeping 6 h per night, 30.8% report
biological functions (Czeisler, 1995; Lucas et al., sleeping 7 h per night, 32.5% report sleeping 8 h
1999; Ralph et al., 1990; Sadun et al., 1984; Watts, per night, and 8.5% of adults report sleeping
1991). In humans, melatonin increases during the more than 9 h per night (Krueger and Friedman,
380

2009). One limitation of large epidemiological indicates that adult sleep duration is significantly
studies is that sleep duration is examined using shorter during weekdays compared to weekends,
self-report which can be subject to inherent biases suggesting that adults attempt to recover sleep
and thereby inaccurate. debt by extending sleep when it is presumably
Indeed, when comparing self-reported sleep more convenient and when schedules are likely
duration to objectively measured sleep duration more flexible (see Fig. 1a; Basner et al., 2007).
using actigraphy (an unobtrusive measure of gross
motor activity which is analyzed to objectively iden-
tify sleep periods), subjectively reported sleep aver- Causes of chronic partial sleep restriction
aged 0.80 h longer than objectively measured sleep
(Bradshaw et al., 2007; Lauderdale et al., 2008). Paid work
Although wrist actigraphy provides an objective
measure of inactivity, it still overestimates poly- Compensated work time may be the most potent
somnography (PSG)-measured sleep duration, determinant of sleep duration (see Fig. 1b and c;
which is the gold standard for determining physio- Basner et al., 2007). There is a higher prevalence
logical sleep. One study showed this overestimation of short sleep duration among full-time employed
was approximately 18 min on average (Blackwell adults when compared to part-time workers,
et al., 2008). Therefore, although epidemiological students, retired individuals, homemakers, or un-
studies are valuable in that they include data based employed adults (Knutson et al., 2010; Luckhaupt
on a large number of subjects, the actual amount of et al., 2010), and longer work hours are associated
physiological sleep that adults receive is typically with shorter sleep duration (Hale, 2005;
less than what is self-reported in these studies. Nakashima et al., 2010; Virtanen et al., 2009).
There is considerable debate as to whether or Adults working more than 8 h/day have the same
not sleep duration has been decreasing among bedtime compared to adults who do not work
adults in recent decades and if so, whether this more than 8 h/day, but they wake up much earlier
reduction is resulting in higher rates of chronic (Basner and Dinges, 2009). Using a prospective
sleep restriction or sleep debt (Dinges, 2004; study design, Virtanen et al. (2009) found that
Horne, 2004). However, according to the Center working more than 55 h/week is a risk factor for
for Disease Control Morbidity and Mortality the development of shortened sleep and for diffi-
Weekly Report (2008), the percentage of adults culty falling asleep. Although the average number
who reported an average of less than or equal to of hours spent working has remained relatively
6 h of sleep within a 24-h period significantly stable during the past few decades, the prevalence
increased from 1985 to 2004 (in both females of individuals who work greater than 48 h/week
and males and among all age groups 18–75). has increased (Rones et al., 1997). The duration
Among employed American adults, two epidemi- of required working hours varies by occupation,
ological studies found that the prevalence of with managerial, professional, manufacturing,
being a short sleeper (either  6 h/day (Knutson and transportation industries typically involving
et al., 2010) or < 6 h/day (Luckhaupt et al., longer work hours; interestingly, the prevalence
2010)) has increased significantly in the past few of short sleep duration is also highest in these
decades. Studies have suggested that habitual occupations (Luckhaupt et al., 2010). By contrast,
short sleepers do not require less sleep than other the lowest prevalence of short sleep duration is
adults; rather, these individuals gradually accrue observed in the “agriculture, forestry, fishing,
sleep debt over time (Aeschbach et al., 1996; and hunting industry” (Luckhaupt et al., 2010).
Bradshaw et al., 2007; Klerman and Dijk, 2005). Although workers within this category (such as
Evidence from the American Time Use Survey farmers) often work long hours and begin work
381

(a) early in the morning, their work schedules are


11.0 Saturday more in sync with the natural light–dark cycle
Sunday
10.5 Monday
compared to other occupations, and thus their
Tuesday work hours presumably interfere less with sleep
10.0 Wednesday quality and duration.
Thursday
Sleep time (h)

Friday In contrast to the schedules of agriculture, for-


9.5
estry, fishing, or hunting industries, shift work
9.0 schedules often require work to occur during the
night when the circadian system is promoting
8.5 sleep and require sleep to occur during the day
when the circadian system is promoting wakeful-
8.0
ness. This work schedule creates a conflict between
7.5 the worker’s internal circadian time and his/her
15–24 25–34 35–44 45–54 55–64 65–74 75+ required sleep–wake schedule, leading to impaired
Age range (years) wakefulness and disturbed sleep (Åkerstedt, 2005;
(b) Kolla and Auger, 2011). According to the Bureau
150 of Labor Statistics (2005), approximately 15% of
Adjusted change in activity time (min)

Mon – Fri full-time wage and salary earners work shifts that
100
are not during the daytime. Night workers sleep
2–4 h less per day than day workers (Åkerstedt,
50
2003) and workers either permanently on a night
0
shift or on a rotation including the night shift are
significantly more likely to sleep less than 6 h and
–50 experience excessive sleepiness during situations
requiring a high degree of attention (Ohayon
–100 et al., 2010). Although some countermeasures,
such as strategic light exposure (Burgess et al.,
–150 2002), or the use of melatonin, stimulants, and other
<4.5 <5.5 <6.5 <7.5 <8.5 <9.5 <10.5 <11.5 ³11.5

(c)
150
Adjusted change in activity time (min)

Sat + Sun on Sunday followed by Saturday and was markedly shorter


100
during the week, gradually decreasing from Monday to
Friday for most age groups. (b and c) Average change in
50
weekday (b) (Mon–Fri) and weekend (c) (SatþSun) waking
activity time depending on sleep time category (N ¼ 23,325)
0 based on multiple linear regression models adjusting for age,
Work
gender, ethnicity, educational attainment, income, presence
–50 Work (>0 min) of partner, and presence of children. The 7.5 h to <8.5 h
Socializing, relaxing, leisure w/o TV
Travel sleep time category served as a reference. For work time,
–100 Personal care w/o sleep separate models were run for the whole group (Work) and
TV
Household activities those only who worked on the interview day (Work > 0 min).
–150 Education The largest reciprocal relationship to sleep was found for
<4.5 <5.5 <6.5 <7.5 <8.5 <9.5 <10.5 <11.5 ³11.5 work time, followed by commute/travel time. Short sleepers
Sleep time (h) spent more time socializing, relaxing, and engaging in
leisure activities, while both short and long sleepers watched
Fig. 1. (a) Average sleep duration with respect to age and day more television than the average sleeper. (Reprinted with
of the week. Average sleep time across age groups was longest permission from Basner et al., 2007.)
382

pharmacological agents related to sleep–wake reg- so portable that they can easily be brought into
ulation can promote wakefulness or improve sleep bed and used right before going to sleep. Among
among shift workers (Kolla and Auger, 2011), cur- adults, computer and mobile phone use in the bed-
rently there is no treatment to completely normal- room is associated with greater variability in
ize a shift worker’s sleep duration that can be sleep–wake schedules and poorer sleep habits
widely used in real-world situations (Åkerstedt (Brunborg et al., 2011). In addition to the time used
and Wright, 2009). engaging in these activities, exposure to light emit-
ted from these screens may also impede sleep. Blue
(short wavelength) light represents the most potent
Commute/travel portion of the light spectrum for suppressing mela-
tonin and thus promoting wakefulness (Brainard
According to the Census American Community et al., 2001; Lockley et al., 2003); moreover, blue
Survey Report (2011), American workers spend light from light-emitting diodes elicits a dose-
nearly an hour commuting to and from work each dependent suppression of melatonin (West et al.,
day and 33.5% of workers have commutes that 2011). Exposure to blue light immediately prior to
are greater than 30 min each way. Travel time bedtime (via flat screen televisions, smartphones,
(composed of traveling to work [commute], and tablet computers) may cause circadian phase
stores, schools, and social events) is negatively delays and disrupt sleep (Cajochen et al., 2006,
associated with sleep time (see Fig. 1b and c; 2011).
Basner et al., 2007). A majority of Americans
(55.4%) leave their homes between 06:00 and
08:30 in order to arrive at work on time (Census Chronotype
American Community Survey Report, 2011),
and they do not arrive home until late in the Although humans are diurnal, some individuals
evening in part due to traffic; the result is an prefer activity in the morning (larks) whereas
increase in the length of the workday and a others prefer activity in the evening (owls)—this
decrease in time for other activities such as soci- preference influences the timing of sleep–wake
alizing, relaxing, and sleeping. cycles. Morning-type and evening-type individuals
differ endogenously in the circadian phase of their
biological clocks (Baehr et al., 2000; Kerkhof and
Leisure VanDongen, 1996), which is partially determined
by genetic polymorphisms (Goel, 2011). In addi-
Socializing, relaxing, and engaging in other leisure tion to genetic factors, age, and gender also
activities are also negatively related to sleep dura- influence morningness–eveningness (Roenneberg
tion (see Fig. 1b and c; Basner et al., 2007). For et al., 2007). The Earth’s light/dark cycle and the
many adults, this time usually involves sitting in work schedules of industrial societies complement
front of some type of screen, such as a television, individuals who function best in the morning
computer, or smartphone. For example, watching rather than in the evening. Owls experience
television was the most common activity during heightened alertness in the late evening and typi-
the 2-h period before bedtime, suggesting that cally stay awake longer and have a delayed bed-
nighttime television viewing and sleep onset are time compared to larks; however, due to typical
tightly correlated (Basner and Dinges, 2009; Basner work schedules owls often use an alarm clock to
et al., 2007). Adults are now able to access e-mail, wake up early in the morning, which produces
engage in social networking and online gaming, chronic sleep restriction, extended wakefulness,
and use the web 24 h a day; devices have become and accumulated sleep debt during the work week
383

(Korczak et al., 2008; Roenneberg et al., 2003; performance. The homeostatic sleep-dependent
Roepke and Duffy, 2010; Taillard et al., 2003). process (process S) balances sleep propensity by
Social jetlag, a term coined by Roenneberg and keeping track of recent sleep history (Daan et al.,
colleagues (Wittmann et al., 2006), describes the 1984). As hours of wakefulness increase, homeo-
misalignment of biological and social time caused static drive increases the propensity for sleep. The
by social activities such as work, commuting, and endogenous circadian process (process C) tracks
television viewing. For many adults, social jetlag changes in light exposure (as well as other zeitgebers
and/or a preference for eveningness can lead to or synchronizers) and entrains sleep propensity to
the accumulation of sleep debt during the work the light–dark cycle; when sleep propensity is
week due to repeated episodes of shortened increased (during the night) waking performance is
sleep. This repeated and chronic partial sleep often degraded (Czeisler et al., 1999). Sleep inertia,
deprivation has been shown to have serious neu- a transient process that occurs immediately after
robehavioral and physiological consequences awakening, causes deficits in waking performance;
unlikely to be reversed by increasing sleep time its effects depend on prior sleep history, circadian
on weekends and days off (Banks et al., 2010; phase, and the depth of sleep at the time of awaken-
Goel et al., 2009a; Van Dongen et al., 2003). ing (Dinges, 1990). Finally, a second homeostatic
process (process U) influences waking performance
by monitoring sleep–wake on a longer time scale
Consequences of chronic partial sleep restriction (nightly sleep duration) and interacting with the
other processes (McCauley et al., 2009). Process U
According to the National Transportation Safety builds up over several days of prolonged wakeful-
Board (1989, 1995), 30–40% of all U.S. truck ness, when sleep debt is accruing and the need for
accidents are fatigue related and several sleep is increasing, and dissipates during sleep. The
catastrophes including the Exxon Valdez accident, status of process U, similar to a sleep reservoir being
Chernobyl and Three Mile Island nuclear plant full or empty, codetermines the rate by which the
meltdowns, and Space Shuttle Challenger tragedy original homeostatic process (process S) increases
were partially due to human error resulting from sleep propensity and thus decreases waking perfor-
sleepiness and fatigue (Mitler et al., 1988). Sleepi- mance (McCauley et al., 2009).
ness due to long work hours has also been All of these processes interact to form a
implicated in medical errors: interns, surgeons, dynamic relationship between sleep propensity
physicians, and residents make significantly more and waking performance (Raslear et al., 2011;
errors, including those that are injurious or fatal, Van Dongen and Dinges, 2005). During a 24-h
after working for an extended period of time period, the interaction of these processes leads
and/or after chronic and acute sleep loss (Czeisler, to a period in the early morning when waking
2009; Lockley et al., 2004). As such, gaining a performance is particularly vulnerable to sleep
greater understanding of the relationship between restriction; neurobehavioral deficits are largest at
sleep and waking performance is an important 08:00 and become progressively smaller during
focus of ongoing research. the day, especially between 16:00 and 20:00
(Mollicone et al., 2010). The incidence of single-
vehicle automobile accidents, human error, and
Regulation of human performance accidents related to work performance and sev-
eral industrial/engineering disasters also reveal
Similar to the regulation of sleep duration and inten- early morning hours as particularly vulnerable to
sity, distinct sleep/wake-related physiological pro- sleep restriction (Mitler et al., 1988). However,
cesses regulate alertness and neurobehavioral when an individual lacks a sufficient amount of
384

sleep, either as a result of acute total sleep depri- 2003; Carskadon and Dement, 1981; Devoto
vation or chronic partial sleep restriction, homeo- et al., 1999; Guilleminault et al., 2003; Rupp
static pressure can increase to the point whereby et al., 2009). Spending 10 h in bed daily for 1
waking cognitive functions will be degraded even week prior to experiencing a week of sleep
during periods of time when the circadian drive restriction may attenuate the effects of sleep
for wakefulness peaks (Doran et al., 2001). restriction on sleep propensity as measured by
MWT (Rupp et al., 2009). By contrast, after
experiencing partial sleep restriction, one recov-
Sleep propensity ery night of up to 10 h time in bed (TIB) is insuf-
ficient to return sleep propensity back to presleep
Individuals who report sleeping less than 7.5 h/ restriction levels (see Fig. 2b; Banks et al., 2010).
day on average are more likely to exhibit a MWT sleep latencies appear normalized after two
stronger sleep propensity (the tendency to fall 8 h TIB nights of recovery sleep (Rupp et al.,
asleep)—they report unintentionally falling 2009); however, due to the small sample size used
asleep during the day and nodding off or falling in this study, replication of this finding is needed.
asleep while driving (McKnight-Eily et al., 2011;
Punjabi et al., 2003). In the laboratory, sleep
propensity—operationalized as the speed of falling Behavioral alertness: Microsleeps and wake-state
asleep in both sleep-conducive and non-conducive instability
conditions—is among the most well-validated
measures of sleepiness (Roehrs et al., 2005). The The increased propensity for sleep due to
effects of chronic sleep restriction on daytime restricted sleep time can lead to the occurrence
physiological sleep propensity have been evaluated of “microsleeps,” very brief sleep episodes that
using the Multiple Sleep Latency Test (MSLT) intrude into wakefulness despite an individual’s
(Carskadon and Dement, 1981) and the Mainte- best effort to stay awake (Åkerstedt et al., 1987;
nance of Wakefulness Test (MWT) (Mitler et al., Bjerner, 1949; Torsvall and Åkerstedt, 1987).
1982). During the MSLT, the subject is instructed Wake-state instability refers to the moment-to-
to close his/her eyes and try to fall asleep while moment shifts in the neurobiological systems
lying in a supine position, whereas during the mediating the motivated desire to sustain waking
MWT, the subject is instructed to look straight alertness and those mediating the involuntary
ahead and try to stay awake while seated upright. homeostatic drive to fall asleep (Doran et al.,
In both tests, PSG recordings are made (including 2001; Saper et al., 2005). This type of interaction
EEG, electrooculogram and electromyogram) and between drives results in unpredictable behavior,
the time taken to fall asleep is a measure of sleep including increased variability in cognitive perfor-
propensity. mance and lapsing (i.e., brief periods of half a sec-
Whether attempting to fall asleep or resist ond to many seconds of no response) (Doran
sleep, the latency from waking to sleeping is et al., 2001). Although individuals may not realize
reduced by chronic partial sleep restriction (see they are experiencing microsleeps and perfor-
Fig. 2a; Banks et al., 2010; Belenky et al., 2003; mance decrements, over time this instability can
Carskadon and Dement, 1981; Dinges et al., progress into full blown sleep attacks—when
1997; Guilleminault et al., 2003; Rupp et al., individuals will not spontaneously wake without
2009). As the number of days of sleep restriction additional stimulation (Dinges and Kribbs, 1991;
increases or the amount of sleep available during Harrison and Horne, 1996; Kleitman, 1963).
each night of sleep restriction decreases, the Sleep restriction can increase the incidence of
latency to fall asleep decreases (Belenky et al., microsleeps and produce decreased behavioral
385

(a) (b)
0 0

5
MWT (min)

10

MWT (min)
10

20 15

20
30
25

(c) (d)
12 25

10 20
8

PVT lapses
PVT lapses

15
6
10
4

2 5

0 0
B2 SR1 SR2 SR3 SR4 SR5 0h 2h 4h 6h 8h 10 h
Protocol night Recovery night TIB

Fig. 2. (a and c) Daily means (SEM) of (a) MWT sleep onset latency and (c) PVT performance in the sleep restriction group
(n ¼ 142, 4 h TIB for five nights [SR1–SR5], solid line), and the control group (n ¼ 17, 10 h TIB on all nights, dashed line). All
subjects had 10 h TIB (22:00–08:00) on baseline day 2 (B2). Sleep restriction on SR1 to SR5 was from 04:00 to 08:00. Data are
plotted to show deficits in neurobehavioral functions increasing (upward) on the ordinate. Relative to the control condition,
sleep restriction degraded neurobehavioral function over days (decreased MWT sleep onset latency and increased PVT lapses
[>500 ms reaction times]). (b and d) Daily means (SEM) of (b) MWT sleep onset latency and (d) PVT performance following
the acute recovery (REC) night as a function of increasing TIB dose (0 –10 h, solid line), controlling for covariates (i.e., baseline,
cumulative deficits during sleep restriction, age, and sex). Sleep-restricted subjects (n ¼ 142) were randomized to either 0 h TIB
(n ¼ 13), 2 h TIB (n ¼ 27), 4 h TIB (n ¼ 29), 6 h TIB (n ¼ 25), 8 h TIB (n ¼ 21), or 10 h TIB (n ¼ 27). For comparison, horizontal
dotted lines show baseline night (B2, 10 h TIB) values, and horizontal dashed lines show control group (n ¼ 17) means on day
8 (10 h TIB), which is the study day equivalent to REC. These neurobehavioral outcomes showed improvement as recovery sleep
doses increased (MWT sleep onset latencies increased and PVT lapses decreased); however, even at the highest sleep recovery
dose, functioning did not return to baseline or control condition levels. (Reprinted with permission from Banks et al., 2010.)

alertness, even during goal-directed behaviors 2010; Trew et al., 2011). Unfortunately, workers in
(e.g., motor vehicle operation). The rapid intrusion these professions are often subjected to chronic
of sleep into wakefulness and consequent wake- sleep restriction due to shift work schedules and
state instability can have profound adverse con- long work hours and are thus particularly vulnera-
sequences, especially if experienced by those in ble to microsleeps and wake-state instability.
safety-sensitive occupations, such as police officers, Behavioral alertness, measured using sustained
firefighters, health care providers, and motor vehi- attention tasks in the laboratory, has been shown
cle operators (Barger et al., 2009; Lombardi et al., to be sensitive to sleep restriction. According to
386

the state instability hypothesis, sleep drive primarily focused on short-term sleep restriction
escalates instability in attention which creates and have found that driving performance on
increasingly variable neurobehavioral perfor- simulators decreases (resulting in more crashes)
mance (periods of accurate responding are inter- after short sleep duration (4–6 h TIB) (De Valck
rupted by errors of omission [lapses] comingled and Cluydts, 2003; Lenne et al., 2004; Macchi
with errors of commission) (Doran et al., 2001). et al., 2002; Otmani et al., 2005; Vakulin et al.,
Studies support this hypothesis; cognitive perfor- 2007). Similarly, the less sleep obtained during
mance variability is influenced by prior wakeful- chronic partial sleep restriction, the more driving
ness, circadian phase, and the amount of time performance (increase in number of driving simu-
spent on the task (Doran et al., 2001; Graw lator accidents) is impaired (Russo et al., 2003).
et al., 2004; Mollicone et al., 2010; Zhou et al.,
2011b). When sleep is chronically restricted, sleep
debt accumulates across each night leading to Cognitive performance
greater impairment over time (Banks et al.,
2010; Belenky et al., 2003; Dinges et al., 1997; In addition to its negative effects on alertness, sleep
Goel et al., 2009a; Rupp et al., 2009; Van Dongen deprivation degrades aspects of cognitive perfor-
et al., 2003; Wu et al., 2010). Thus, as the need for mance (Harrison and Horne, 2000; Kleitman,
sleep increases, the brain’s ability to maintain 1963; McCoy and Strecker, 2011). There are
alertness becomes progressively more over- hundreds of published studies on the effects of total
whelmed by the activation of sleep processes sleep deprivation on cognitive performance (Alhola
which leads to microsleeps and consequent neuro- and Polo-Kantola, 2007; Goel et al., 2009b; Lim
behavioral instability (Doran et al., 2001). Even and Dinges, 2010), but far fewer studies on the
when highly motivated, an individual’s attempt cognitive-impairing effects of chronic partial sleep
to compensate for excessive sleepiness by restriction.
engaging in various behaviors ultimately fails to Sleep deprivation induces a wide range of
prevent intrusions of microsleeps and impaired effects on cognitive functions (Killgore, 2010;
neurobehavioral performance (Horne and Pettitt, Poe et al., 2010); however, cognitive tasks vary
1985; Hsieh et al., 2010). considerably in their sensitivity to sleep loss.
In general, regardless of the task, cognitive per-
formance becomes progressively worse when time
Driving performance on task is extended; this is the classic “fatigue”
effect that is exacerbated by sleep loss (Bjerner,
Driving a vehicle requires sustained attention and 1949; Wilkinson, 1969). However, performance
is sensitive to the reduced alertness associated on brief cognitive tasks that measure speed of
with sleepiness. Every year, thousands of automo- cognitive “throughput,” working memory, and
bile crashes, injuries, and fatalities are due to other aspects of attention are also sensitive to sleep
drivers falling asleep (Strohl et al., 1997). Adults deprivation (Dinges, 1992). Two confounding
who are employed, work more than 60 h/week, factors that can obscure the effects of sleep loss on
work irregular hours, or work at night, as well as many cognitive tasks are intersubject variability
those who are sleep deprived or chronically sleep and intrasubject variability (Dorrian et al., 2005).
restricted are more likely to have a car accident The intersubject confound describes differences in
(Abe et al., 2010; Åkerstedt et al., 2005; Brown, aptitude: one individual’s poorest performance
1994; Gander et al., 2005; Horne and Reyner, during sleep deprivation may be superior to the
1999; Philip et al., 2003, 2005; Scott et al., 2007; best performance of a non-sleep-deprived individ-
Stutts et al., 2003). In the laboratory, studies have ual (Goel et al., 2009b). The intrasubject confound
387

describes a learning effect: a person may be cogni- deprivation (Van Dongen et al., 2003). This
tively diminished by sleep loss, but continue to finding has been replicated several times (Axelsson
improve on a repeated task due to the effects of et al., 2008; Banks et al., 2010; Bliese et al., 2006;
learning (Goel et al., 2009b). The nature of the Cote et al., 2008; Dinges et al., 1997; Fafrowicz
dependent variables selected for examining the et al., 2010; Goel et al., 2009a; Rupp et al., 2009;
cognitive effects of sleep deprivation can also be Wu et al., 2010).
problematic. Sleep deprivation increases variability Research has also shown that the recovery rate
within subjects (i.e., state instability) and between from chronic sleep restriction may be slower than
subjects (i.e., differential vulnerability to the effects from acute total sleep loss (Banks et al., 2010;
of sleep deprivation); therefore, the effects of sleep Rupp et al., 2009). In a large-scale experimental
loss on cognitive measures may be missed due to study examining dose–response effects of one
insensitive metrics or data analyses (Olofsen et al., night of recovery sleep after five nights of sleep
2004; Van Dongen et al., 2004a; Whitney and restriction on PVT performance, recovery to
Hinson, 2010). To provide an accurate and useful either a subject’s own baseline values or values
measure of performance during sleep loss as well recorded for the sleep-satiated control group
as the dynamic expression of waking neuro- was not achieved at any of the doses examined,
behavioral integrity as it changes over time, cogni- including the highest dose of 10 h TIB (see Fig. 2d;
tive assessments must be valid and reliable Banks et al., 2010). Due to circadian limitations
reflections of fundamental waking functions altered on sleep duration, it is unlikely that recovery from
by sleep deprivation (Goel et al., 2009b). As such, 5 days of chronic sleep restriction can occur in
measures of attention, vigilance, and declarative one night of more than 10 h TIB; therefore, resid-
memory are often used, with reaction time as the ual attentional deficits still present after one night
dependent variable. of recovery may potentiate the effects of a
The Psychomotor Vigilance Test (PVT), a mea- subsequent sleep restriction period (Banks et al.,
sure of sustained attention, is free of aptitude and 2010). Thus, attempting to recover lost sleep from
learning effects and is widely used in sleep studies a work week by extending sleep on a weekend
due to its sensitivity to sleep loss, sleep pathology, night is likely insufficient for recuperating impaired
and functioning at an adverse circadian phase alertness and sustained attention. Other studies
(Balkin et al., 2004; Dorrian et al., 2005; Lim examining recovery after chronic sleep restriction
and Dinges, 2008). The primary outcome measures have found that PVT performance remains sub-
of the PVT are reaction time and errors, includ- standard even after 3, 5, or 7 recovery nights of
ing omissions (lapses) and commissions (false 8 h TIB (Axelsson et al., 2008; Belenky et al.,
responses). In 2003, two large-scale experimental 2003; Rupp et al., 2009). However, similar to results
studies found that when sleep is reduced to 3, 4, for sleep propensity, prophylactic napping prior to
5, or 6 h TIB for several nights, sustained attention sleep deprivation significantly improved reaction
performance (PVT) as well as working memory time performance and spending 10 h TIB daily for
and cognitive throughput (Serial Addition/Sub- 1 week prior to experiencing chronic partial sleep
traction Task, Digit Symbol Substitution Task) restriction attenuated the decrement in PVT per-
decreases in a dose-dependent manner (see Fig. 2c; formance during sleep loss and facilitated improve-
Belenky et al., 2003; Van Dongen et al., 2003). ment of PVT performance during recovery (Dinges
As sleep debt accumulates across days, perfor- et al., 1987; Rupp et al., 2009). Thus, chronic sleep
mance becomes progressively worse over time; restriction induces slow changes in neural processes
14 days of chronic sleep restriction (4 or 6 h TIB mediating alertness and attention that cause perfor-
each night) produced comparable cognitive deficits mance to become progressively worse over time,
to those produced by 24–48 h of total sleep producing the accumulation of severe performance
388

decrements. Similarly, the slow recovery rate from and Harrison, 2001). In experimental studies,
chronic sleep restriction suggests that this type of individuals who are sleep restricted exhibit deficits
sleep loss may induce long-term neuromodulatory on the Wisconsin Card Sorting Task (Herscovitch
changes in brain physiology (Banks et al., 2010; et al., 1980) and Stroop test (Stenuit and Kerkhofs,
Rupp et al., 2009). 2008) but do not exhibit deficits on the Hayling,
Researchers have proposed that attention and Brixton, COWAT, or Tower of London Tasks
working memory, considered basic cognitive (Goel et al., 2009a). Although subjects are able
actions, are essential to virtually all other cogni- to accurately perform many cognitive assessments,
tive processes; therefore, sufficient attention and compensatory mechanisms become engaged in
memory performance are necessary for optimal response to chronic sleep restriction which
functioning of neural circuits that mediate higher impedes the ability to think flexibly and increases
cognitive functions, such as executive function the use of automatic processing, impulsivity, and
(Balkin et al., 2008). Interestingly, there is a posi- rigid rule adherence to complete tasks (Stenuit
tive relationship between the level of sustained and Kerkhofs, 2008; Swann et al., 2006).
attention required to perform a task and the Plessow and colleagues studied new parents, a
degree to which task performance is impaired by population of healthy adults who are often sub-
sleep loss (Jennings et al., 2003). Although there jected to chronic sleep restriction for several months
is evidence that sleep deprivation adversely due to nightly feedings and duties related to caring
affects prefrontal cortex-related executive atten- for a new infant. Parents were administered an
tion and working memory abilities, these cogni- explicit-cueing version of the task-switching para-
tive effects are often not as prominent or easy to digm to assess their ability to flexibly adapt to
measure as those involving basic processes such changing environmental demands. Sleep-deprived
as cognitive and psychomotor speed (Goel et al., new parents exhibited significantly slower reaction
2009b). Therefore, more complex cognitive tasks times during task switches compared to new parents
involving higher cognitive functions have been reporting adequate sleep; correlational analyses
regarded as insensitive to sleep deprivation by showed that greater self-reported sleep debt was
some researchers (Harrison and Horne, 2000). associated with greater impairment on the task
Sleep restriction affects mental flexibility, (Plessow et al., 2010). These results indicate that
attention shifting, and the inhibition of automatic chronic sleep restriction can impair an individual’s
responses, all reflective of executive function ability to implement task goals in order to switch
(McCoy and Strecker, 2011). Executive function tasks in a fast-changing environment—factors
can be defined as “the ability to plan and coordi- important for doctors, pilots, and individuals in the
nate a willful action in the face of alternatives, to armed forces, who constantly rely on fast and flexi-
monitor and update action as necessary and sup- ble goal shifting and are likely to experience chronic
press distracting material by focusing attention on sleep restriction.
the task at hand” (Jones and Harrison, 2001).
Many tasks believed to engage different aspects
of executive function have been used in sleep dep- Interindividual differences in response to sleep
rivation studies. Examples of such tasks include restriction
the Brixton Spatial Anticipation Test, Controlled
Oral Word Association Test (COWAT), Hayling As stated previously, average sleep duration is
Sentence Completion Task, Torrance Tests of Cre- normally distributed and widely differs depending
ative Thinking, Tower of London, Thurstone’s on the individual (Van Dongen et al., 2005).
Verbal Learning Task, and Wisconsin Card In addition, although humans are diurnal
Sorting Task (Harrison and Horne, 2000; Jones mammals, there are individual differences in the
389

timing of their behaviors; some prefer to be active Individuals are highly vulnerable, somewhat vul-
in the morning whereas others prefer to be active nerable, or highly resistant to the neurobehavioral
in the late evening (Kleitman, 1963; Roenneberg effects of sleep restriction (Goel and Dinges,
et al., 2003). Similarly, studies of total sleep depri- 2011). It remains unknown whether the same
vation as well as chronic sleep restriction have individuals are vulnerable to the adverse effects of
reported substantial interindividual differences in both acute total sleep deprivation and chronic sleep
the magnitude of sensitivity/resilience to the restriction.
effects of sleep loss on neurobehavioral functions A few studies have found that these interindivid-
(see Fig. 3a; Goel et al., 2009a, 2010; Van Dongen ual differences may be task dependent; one
et al., 2004b). Prior sleep history and endogenous individual’s sustained attention may be most
circadian rhythms likely play an important role in affected by sleep loss whereas another individual’s
an individual’s response to sleep loss; however, lab- working memory may be most affected (Frey
oratory studies that carefully control for these et al., 2004; Van Dongen et al., 2004b). In addition,
factors still find marked individual differences in interindividual sensitivity to sleep loss is trait like;
neurobehavioral responses to sleep loss (Leproult when neurobehavioral performance was assessed
et al., 2003; Van Dongen et al., 2004b). In the stud- repeatedly (during separate laboratory visits), the
ies described earlier, when sleep duration was same individuals remained particularly vulnerable
limited to less than 7 h per night for several consec- or resistant to the effects of sleep loss even when
utive nights, performance on sustained attention, previously sleep satiated or sleep restricted (Van
working memory, and cognitive throughput was Dongen et al., 2004b). Finally, these stable interin-
significantly impaired; however, not all subjects dividual differences account for a substantial
were affected to the same degree (Axelsson et al., proportion of variance in cognitive performance
2008; Belenky et al., 2003; Bliese et al., 2006; Goel decrements induced by sleep loss. Demographic
et al., 2009a, 2010; Van Dongen et al., 2003). factors (age, sex, IQ), baseline functioning, various

(a) (b)
13 7
Mean KSS sleepiness score

11 6
Mean PVT lapses

9
5
7
4
5

3 3

1 2
B SR1 SR2 SR3 SR4 SR5 B SR1 SR2 SR3 SR4 SR5
Day Day

Fig. 3. Neurobehavioral performance at baseline and during chronic partial sleep deprivation for the PER3 genotype groups. Mean
(SEM) (a) PVT lapses (> 500 ms reaction times) per trial and (b) Karolinska Sleepiness Scale (KSS) scores at baseline (b) for
each partial sleep restriction night (SR1–SR5) for PER34/4 (open circles), PER34/5 (gray triangles), and PER35/5 (closed circles)
subjects. All genotypes demonstrated large but equivalent cumulative decreases in cognitive performance (PVT) and increases in
sleepiness (KSS) across sleep restriction nights. As illustrated by increasing standard error bars in (a), interindividual differences
in PVT lapses increased across sleep restriction nights in the entire study group (from Goel et al., 2009a).
390

aspects of habitual sleep timing, and circadian in cognitive performance and increases in sleepiness
chronotype have not accounted for the robust and in response to chronic sleep restriction (Goel et al.,
stable differences in neurobehavioral responses 2010). Additional investigations are underway to
to sleep restriction (Doran et al., 2001; Goel and identify other genetic polymorphisms that may medi-
Dinges, 2011; Van Dongen and Dinges, 2003; Van ate an individual’s neurobehavioral vulnerability to
Dongen et al., 2004b). sleep restriction and acute total sleep deprivation
Due to the trait like, or phenotypic, nature of these (Goel, 2011).
interindividual differences, research has focused on Identifying correlates of interindividual differ-
understanding which genetic polymorphisms may ences in neurobehavioral vulnerability to chronic
underlie vulnerability to sleep loss. Several genetic sleep restriction, such as biomarkers (including
mechanisms have been examined in relation to the genetic polymorphisms as well as adenosine and
effects of chronic sleep restriction including the salivary amylase levels) and behavioral predictors,
PERIOD3 (PER3) variable number tandem repeat will provide a viable means to determine those
(VNTR) polymorphism and the DQB1*0602 allele. individuals in the general population who need lon-
Although the VNTR polymorphism of the circadian ger habitual sleep durations and/or who require
gene PER3 (PER34/4, PER34/5, and PER35/5) has effective interventions and countermeasures for
been shown to influence executive function deficits unavoidable sleep loss (Goel and Dinges, 2011). This
in response to total sleep loss (Groeger et al., 2008; is a particularly important area of research since
Vandewalle et al., 2009; Viola et al., 2007), it does adults are not adept at accurately judging how
not appear to influence neurobehavioral effects of affected they are by chronic sleep restriction (see
chronic sleep restriction. A large-scale experimental Figs. 3 and 4; Banks and Dinges, 2007; Van Dongen
study examining individuals with different PER3 et al., 2003; Zhou et al., 2011b). Subjective sleepiness
genotypes found that PER34/4, PER34/5, and is measured using two widely used and well-
PER35/5 genotypes exhibited similar baseline per- validated questionnaires: the Stanford Sleepiness
formance on neurobehavioral assessments and Scale (SSS) which requires subjects to rate how
demonstrated large but equivalent cumulative sleepy they feel on a scale of 1–7 (Hoddes et al.,
increases in sleepiness and sleep propensity and 1973) and the Karolinska Sleepiness Scale (KSS)
cumulative decreases in cognitive performance and which requires subjects to rate how sleepy they feel
physiologic alertness across five nights of sleep on a scale of 1–9 (Åkerstedt and Gillberg, 1990).
restricted to 4 h per night (see Fig. 3; Goel et al., Although subjects are able to detect the rapid and
2009a). Daily intersubject variability also increased severe changes in levels of alertness and sleepiness
across sleep restriction days similarly in all genotypes in response to total sleep deprivation, subjects are
(Goel et al., 2009a). Although PER3 genotypes did much less sensitive in detecting the smaller changes
not differ at baseline in habitual sleep, physiological in levels of alertness and sleepiness that accumulate
sleep structure, or sleepiness, during sleep restriction, during each day with sleep restriction (see Figs. 3
PER35/5 subjects had elevated sleep homeostatic and 4; Banks and Dinges, 2007; Goel et al., 2009a;
pressure, measured physiologically by EEG SWE Van Dongen et al., 2003; Zhou et al., 2011a). After
during NREM compared with PER34/4 subjects a week (or two) of sleep restriction, subjects show
(Goel et al., 2009a). Goel and colleagues also exam- marked cognitive impairment and severe decreases
ined the human leukocyte antigen DQB1*0602 allele in alertness but rate themselves as only moderately
which is closely associated with narcolepsy. Subjects sleepy (Belenky et al., 2003; Van Dongen et al.,
who were either positive or negative for the 2003). Consequently, a subject’s rating of subjective
DQB1*0602 allele differed in levels of sleep homeo- sleepiness does not accurately parallel the
static pressure, baseline sleepiness, and sleep physi- continuing accumulation of cognitive performance
ology but showed comparable cumulative decreases deficits associated with sleep loss. This finding
391

Objective Subjective

PVT performance lapses SSS rating

15.0 4h TIB 3.0

10.0 2.0
4h TIB
6h TIB
6h TIB
5.0 1.0

8h TIB
0.0 0.0 8h TIB

BL 2 4 6 8 10 12 14 BL 2 4 6 8 10 12 14
Days of sleep restriction Days of sleep restriction

Fig. 4. Restriction of nocturnal sleep in healthy adults resulted in near-linear increases in PVT lapses (>500 ms reaction times) of
attention across 14 days (coefficients of change near 1.0), but subjective ratings of sleepiness, measured using the Stanford
Sleepiness Scale (SSS), showed a nonlinear coefficient below 0.5 for change over days. As objective performance continued to
decline near-linearly, there were only minor further increases in the subjective ratings of sleepiness. By the end of the 14 days of
sleep restriction, when PVT performance was at its worst levels, subjects in the 4 and 6-h sleep period conditions reported
feeling only slightly sleepy. Therefore, unlike performance measures, sleepiness ratings appeared to show adaptation to sleep
restriction (N ¼ 35 subjects; 8 h condition n ¼ 9, 6 h condition n ¼ 13, and 4 h condition n ¼ 13). (Reprinted with permission from
Van Dongen et al., 2003.)

suggests that people who are chronically sleep regulatory organization imposes limits to daily/
restricted underestimate the impact of sleep restric- weekly shift duration as well as the number of
tion and may overestimate their levels of alertness consecutive shifts or rest hours (Czeisler, 2009;
and ability to perform various cognitive tasks Gander et al., 2011). Regulatory organizations
(Banks and Dinges, 2007; Czeisler, 2009; Van may also screen potential employers for sleep dis-
Dongen et al., 2003; Zhou et al., 2011a). orders when occupational duties pose safety risks
(Czeisler, 2009; Gander et al., 2011). Although
these policies can improve workplace safety and
Management of sleep restriction-induced productivity, many regulations are not strictly
neurobehavioral impairment enforced or adhered to and it is nearly impossible
for an employer to ensure that workers are actu-
Due to the high prevalence of chronic partial sleep ally obtaining an appropriate amount of sleep each
restriction in industrialized societies, the severe night (Czeisler, 2009; Gander et al., 2011).
consequences of sleep loss on alertness and cogni- Balkin et al. (2011) summarized various tech-
tive function, and our inability to accurately moni- nological approaches to fatigue management that
tor our own sensitivity to chronic sleep loss, accurately predict and/or objectively monitor
objective actions need to be employed to manage sleepiness and fatigue effects in real time and that
sleep debt. Currently, fatigue management within allow for optimally timed administration of inter-
transportation, health professions, and other occu- ventions or countermeasures (such as the use of
pational settings are primarily based on policy; a properly dosed pharmaceuticals or napping).
392

When discussing the ideal monitoring system, the sleep loss. These effects can have serious adverse
authors state that technological approaches consequences including fatal medical errors, cata-
should be valid, reliable, sensitive, specific, and gen- strophic oil spills, and vehicle crashes. Finding
eralizable. In order to adhere to these criteria, the biological and behavioral markers of sleep loss
approach would have (a) the ability to predict vulnerability and developing technological
fatigue, based on the factors that produce it (sleep approaches to fatigue management are two
history, hours of wakefulness, circadian rhythms, domains of research essential for offsetting the
time of day, and biomarkers or behavioral correlates negative effects of societal factors on the
of sensitivity to sleep loss); (b) the ability to measure biological need for sleep.
and monitor fatigue/performance online in the oper-
ational environment in order to detect downward
trends in alertness or performance before such Acknowledgments
trends reach the threshold at which operational
performance is actually impacted; and (c) the ability Supported by NIH NR004281; the National Space
to effectively intervene when potential deficits Biomedical Research Institute through NASA
are identified or anticipated, with interventions NCC 9-58; and the Department of the Navy, Office
calibrated so as to restore and sustain alertness/per- of Naval Research (Award No. N00014-11-1-0361).
formance as long as needed (e.g., until the operator
returns home after the work shift and can safely
obtain adequate, recuperative sleep) (Balkin et al., Abbreviations
2011). Examples of instruments currently being
examined for such use include EEG measurement COWAT Controlled Oral Word Association
of brain wave activity, ocular measures, video-based Test
monitoring systems and portable assessments of EEG electroencephalography
attention, and reaction time. Fatigue detection and KSS Karolinska Sleepiness Scale
prediction instruments, implemented in addition to MSLT Multiple Sleep Latency Test
fatigue-related regulations, will help ensure work- MWT Maintenance of Wakefulness Test
place and road safety and adequate levels of alert- NREM non-rapid eye movement
ness in workers while on duty (Balkin et al., 2011). PSG polysomnography
PVT Psychomotor Vigilance Test
Conclusion REM rapid eye movement
SSS Stanford Sleepiness Scale
Despite being treated as a commodity that can be SWE slow wave energy
traded for other activities, sleep is essential for TIB time in bed
optimal neurobehavioral functioning. The 24/7
schedule of many occupations in current indus-
trialized societies as well as other social- and References
work-related expediencies has produced a high
percentage of habitually sleep-restricted adults, a Abe, T., Komada, Y., Nishida, Y., Hayashida, K., & Inoue, Y.
trend that will to continue in the decades ahead. (2010). Short sleep duration and long spells of driving are
associated with the occurrence of Japanese drivers’ rear-
Although chronic partial sleep restriction leads
end collisions and single-car accidents. Journal of Sleep
to increased sleep propensity and neuro- Research, 19, 310–316.
behavioral deficits, adults are not adept at accu- Aeschbach, D., Cajochen, C., Landolt, H., & Borbély, A. A.
rately judging how affected they are by this (1996). Homeostatic sleep regulation in habitual short
393

sleepers and long sleepers. American Journal of Physiology, Banks, S., & Dinges, D. F. (2007). Behavioral and physiologi-
270, R41–R53. cal consequences of sleep restriction. Journal of Clinical
Åkerstedt, T. (2003). Shift work and disturbed sleep/wakeful- Sleep Medicine, 3, 519–528.
ness. Occupational Medicine (Oxford), 53, 89–94. Banks, S., Van Dongen, H. P. A., Maislin, G., & Dinges, D. F.
Åkerstedt, T. (2005). Shift work and sleep disorders. Sleep, 28, (2010). Neurobehavioral dynamics following chronic sleep
9–11. restriction: Dose-response effects of one night for recovery.
Åkerstedt, T., & Froberg, J. E. (1978). Persistence of circadian- Sleep, 33, 1013–1026.
rhythms in phenomenological and physiological arousal Barger, L. K., Lockley, S. W., Rajaratnam, S. M. W., &
under conditions of continuous activity without sleep. Ergo- Landrigan, C. P. (2009). Neurobehavioral, health, and safety
nomics, 21, 866. consequences associated with shift work in safety-sensitive
Åkerstedt, T., & Gillberg, M. (1990). Subjective and objective professions. Current Neurology and Neuroscience Reports,
sleepiness in the active individual. International Journal of 9, 155–164.
Neuroscience, 52, 29–37. Basner, M., & Dinges, D. F. (2009). Dubious bargain: Trading
Åkerstedt, T., Kecklund, G., Ingre, M., Lekander, M., & sleep for Leno and Letterman. Sleep, 32, 747–752.
Axelsson, J. (2009). Sleep homeostasis during repeated Basner, M., Fomberstein, K. M., Razavi, F. M., Banks, S.,
sleep restriction and recovery: Support from EEG dynamics. William, J. H., Rosa, R. R., et al. (2007). American time
Sleep, 32, 217–222. use survey: Sleep time and its relationship to waking
Åkerstedt, T., Peters, B., Anund, A., & Kecklund, G. (2005). activities. Sleep, 30, 1085–1095.
Impaired alertness and performance driving home from Basner, M., Muller, U., & Elmenhorst, E. M. (2011). Single
the night shift: A driving simulator study. Journal of Sleep and combined effects of air, road, and rail traffic noise on
Research, 14, 17–20. sleep and recuperation. Sleep, 34, 11–23.
Åkerstedt, T., Torsvall, L., & Gillberg, M. (1987). Sleepiness in Belenky, G., Wesensten, N. J., Thorne, D. R., Thomas, M. L.,
shiftwork. A review with emphasis on continuous monitoring Sing, H. C., Redmond, D. P., et al. (2003). Patterns of per-
of EEG and EOG. Chronobiology International, 4, 129–140. formance degradation and restoration during sleep restric-
Åkerstedt, T., & Wright, K. P. Jr. (2009). Sleep loss and tion and subsequent recovery: A sleep dose-response
fatigue in shift work and shift work disorder. Sleep Medicine study. Journal of Sleep Research, 12, 1–12.
Clinics, 4, 257–271. Bjerner, B. (1949). Alpha depression and lowered pulse rate
Alhola, P., & Polo-Kantola, P. (2007). Sleep deprivation: during delayed actions in a serial reaction test. Acta Physi-
Impact on cognitive performance. Neuropsychiatric Disease ologica Scandinavica, 19, 1–93.
and Treatment, 3, 553–567. Blackwell, T., Redline, S., Ancoli-Israel, S., Schneider, J. L.,
Aschoff, J., & Wever, R. (1981). The circadian system of man. Surovec, S., Johnson, N. L., et al. (2008). Comparison of
In J. Aschoff (Ed.), Handbook of behavioral neurobiology sleep parameters from actigraphy and polysomnography in
(pp. 311–331). New York: Plenum Press. older women: The SOF study. Sleep, 31, 283–291.
Axelsson, J., Kecklund, G., Åkerstedt, T., Donofrio, P., Bliese, P. D., Wesensten, N. J., & Balkin, T. J. (2006). Age
Lekander, M., & Ingre, M. (2008). Sleepiness and perfor- and individual variability in performance during sleep
mance in response to repeated sleep restriction and restriction. Journal of Sleep Research, 15, 376–385.
subsequent recovery during semi-laboratory conditions. Borbély, A. (1982). Endogenous sleep-promoting substances.
Chronobiology International, 25, 297–308. Trends in Pharmacological Sciences, 3, 350.
Baehr, E. K., Revelle, W., & Eastman, C. I. (2000). Individual Borbély, A. (1994). Sleep homeostasis and models of sleep
differences in the phase and amplitude of the human circadian regulation. In M. Kryger, T. Roth & W. Dement (Eds.),
temperature rhythm: With an emphasis on morningness- Principles and practice of sleep medicine. Philadelphia: W.
eveningness. Journal of Sleep Research, 9, 117–127. B. Saunders.
Balkin, T. J., Horrey, W. J., Graeber, R. C., Czeisler, C. A., & Borbély, A. A. (1998). Processes underlying sleep regulation.
Dinges, D. F. (2011). The challenges and opportunities of Hormone Research, 49, 114–117.
technological approaches to fatigue management. Accident Borbély, A. A., & Tobler, I. (1996). Sleep regulation: Relation
Analysis and Prevention, 43, 565–572. to photoperiod, sleep duration, waking activity, and torpor.
Balkin, T. J., Kamimori, G. H., Redmond, D. P., Progress in Brain Research, 111, 343–348.
Vigneulle, R. M., Thorne, D. R., Belenky, G., et al. (2004). Bradshaw, D. A., Yanagi, M. A., Pak, E. S., Peery, T. S., &
On the importance of countermeasures in sleep and perfor- Ruff, G. A. (2007). Nightly sleep duration in the 2-week
mance models. Aviation, Space, and Environmental Medicine, period preceding multiple sleep latency testing. Journal of
75, A155–A157. Clinical Sleep Medicine, 3, 613–619.
Balkin, T. J., Rupp, T., Picchioni, D., & Wesensten, N. J. (2008). Brainard, G. C., Hanifin, J. P., Rollag, M. D., Greeson, J.,
Sleep loss and sleepiness—Current issues. Chest, 134, 653–660. Byrne, B., Glickman, G., et al. (2001). Human melatonin
394

regulation is not mediated by the three cone photopic visual Czeisler, C. A. (2009). Medical and genetic differences in
system. Journal of Clinical Endocrinology and Metabolism, the adverse impact of sleep loss on performance: Ethical
86, 433–436. considerations for the medical profession. Transactions of
Brown, I. D. (1994). Driver fatigue. Human Factors, 36, the American Clinical and Climatological Association, 120,
298–314. 249–285.
Brunborg, G. S., Mentzoni, R. A., Molde, H., Myrseth, H., Czeisler, C. A., Duffy, J. F., Shanahan, T. L., Brown, E. N.,
Skouverøe, K. J., Bjorvatn, B., et al. (2011). The relation- Mitchell, J. F., Rimmer, D. W., et al. (1999). Stability, preci-
ship between media use in the bedroom, sleep habits and sion, and near-24-hour period of the human circadian
symptoms of insomnia. Journal of Sleep Research, 4, pacemaker. Science, 284, 2177–2181.
569–575. Daan, S., Beersma, D. G. M., & Borbély, A. A. (1984). Timing
Brunner, D. P., Dijk, D. J., & Borbély, A. A. (1993). Repeated of human sleep—Recovery process gated by a circadian
partial sleep-deprivation progressively changes the EEG pacemaker. American Journal of Physiology, 246,
during sleep and wakefulness. Sleep, 16, 100–113. R161–R178.
Brunner, D. P., Dijk, D. J., Tobler, I., & Borbély, A. A. (1990). De Valck, E., & Cluydts, R. (2003). Sleepiness as a state-trait
Effect of partial sleep deprivation on sleep stages and EEG phenomenon, comprising both a sleep drive and a wake
power spectra: Evidence for non-REM and REM sleep drive. Medical Hypotheses, 60, 509–512.
homeostasis. Electroencephalography and Clinical Neuro- Dement, W., & Greenber, S. (1966). Changes in total amount
physiology, 75, 492–499. of stage 4 sleep as a function of partial sleep deprivation.
Bureau of Labor Statistics, (2005). Workers on Flexible and Electroencephalography and Clinical Neurophysiology, 20,
Shift Schedules in 2004 Summary. Washington, DC: U.S. 523–526.
Department of Labor Economic News Release. Devoto, A., Lucidi, F., Violani, C., & Bertini, M. (1999).
Burgess, H. L., Sharkey, K. M., & Eastman, C. I. (2002). Bright Effects of different sleep reductions on daytime sleepiness.
light, dark and melatonin can promote circadian adaptation Sleep, 22, 336–343.
in night shift workers. Sleep Medicine Reviews, 6, 407–420. Dijk, D. J., Beersma, D. G., Daan, S., & Lewy, A. J. (1989).
Cajochen, C., Frey, S., Anders, D., Spati, J., Bues, M., Bright morning light advances the human circadian system
Pross, A., et al. (2011). Evening exposure to a light-emitting without affecting NREM sleep homeostasis. American
diodes (LED)-backlit computer screen affects circadian Journal of Physiology, 256, R106–R111.
physiology and cognitive performance. Journal of Applied Dinges, D. F. (1990). Are you awake? In R. Bootzin, J.
Physiology, 110, 1432–1438. Kihlstrom & D. Schacter (Eds.), Sleep and cognition
Cajochen, C., Jud, C., Munch, M., Kobialka, S., Wirz- (pp. 159–175). Washington, DC: American Psychological
Justice, A., & Albrecht, U. (2006). Evening exposure to blue Association.
light stimulates the expression of the clock gene PER2 in Dinges, D. F. (1992). Probing the limits of functional capabil-
humans. European Journal of Neuroscience, 23, 1082–1086. ity: The effects of sleep loss on short-duration tasks. In
Carskadon, M. A., & Dement, W. C. (1981). Cumulative R. J. Broughton & R. D. Oglivie (Eds.), Sleep, arousal,
effects of sleep restriction on daytime sleepiness. Psycho- and performance (pp. 176–188). Boston: Birkhauser.
physiology, 18, 107–113. Dinges, D. F. (2004). Sleep debt and scientific evidence. Sleep,
Centers for Disease Control and Prevention, (2008). 27, 1050–1052.
QuickStats: Percentage of adults aged > 18 years who Dinges, D. F., & Kribbs, N. B. (1991). Performing while sleepy:
reported an average of < 6 hours of sleep per 24-Hour Effects of experimentally-induced sleepiness. In T. H. Monk
period, by sex and age group - National Health Interview (Ed.), Sleep, sleepiness and performance. Chister: Wiley.
Survey, United States, 1985 and 2006. Morbidity and Mor- Dinges, D. F., Orne, M. T., Whitehouse, W. G., & Orne, E. C.
tality Weekly Report, 57, 209. (1987). Temporal placement of a nap for alertness: Con-
Census American Community Survey Report, (2011). tributions of circadian phase and prior wakefulness. Sleep,
Workers on Flexible and Shift Schedules in 2004 Summary. 10, 313–329.
Washington, DC: U.S. Department of Commerce American Dinges, D. F., Pack, F., Williams, K., Gillen, K. A.,
Community Survey Reports. Powell, J. W., Ott, G. E., et al. (1997). Cumulative sleepi-
Cote, K. A., Milner, C. E., Osip, S. L., Baker, M. L., & ness, mood disturbance, and psychomotor vigilance perfor-
Cuthbert, B. P. (2008). Physiological arousal and attention mance decrements during a week of sleep restricted to 4-5
during a week of continuous sleep restriction. Physiology hours per night. Sleep, 20, 267–277.
and Behavior, 95, 353–364. Doran, S. M., Van Dongen, H. P. A., & Dinges, D. F. (2001).
Czeisler, C. A. (1995). The effect of light on the human circa- Sustained attention performance during sleep deprivation:
dian pacemaker. Circadian Clocks and Their Adjustment, Evidence of state instability. Archives Italiennes de Biologie,
183, 254–290. 139, 253–267.
395

Dorrian, J., Rogers, N., & Dinges, D. (2005). Psychomotor vig- Harrison, Y., & Horne, J. A. (1996). Occurrence of
ilance performance: Neurocognitive assay sensitive to sleep ‘microsleeps’ during daytime sleep onset in normal subjects.
loss. In C. Kushida (Ed.), Sleep deprivation. New York: Electroencephalography and Clinical Neurophysiology, 98,
Marcel Dekker. 411–416.
Everson, C. A., Bergmann, B. M., & Rechtschaffen, A. (1989). Harrison, Y., & Horne, J. A. (2000). The impact of sleep dep-
Sleep-deprivation in the rat. 3. Total sleep-deprivation. rivation on decision making: A review. Journal of Experi-
Sleep, 12, 13–21. mental Psychology: Applied, 6, 236–249.
Fafrowicz, M., Oginska, H., Mojsa-Kaja, J., Marek, T., Herscovitch, J., Stuss, D., & Broughton, R. (1980). Changes
Golonka, K., & Tucholska, K. (2010). Chronic sleep deficit in cognitive processing following shortterm cumulative
and performance of a sustained attention task-an electrooc- partial sleep deprivation and recovery oversleeping.
ulography study. Chronobiology International, 27, 934–944. Journal of Clinical and Experimental Neuropsychology, 2,
Frey, D. J., Badia, P., & Wright, K. P. (2004). Inter- and intra- 301–319.
individual variability in performance near the circadian Hoddes, E., Zarcone, V., Smythe, H., Phillips, R., &
nadir during sleep deprivation. Journal of Sleep Research, Dement, W. C. (1973). Quantification of sleepiness: A new
13, 305–315. approach. Psychophysiology, 10, 431–436.
Gander, P., Hartley, L., Powell, D., Cabon, P., Hitchcock, E., Horne, J. A. (1985). Sleep function, with particular reference to
Mills, A., et al. (2011). Fatigue risk management: Organiza- sleep-deprivation. Annals of Clinical Research, 17, 199–208.
tional factors at the regulatory and industry/company level. Horne, J. (1988). The substance of sleep. New Scientist, 117,
Accident Analysis and Prevention, 43, 573–590. 60–62.
Gander, P. H., Marshall, N. S., Harris, R. B., & Reid, P. Horne, J. (2004). Is there a sleep debt? Sleep, 27, 1047–1049.
(2005). Sleep, sleepiness and motor vehicle accidents: Horne, J. A., & Pettitt, A. N. (1985). High incentive effects
A national survey. Australian and New Zealand Journal of on vigilance performance during 72 hours of total sleep-
Public Health, 29, 16–21. deprivation. Acta Psychologica, 58, 123–139.
Goel, N. (2011). Genetics of sleep timing, duration, and Horne, J., & Reyner, L. (1999). Vehicle accidents related to
homeostasis in humans. Sleep Medicine Clinics, 6, 171–182. sleep: A review. Occupational and Environmental Medicine,
Goel, N., Banks, S., Mignot, E., & Dinges, D. F. (2009). PER3 56, 289–294.
polymorphism predicts cumulative sleep homeostatic but Hsieh, S. L., Li, T. H., & Tsai, L. L. (2010). Impact of mone-
not neurobehavioral changes to chronic partial sleep depriva- tary incentives on cognitive performance and error monitor-
tion. PLoS One, 4, e5874. http://dx.doi.org/10.1371/journal. ing following sleep deprivation. Sleep, 33, 499–507.
pone.0005874. Jennings, J. R., Monk, T. H., & van der Molen, M. W. (2003).
Goel, N., Banks, S., Mignot, E., & Dinges, D. F. (2010). Sleep deprivation influences some but not all processes of
DQB1*0602 predicts interindividual differences in physio- supervisory attention. Psychological Science, 14, 473–479.
logic sleep, sleepiness, and fatigue. Neurology, 75, 1509–1519. Jones, K., & Harrison, Y. (2001). Frontal lobe function, sleep
Goel, N., & Dinges, D. F. (2011). Behavioral and genetic loss and fragmented sleep. Sleep Medicine Reviews, 5,
markers of sleepiness. Journal of Clinical Sleep Medicine, 463–475.
7, S19–S21. Kalinchuk, A. V., Lu, Y., Stenberg, D., Rosenberg, P. A., &
Goel, N., Rao, H., Durmer, J. S., & Dinges, D. F. (2009). Porkka-Heiskanen, T. (2006). Nitric oxide production in
Neurocognitive consequences of sleep deprivation. Seminars the basal forebrain is required for recovery sleep. Journal
in Neurology, 29, 320–339. of Neurochemistry, 99, 483–498.
Graw, P., Krauchi, K., Knoblauch, V., Wirz-Justice, A., & Kerkhof, G. A., & VanDongen, H. P. A. (1996). Morning-type
Cajochen, C. (2004). Circadian and wake-dependent modula- and evening-type individuals differ in the phase position of
tion of fastest and slowest reaction times during the psychomo- their endogenous circadian oscillator. Neuroscience Letters,
tor vigilance task. Physiology and Behavior, 80, 695–701. 218, 153–156.
Groeger, J. A., Viola, A. U., Lo, J. C. Y., von Schantz, M., Killgore, W. D. (2010). Effects of sleep deprivation on cogni-
Archer, S. N., & Dijk, D. J. (2008). Early morning executive tion. Progress in Brain Research, 185, 105–129.
functioning during sleep deprivation is compromised by a Kleitman, N. (1963). Sleep and wakefulness. Chicago: Univer-
PERIOD3 polymorphism. Sleep, 31, 1159–1167. sity of Chicago.
Guilleminault, C., Powell, N. B., Martinez, S., Kushida, C., Klerman, E. B., & Dijk, D. J. (2005). Interindividual variation
Raffray, T., Palombini, L., et al. (2003). Preliminary in sleep duration and its association with sleep debt in young
observations on the effects of sleep time in a sleep restric- adults. Sleep, 28, 1253–1259.
tion paradigm. Sleep Medicine, 4, 177–184. Knutson, K. L., Van Cauter, E., Rathouz, P. J., DeLeire, T., &
Hale, L. (2005). Who has time to sleep? Journal of Public Lauderdale, D. S. (2010). Trends in the prevalence of short
Health Medicine, 27, 205–211. sleepers in the USA: 1975-2006. Sleep, 33, 37–45.
396

Kolla, B. P., & Auger, R. R. (2011). Jet lag and shift work Luckhaupt, S. E., Tak, S., & Calvert, G. M. (2010). The prev-
sleep disorders: How to help reset the internal clock. alence of short sleep duration by industry and occupation in
Cleveland Clinic Journal of Medicine, 78, 675–684. the National Health Interview Survey. Sleep, 33, 149–159.
Korczak, A. L., Martynhak, B. J., Pedrazzoli, M., Brito, A. F., Macchi, M. M., Boulos, Z., Ranney, T., Simmons, L., &
& Louzada, F. M. (2008). Influence of chronotype and social Campbell, S. S. (2002). Effects of an afternoon nap on night-
zeitgebers on sleep/wake patterns. Brazilian Journal of time alertness and performance in long-haul drivers. Acci-
Medical and Biological Research, 41, 914–919. dent Analysis and Prevention, 34, 825–834.
Krueger, P. M., & Friedman, E. M. (2009). Sleep duration in McCauley, P., Kalachev, L. V., Smith, A. D., Belenky, G.,
the United States: A cross-sectional population-based study. Dinges, D. F., & Van Dongen, H. P. A. (2009). A new math-
American Journal of Epidemiology, 169, 1052–1063. ematical model for the homeostatic effects of sleep loss on
Kryger, M. H., Roth, T., & Dement, W. C. (Eds.), (2011). neurobehavioral performance. Journal of Theoretical Biol-
Principles and practice of sleep medicine. St. Louis: Elsevier ogy, 256, 227–239.
Saunders. McCoy, J., & Strecker, R. (2011). The cognitive cost of sleep
Lauderdale, D. S., Knutson, K. L., Yan, L. L., Liu, K., & lost. Neurobiology of Learning and Memory, doi:10.1016/j.
Rathouz, P. J. (2008). Self-reported and measured sleep nlm.2011.07.004.
duration: How similar are they? Epidemiology, 19, 838–845. McKnight-Eily, L. R., Liu, Y., Croft, J. B., Perry, G. S., &
Lenne, M. G., Dwyer, F., Triggs, T. J., Rajaratnam, S., & Strine, T. (2011). Unhealthy sleep-related behaviors—12
Redman, J. R. (2004). The effects of a nap opportunity in States, 2009 (pp. 223–238). Atlanta, GA: U.S. Department
quiet and noisy environments on driving performance. of Health and Human Services Centers for Disease Control
Chronobiology International, 21, 991–1001. and Prevention.
Leproult, R., Colecchia, E. F., Berardi, A. M., Stickgold, R., Mitler, M. M., Carskadon, M. A., Czeisler, C. A.,
Kosslyn, S. M., & Van Cauter, E. (2003). Individual Dement, W. C., Dinges, D. F., & Graeber, R. C. (1988).
differences in subjective and objective alertness during sleep Catastrophes, sleep, and public-policy—Consensus report.
deprivation are stable and unrelated. American Journal of Sleep, 11, 100–109.
Physiology. Regulatory, Integrative and Comparative Physi- Mitler, M. M., Gujavarty, K. S., & Browman, C. P. (1982).
ology, 284, 280–290. Maintenance of Wakefulness Test—A polysomnographic
Leproult, R., & Van Cauter, E. (2010). Role of sleep and sleep technique for evaluating treatment efficacy in patients with
loss in hormonal release and metabolism. Endocrine Devel- excessive somnolence. Electroencephalography and Clinical
opment, 17, 11–21. Neurophysiology, 53, 658–661.
Lim, J., & Dinges, D. F. (2008). Sleep deprivation and vigilant Mollicone, D. J., Van Dongen, H. P. A., Rogers, N. L.,
attention. Annals of the New York Academy of Sciences, Banks, S., & Dinges, D. F. (2010). Time of day effects on
1129, 305–322. neurobehavioral performance during chronic sleep restric-
Lim, J., & Dinges, D. F. (2010). A meta-analysis of the impact tion. Aviation, Space, and Environmental Medicine, 81,
of short-term sleep deprivation on cognitive variables. 735–744.
Psychological Bulletin, 136, 375–389. Nakashima, M., Morikawa, Y., Sakurai, M., Nakamura, K.,
Lockley, S. W., Brainard, G. C., & Czeisler, C. A. (2003). High Miura, K., Ishizaki, M., et al. (2010). Association between
sensitivity of the human circadian melatonin rhythm to long working hours and sleep problems in white-collar
resetting by short wavelength light. Journal of Clinical workers. Journal of Sleep Research, 20, 110–116.
Endocrinology and Metabolism, 88, 4502–4505. Nansen, F. (Ed.), (1999). Farthest north. New York: The Mod-
Lockley, S. W., Cronin, J. W., Evans, E. E., Cade, B. E., ern Library.
Lee, C. J., Landrigan, C. P., et al. (2004). Effect of reducing National Transportation Safety Board, (1995). Factors that
interns’ weekly work hours on sleep and attentional failures. affect fatigue in heavy truck accidents. Washington, DC:
The New England Journal of Medicine, 351, 1829–1837. National Transportation Safety Board.
Lombardi, D. A., Folkard, S., Willetts, J. L., & Smith, G. S. Ohayon, M. M., Smolensky, M. H., & Roth, T. (2010).
(2010). Daily sleep, weekly working hours, and risk Consequences of shiftworking on sleep duration, sleepi-
of work-related injury: US National Health Interview ness, and sleep attacks. Chronobiology International, 27,
Survey (2004-2008). Chronobiology International, 27, 575–589.
1013–1030. Olofsen, E., Dinges, D. F., & Van Dongen, H. P. (2004). Non-
Lucas, R. J., Stirland, J. A., Darrow, J. M., Menaker, M., & linear mixed-effects modeling: Individualization and predic-
Loudon, A. S. I. (1999). Free running circadian rhythms tion. Aviation, Space, and Environmental Medicine, 75,
of melatonin, luteinizing hormone, and cortisol in Syrian A134–A140.
hamsters bearing the circadian tau mutation. Endocrinology, Otmani, S., Pebayle, T., Roge, J., & Muzet, A. (2005). Effect
140, 758–764. of driving duration and partial sleep deprivation on
397

subsequent alertness and performance of car drivers. Physi- Roepke, S. E., & Duffy, J. F. (2010). Differential impact of
ology and Behavior, 84, 715–724. chronotype on weekday and weekend sleep timing and
Philip, P., Sagaspe, P., Taillard, J., Moore, N., Guilleminault, C., duration. Nature and Science of Sleep, 2010, 213–220.
Sanchez-Ortuno, M., et al. (2003). Fatigue, sleep restriction, Rones, P. L., Iig, R. E., & Gardner, J. M. (1997). Trends in
and performance in automobile drivers: A controlled study hours of work since the mid-1970s. Monthly Labor Review,
in a natural environment. Sleep, 26, 277–280. 120, 3–14.
Philip, P., Sagaspe, P., Taillard, J., Valtat, C., Moore, N., Rupp, T. L., Wesensten, N. J., Bliese, P. D., & Balkin, T. J.
Åkerstedt, T., et al. (2005). Fatigue, sleepiness, and perfor- (2009). Banking sleep: Realization of benefits during
mance in simulated versus real driving conditions. Sleep, subsequent sleep restriction and recovery. Sleep, 32,
28, 1511–1516. 311–321.
Pittendrigh, C. S. (1981). Circadian systems: Entrainment. In Russo, A., Thomas, A., Thorne, D., Sing, H., Redmond, D.,
J. Aschoff (Ed.), Handbook of behavioral neurobiology Rowland, L., et al. (2003). Oculomotor impairment during
(pp. 95–124). New York: Plenum Press. chronic partial sleep deprivation. Clinical Neurophysiology,
Plessow, F., Kiesel, A., Petzold, A., & Kirschbaum, C. (2010). 114, 723–736.
Chronic sleep curtailment impairs the flexible implementa- Sadun, A. A., Schaechter, J. D., & Smith, L. E. H. (1984).
tion of task goals in new parents. Journal of Sleep Research, A retinohypothalamic pathway in man—Light mediation
20, 279–287. of circadian-rhythms. Brain Research, 302, 371–377.
Poe, G. R., Walsh, C. M., & Bjorness, T. E. (2010). Cognitive Saper, C. B., Scammell, T. E., & Lu, J. (2005). Hypothalamic
neuroscience of sleep. Progress in Brain Research, 185, regulation of sleep and circadian rhythms. Nature, 437,
1–19. 1257–1263.
Porkka-Heiskanen, T., & Kalinchuk, A. V. (2011). Adenosine, Scharf, M. T., Naidoo, N., Zimmerman, J. E., & Pack, A. I.
energy metabolism and sleep homeostasis. Sleep Medicine (2008). The energy hypothesis of sleep revisited. Progress
Reviews, 15, 123–135. in Neurobiology, 86, 264–280.
Punjabi, N. M., Bandeen-Roche, K., & Young, T. (2003). Pre- Scott, L. D., Hwang, W. T., Rogers, A. E., Nysse, T.,
dictors of objective sleep tendency in the general popula- Dean, G. E., & Dinges, D. F. (2007). The relationship
tion. Sleep, 26, 678–683. between nurse work schedules, sleep duration, and drowsy
Ralph, M. R., Foster, R. G., Davis, F. C., & Menaker, M. driving. Sleep, 30, 1801–1807.
(1990). Transplanted suprachiasmatic nucleus determines Seugnet, L., Boero, J., Gottschalk, L., Duntley, S. P., &
circadian period. Science, 247, 975–978. Shaw, P. J. (2006). Identification of a biomarker for sleep
Raslear, T. G., Hursh, S. R., & Van Dongen, H. P. (2011). drive in flies and humans. Proceedings of the National Acad-
Predicting cognitive impairment and accident risk. Progress emy of Sciences of the United States of America, 103,
in Brain Research, 190, 155–167. 19913–19918.
Rechtschaffen, A. (1998). Current perspectives on the function Shanahan, T. L., & Czeisler, C. A. (1991). Light exposure
of sleep. Perspectives in Biology and Medicine, 41, 359–390. induces equivalent phase-shifts of the endogenous
Rechtschaffen, A., Gilliland, M. A., Bergmann, B. M., & circadian-rhythms of circulating plasma melatonin and core
Winter, J. B. (1983). Physiological correlates of prolonged body-temperature in men. Journal of Clinical Endocrinol-
sleep-deprivation in rats. Science, 221, 182–184. ogy and Metabolism, 73, 227–235.
Rechtschaffen, A., & Kales, A. (1968). A manual of Shaw, P. J., Tononi, G., Greenspan, R. J., & Robinson, D. F.
standardized terminology, techniques and scoring system of (2002). Stress response genes protect against lethal
sleep stages in human subjects. Los Angeles: Brain Informa- effects of sleep deprivation in Drosophila. Nature, 417,
tion Service/Brain Research Institute, University of 287–291.
California. Siegel, J. M. (2005). Clues to the functions of mammalian
Roehrs, T., Carskadon, M. A., Dement, W. C., & Roth, T. sleep. Nature, 437, 1264–1271.
(2005). Daytime sleepiness and alertness. In M. H. Silber, M. H., Ancoli-Israel, S., Bonnet, M. H., Chokroverty, S.,
Kryger, T. Roth & W. C. Dement (Eds.), Principles and Grigg-Damberger, M. M., Hirshkowitz, M., et al. (2007). The
practice of sleep medicine. New York: Saunders Company. visual scoring of sleep in adults. Journal of Clinical Sleep
Roenneberg, T., Kuehnle, T., Juda, M., Kantermann, T., Medicine, 3, 121–131.
Allebrandt, K., Gordijn, M., et al. (2007). Epidemiology of Stenuit, P., & Kerkhofs, M. (2008). Effects of sleep
the human circadian clock. Sleep Medicine Reviews, 11, restriction on cognition in women. Biological Psychology,
429–438. 77, 81–88.
Roenneberg, T., Wirz-Justice, A., & Merrow, M. (2003). Life Stenuit, K. P., Blatt, J., Forrest, C., Georges, K., Kiley, J.,
between clocks: Daily temporal patterns of human Kurrus, R., et al. (1997). NCSDR/NHTSA Expert Panel on
chronotypes. Journal of Biological Rhythms, 18, 80–90. Driver Fatigue and Sleepiness.
398

Stutts, J. C., Wilkins, J. W., Scott Osberg, J., & Vaughn, B. V. Van Dongen, H. P., Vitellaro, K. M., & Dinges, D. F. (2005).
(2003). Driver risk factors for sleep-related crashes. Acci- Individual differences in adult human sleep and wakeful-
dent Analysis and Prevention, 35, 321–331. ness: Leitmotif for a research agenda. Sleep, 28, 479–496.
Swann, C. E., Yelland, G. W., Redman, J. R., & Vandewalle, G., Archer, S. N., Wuillaume, C., Balteau, E.,
Rajaratnam, S. M. W. (2006). Chronic partial sleep loss Degueldre, C., Luxen, A., et al. (2009). Functional magnetic
increases the facilitatory role of a masked prime in a word resonance imaging-assessed brain responses during an exec-
recognition task. Journal of Sleep Research, 15, 23–29. utive task depend on interaction of sleep homeostasis, circa-
Taillard, J., Philip, P., Coste, O., Sagaspe, P., & Bioulac, B. dian phase, and PER3 genotype. Journal of Neuroscience,
(2003). The circadian and homeostatic modulation of sleep 29, 7948–7956.
pressure during wakefulness differs between morning and Viola, A. U., Archer, S. N., James, L. M., Groeger, J. A.,
evening chronotypes. Journal of Sleep Research, 12, 275–282. Lo, J. C. Y., Skene, D. J., et al. (2007). PER3 polymorphism
Tobler, I. (1995). Is sleep fundamentally different between predicts sleep structure and waking performance. Current
mammalian species? Behavioural Brain Research, 69, 35–41. Biology, 17, 613–618.
Torsvall, L., & Åkerstedt, T. (1987). Sleepiness on the job— Virtanen, M., Ferrie, J. E., Gimeno, D., Vahtera, J.,
Continuously measured EEG changes in train drivers. Elovainio, M., Singh-Manoux, A., et al. (2009). Long work-
Electroencephalography and Clinical Neurophysiology, 66, ing hours and sleep disturbances: The Whitehall II prospec-
502–511. tive cohort study. Sleep, 32, 737–745.
National Transportation Safety Board, (1989). Marine Accident Watts, A. G. (1991). The efferent projections of the
Report: Grounding of the U.S. Tankship EXXON VALDEZ suprachiasmatic nucleus. Anatomical insights into the con-
on Bligh Reef. Prince William Sound near Valdez, Alaska. trol of circadian rhythms. In D. C. Klein, R. Y. Moore &
Washington, DC: National Transportation Safety Board. S. M. Reppert (Eds.), Suprachiasmatic nucleus: The mind’s
Trew, A., Searles, B., Smith, T., & Darling, E. M. (2011). clock (pp. 77–106). New York: Oxford University Press.
Fatigue and extended work hours among cardiovascular Wehr, T. A., Aeschbach, D., & Duncan, W. C. Jr. (2001). Evi-
perfusionists: 2010 Survey. Perfusion, 26, 361–370. dence for a biological dawn and dusk in the human circadian
Vakulin, A., Baulk, S. D., Catcheside, P. G., Anderson, R., timing system. The Journal of Physiology, 535, 937–951.
van den Heuvel, C. J., Banks, S., et al. (2007). Effects of ]West, K. E., Jablonski, M. R., Warfield, B., Cecil, K. S.,
moderate sleep deprivation and low-dose alcohol on driving James, M., Ayers, M. A., et al. (2011). Blue light from light-
simulator performance and perception in young men. Sleep, emitting diodes elicits a dose-dependent suppression of mela-
30, 1327–1333. tonin in humans. Journal of Applied Physiology, 110, 619–626.
Van Dongen, H. P., Baynard, M. D., Maislin, G., & Whitney, P., & Hinson, J. M. (2010). Measurement of cogni-
Dinges, D. F. (2004). Systematic interindividual differences tion in studies of sleep deprivation. Progress in Brain
in neurobehavioral impairment from sleep loss: Evidence of Research, 185, 37–48.
trait-like differential vulnerability. Sleep, 27, 423–433. Wilkinson, R. T. (1969). Loss of sleep. Proceedings of the
Van Dongen, H. P., & Dinges, D. F. (2003). Investigating the Royal Society of Medicine, 62, 903–904.
interaction between the homeostatic and circadian processes Wittmann, M., Dinich, J., Merrow, M., & Roenneberg, T.
of sleep-wake regulation for the prediction of waking (2006). Social jetlag: Misalignment of biological and social
neurobehavioural performance. Journal of Sleep Research, time. Chronobiology International, 23, 497–509.
12, 181–187. Wu, H., Stone, W. S., Hsi, X., Zhuang, J., Huang, L., Yin, Y.,
Van Dongen, H. P., & Dinges, D. F. (2005). Sleep, circadian et al. (2010). Effects of different sleep restriction protocols
rhythms, and psychomotor vigilance. Clinics in Sports Medi- on sleep architecture and daytime vigilance in healthy
cine, 24, 237–249 vii–viii. men. Physiological Research, 59, 821–829.
Van Dongen, H., Dinges, D. F., & Olofsen, E. (2004). Bayesian Zhou, X., Ferguson, S. A., Matthews, R. W., Sargent, C.,
individualization of biomathematical predictions for cognitive Darwent, D., Kennaway, D. J., et al. (2011a). Sleep, wake
performance impairment from sleep loss. Sleep, 27, 144. and phase dependent changes in neurobehavioral function
Van Dongen, H. P., Maislin, G., Mullington, J. M., & under forced desynchrony. Sleep, 34, 931–941.
Dinges, D. F. (2003). The cumulative cost of additional Zhou, X. A., Ferguson, S. A., Matthews, R. W., Sargent, C.,
wakefulness: Dose-response effects on neurobehavioral Darwent, D., Kennaway, D. J., et al. (2011b). Dynamics of
functions and sleep physiology from chronic sleep restriction neurobehavioral performance variability under forced
and total sleep deprivation. Sleep, 26, 117–126. desynchrony: Evidence of state instability. Sleep, 34, 57–61.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 22

Noisy and individual, but doable: Shift-work


research in humans

Thomas Kantermann{,1, Sophie M. T. Wehrens{, Melissa A. Ulhôa{, Claudia Moreno{


and Debra J. Skene{,*

{
Centre for Chronobiology, Faculty of Health and Medical Sciences, University of Surrey, Guildford, Surrey,
United Kingdom
{
University of São Paulo, School of Public Health, São Paulo, Brazil

Abstract: Working around the clock is common for many occupations, as diverse as nurses, truck
drivers, physicians, steel workers, and pilots. Each shift-work profession is individual in more aspects
than just work hours and individual work scenarios, each posing a different impact on the health of
workers. Related health problems in shift workers, therefore, are also diverse and encompass sleep
problems, metabolic and cardiovascular system disturbances, as well as cancer. Little is known about
how all these individual factors influence a shift worker’s health status, partly because many shift-work
studies show inconsistent results. In addition, these individual factors create many methodological
difficulties for researchers who investigate such work scenarios. This chapter presents examples from
our laboratory and field studies of shift workers, which emphasize the importance of taking individual
circumstances into account. Both study approaches, laboratory and field based, are needed to fully
account for the difficulties that shift-work studies pose on both workers and researchers. Finally,
understanding the mechanisms that underpin interindividual differences in response to shift work will
advance our understanding of how to design better and healthier shift-work schedules in the future.

Keywords: chronotype; social jetlag; stress; sleep; health; shift work; cardiovascular system.

*Corresponding author. Introduction


Tel.: þ44 1483 689706; Fax: þ44 1483 686401
E-mail: d.skene@surrey.ac.uk
Studying shift workers means studying individuals
at a variety of workplaces employed in different
1
Current address: University of Groningen, Chronobiology and often flexible working time arrangements.
Research Unit, Life Science building, Nijenborgh 7, 9747AG Studying the effects of shift work on health is thus
Groningen, The Netherlands complex and poses challenges to researchers to

http://dx.doi.org/10.1016/B978-0-444-59427-3.00022-8 399
400

control the many individual factors—economic, timing system by shift work. Human daily behav-
social, and biological—that characterize modern ior and physiology, including our mental and
shift-work settings. Each of these individual factors physical performance, shows a clear time-of-day
by themselves and in combination may negatively variation, which is governed by an endogenous
affect health (Kantermann et al., 2010; Knutsson, circadian timing system. In temporal isolation,
2004; Saksvik et al., 2011). Therefore, the effects of such internal clocks produce internal days, which
shift work on health are very diverse (Costa, in humans are usually longer than 24 h (Aschoff,
2003b; Kantermann, 2008; Sookoian et al., 2007; 1965; Brown et al., 2008). Under normal con-
Wang et al., 2011), and studies on shift work and ditions, circadian clocks are synchronized to the
health are often inconsistent in their findings. 24-h day by zeitgebers (Roenneberg et al., 2003).
Related health problems in shift workers broadly The light–dark cycle is the most important zeitge-
fall into four categories: (i) sleep problems ber for the human circadian timing system
(Åkerstedt and Wright, 2009; Åkerstedt et al., (Freedman et al., 1999; Kantermann et al., 2007;
2007; Folkard et al., 2005; Lemos et al., 2009; Panda et al., 2002; Roenneberg et al., 2007b).
Moreno et al., 2004), (ii) metabolic problems Synchronization of circadian clocks to a zeitgeber
(Esquirol et al., 2009; Sookoian et al., 2007; Wang cycle, called entrainment, is an active process that
et al., 2011), (iii) cardiovascular system disturbances works via response characteristics which, in turn,
(Boggild and Knutsson, 1999; Hublin et al., 2010; depend on internal time of day (Roenneberg
Puttonen et al., 2010), and (iv) cancer (Erren et al., et al., 2003, 2010): during the late internal night or
2008; Stevens, 2009). Notably, none of the reported early morning, light shortens the internal day, and
health problems constitute an occupational disease lengthens it during internal afternoon or evening,
unique to shift workers. By contrast, all of these whereas there is little or no response to light
health problems belong to a group of “normal” epi- around internal midday. These principles of
demiological and age-related health problems. entrainment are important for understanding the
Interestingly, many of these epidemiological health disruption of the body clock in rotating shift
problems occur more frequently and at an earlier workers when they are exposed to light at unusual
age in shift workers than in the normal non-shift- and continually changing times. The relationship
working population. To date, however, no study between external and internal time is called phase
has tested the hypothesis that exposure to shift work of entrainment, which shows large interindividual
is associated with advanced aging. Obviously, such a differences, often referred to as chronotype
broad range of health problems affects a worker’s (Roenneberg et al., 2007a). In non-shift-work
quality-of-life and life expectancy, which makes populations, for example, late chronotypes have
shift-work studies highly relevant not only for the been shown to be at higher risk of suffering from
health care system but also for society as a whole. bipolar disorders (Wood et al., 2009), headaches
Little is known about how these individual factors (Bruni et al., 2008), seasonal depression (Natale
interact at an economic, social, and biological level et al., 2005), depression (Levandovski et al.,
and how these interacting factors then affect health 2011), and fibromyalgia syndrome (Kantermann
(Kantermann and Roenneberg, 2009; Kantermann et al., 2012).
et al., 2010; Knutsson, 2004; Wang et al., 2011). To date, only a few shift-work studies have
Therefore, there is a need to understand the implemented the concept of internal time into
mechanisms underlying shift work associated their analyzes. Complete adjustment of the circa-
health deteriorations to design better shift-work dian clock to shift work is rare, even in permanent
conditions. night workers in conventional shift-work settings
The complexity of the issue extends to under- (Folkard, 2008). Circadian adjustment supporting
standing disruptions to the internal circadian work environments have been identified only in
401

highly specialized settings such as offshore oilrigs morningness, the more satisfied the workers were
(Barnes et al., 1998; Gibbs et al., 2002) or work- with their job. By contrast, among night workers,
stations in Antarctica (Lund et al., 2001; Midwin- job satisfaction was associated with sleep quality
ter and Arendt, 1991; Ross et al., 1995). As an and hospital seniority rather than diurnal prefer-
alternative, recent shift-work simulation studies ence. Although this study has been conducted
have demonstrated that establishing a compro- with night workers, these findings reinforce the
mise circadian phase position facilitated adjust- idea that individual preferences for evening
ment in people switching between night shifts activities cannot be seen as the sole factor leading
and day work (Lee et al., 2006; Smith and to shift-work adaptation. Currently, however,
Eastman, 2008; Smith et al., 2008, 2009). These there are no conclusive data showing that later
observations may be of interest in the future chronotypes would develop less (chronic) health
design of shift schedules. In addition, the few problems than earlier chronotypes when working
studies that have implemented the concept of in alternating shifts.
internal time into their study design show evi- There are a number of comprehensive review
dence of chronotype-specific differences in shift- articles on shift work and health (Costa, 2003b;
work adjustment: later chronotypes exhibited Kantermann et al., 2010; Knutsson, 2004; Wang
higher shift-work tolerance and less rigid sleep- et al., 2011). Aiming to add to this existing pool
wake behavior compared to earlier chronotypes of knowledge, in this chapter, we present results
(Duffy et al., 1999; Härmä, 1993; Östberg, 1973; from both our field and laboratory studies on shift
Vidacek et al., 1988). Further, later chronotypes workers. These studies emphasize the importance
appeared to accumulate less sleep deficit during of taking individual factors—especially internal
a shift-work cycle (Folkard and Monk, 1981). time—into account. Clearly, the insights gained
These findings were also supported by reports of when we study people under controlled labora-
better sleep quality in later chronotypes trying tory conditions (e.g., by strictly controlling circa-
to sleep during the daytime (Khaleque, 1999). dian phase, posture, diet, light exposure, etc.)
Later chronotypes are therefore assumed to suf- helps to design better field studies and interpret
fer less from night shifts (Burgess et al., 2002; findings gathered in noisy and less controllable
Folkard and Monk, 1981; Hildebrandt and real-life settings. Therefore, we have structured
Stratmann, 1979). By contrast, later chronotypes this chapter into two sections: (1) studies per-
appear to be more challenged by early morning formed in the field, and (2) studies performed
shifts (Åkerstedt and Torsvall, 1981; Griefahn under controlled conditions in the laboratory.
et al., 2002). This latter result has also been con-
firmed in our study discussed in more detail
below (Kantermann et al., 2012a). Another Studying shift workers in real-life settings
important factor in this context is the effect of
age, since younger people have later chronotypes Stress and adverse effects on health in shift
and chronotype, in turn, advances with age workers has been related to adverse working
(Roenneberg et al., 2007a). These findings have conditions, poor adaptation of work and social
been used to explain why older, earlier life, sleep deprivation and interference in the syn-
chronotypes, on average, withdraw sooner from chronization of the internal circadian clock by the
shift work than younger, later chronotypes (Bohle light–dark cycle (Costa, 2003a; Kantermann et al.,
and Tilley, 1989; Costa et al., 1989; Hauke et al., 2010; Knutsson, 2003; Thompson, 2009; Wang
1979). In addition, a study (Moreno et al., 2012) et al., 2011). Therefore, we have performed a
involving 514 nursing professionals, found among number of shift-work field studies, to investigate
the day workers (DW) that the higher the acute stress (measured via levels of morning and
402

evening cortisol), as well as chronic stress ratio (WHR) and elevated VLDL-cholesterol.
(measured via arterial stiffness as a surrogate for In addition, the “irregular” shift workers reported
atherosclerotic risk), and consequences of shift- more tiredness after work, more disturbances due
work demand as estimated by the incidence rate to truck vibration, and less job demand compared
at work. to DW. In addition, the “irregular” shift workers
were employed as truck drivers for longer than
the DW. Cortisol collected both in the morning
Cortisol awakening response and stress in shift and at bedtime in the “irregular” shift workers
working truck drivers was positively correlated to short sleep duration,
low job satisfaction, total cholesterol, HDL,
In a cross-sectional study (Ulhôa et al., 2011), we LDL, VLDL, and triglycerides. DW showed
investigated the individual cortisol awakening higher cortisol at 30 min after waking and a
response (CAR) (Dahlgren et al., 2009) in truck higher CAR on their workdays compared to their
drivers working day shifts (n ¼ 21) or “irregular days off. The “irregular” shift workers had higher
shifts” including night work (n ¼ 21). In addition, cortisol levels on their days off compared to the
we investigated subjective and objective sleep DW. These results suggest that the “irregular”
and cardiovascular blood parameters in these shift workers might have had a prolonged stress
workers. All participants worked for the same response on their days off, since they had higher
transportation company in Brazil. Objective sleep cortisol levels on these days compared to their
was obtained from actigraphy measurements. Sal- workdays (Fig. 1). An earlier study of 470 truck
ivary cortisol samples for the CAR analysis were drivers reported extended working hours as a
obtained at two time points: at waking time and stressor at work, and this was associated with
at 30 min after waking time. In addition, cortisol minor psychiatric disorders (Ulhôa et al., 2010).
at bedtime was also measured. These three corti- Regulation of working hours and avoidance of
sol samples were collected both during a workday extended work hours may thus be an important
and on a day off work. The “irregular” shift way to reduce stress in truck drivers. Future stud-
workers showed a significantly higher waist-hip ies of shift workers with irregular work hours are

1.25
Salivary cortisol level (µg/dl)

1.00

0.75

0.50

0.25

0.00
Upon waking 30 min after waking At bedtime
Time of sample collection

Fig. 1. Salivary cortisol levels (mean  SEM) according to the time of sample collection in irregular shift workers. Samples were
collected upon awaking, 30 min after waking and at bedtime during a work day (■) and during a day off (□).
403

warranted for a more detailed investigation of addition to blood pressure (BP) and heart rate
stress responses. (HR) on one morning shift between 08:00 and
12:30 h (no caffeine/smoking/exercise). There
were no differences in chronotype, age, body mass
Individual “shift-work load” in rotating shift index (BMI), WHR, BP, HR, smoking, and coffee
workers consumption between shift workers and DW. By
contrast to DW, however, shift workers (CW and
Regarding the effects of shift work on the individ- CC combined) reported more stomach and diges-
ual, the different shift types have become subject tion problems, and more weight fluctuations. In
to investigation. For example, clockwise (CW)- addition, as expected, shift workers had more
rotating schedules (e.g., rotating from the morning SJL. In previous investigations, SJL has been
shift to the late shift to the night shift) compared to found to vary with chronotype and work shift
counter-clockwise (CC) rotation (e.g., rotating (Juda, 2010; Vetter, 2010). This finding was also
from the night shift to the late shift to the morning corroborated in our study. The highest level of
shift) is considered the shift direction of choice as it SJL was found for early chronotypes while doing
is assumed that CW rotation causes less sleep and night shifts, with only a small amount of SJL
circadian rhythm disruption (Costa, 2003a; Härmä, observed for early chronotypes during morning
1993; Knauth, 1993). Conclusive evidence, how- and late shifts. Late chronotypes, however, showed
ever, in particular on chronic health parameters, the highest amount of SJL on the morning shift,
is limited. For the first time in a field study of rotat- although this was less than observed in the early
ing shift workers, we have measured pulse wave chronotypes on night shifts. The two shift-work
velocity (PWV) in fast CW and slow CC shift groups (CW, CC) did not differ in the ratings of
workers compared to DW (Kantermann et al., how shift work affected their sleep, social, and
2012a). The aim of this study was to assess cardio- work life. In all workers combined (CW, CC, and
vascular risk using arterial stiffness in workers DW), HR and average SJL were significantly posi-
undertaking different shift rotations. Male workers tively associated (r ¼ 0.309, p ¼ 0.021 adjusted for
(n ¼ 77, mean ( SD) age 42  7.6 yrs) in a Belgian age). This association also was significant for DW
steel factory with at least 5 years experience in their alone (r ¼ 0.679, p ¼ 0.015 adjusted for age). There
current work schedule participated. All par- was no statistically significant difference in PWV
ticipants completed questionnaires covering between fast CW, slow CC, and DW.
demographics, details about their shift work sched- In a subsequent step, and to introduce a new
ule, health, and stimulant consumption. All shift and simple concept to shift-work research, we cal-
and day workers also completed the Munich culated the “individual shift-work load” (ISL), to
Chronotype Questionnaire for shift workers measure the impact of a worker’s schedule, using
(MCTQshift) (Juda, 2010). Data obtained from the the following two formulae:
MCTQshift were used to calculate each worker’s
chronotype (time of mid-sleep on free days (i) For shift workers:
(MSFsc), corrected for the sleep deficit on ISLshift ¼ (SJL/#WD)*#SWyears
workdays) and the amount of their social jetlag (ii) For DW: ISLday ¼ (SJL/#WD)*#Wyears
(SJL) (difference between the time of mid-sleep
on workdays (MSW) and MSF) (Wittmann et al., The ISL is composed of a worker’s individual
2006) in each work shift (early-, late-, night-, and SJL (difference between time of MSW and
day shift). Shift workers also self-rated how sleep, MSF; Wittmann et al., 2006) divided by the
social, and work life were affected due to working total number of days per shift cycle (#WD;
shifts. In 63 workers, we measured PWV, in representing speed of rotation) and multiplied
404

by the individual’s total number of years of expo- Franzen et al., 2008; Sallinen et al., 2008; Van
sure to shift work (#SWyears). As the DW were Dongen et al., 2003). We investigated the incidence
not exposed to shift work, the respective number rate in rotating shift workers (undertaking
of years of employment at the current company morning, late, and night shifts) employed in two
(#Wyears) was used instead. There was a signifi- different shift-work rotations (Kantermann et al.,
cant positive correlation (r ¼ 0.493, p ¼ 0.005) 2012b). A retrospective analysis of the incidence
between ISL and PWV (PWVaBP, which is PWV data from 730 male shift workers employed in
adjusted for age and BP). This finding provides either a CW (e.g., rotating from the morning shift
first evidence that arterial stiffness assessed by to the late shift to the night shift) or CC rotation
PWV and the chronic strain of shift work (e.g., rotating from the night shift to the late shift
(estimated from measures of SJL) may be interre- to the morning shift) with comparable work
lated. The group differences in both speed and conditions at the same steel factory over a 5-year
direction of shift rotation, the small number of period has been performed. Morning shifts
subjects, and its cross-sectional design are exhibited a significantly higher incidence rate com-
limitations of this study. We hope, however, that pared to night shifts, independent of shift-work
the findings from this pilot study will initiate rotation. The incidence rate across the 24-h day
future studies of this kind. Indeed, more studies did not differ between CW and CC rotation. The
in workers employed in shift schedules with dif- elevated incidence rate in the morning shift at this
ferent speeds and directions of rotation but work- steel factory could be related to the morning shift
ing in the same work setting are needed. being the most labor-intensive one of the three
work shifts in both the CW and CC rotation. In
addition, there could be an effect of sleep depriva-
Effect of shift type on incidence rate in rotating tion due to the early start of the morning shift at
shift workers 06:00 h, which means an even earlier get up time
for the workers. These findings suggest that, in
Previous studies have shown that the night shift addition to, or even irrespective of, work shift and
in rotating shift schedules is characterized by the direction of shift rotation, the impact of work time
lowest levels of alertness and vigilance, and the and work demand may have a strong modulating
highest amount of sleep deprivation and hence has effect on the incidence rate.
been suggested to lead to an elevated incidence risk
in night workers (Costa, 2003b; Folkard, 1997;
Mitler et al., 1988). Studies have shown influences Studying shift workers in controlled laboratory
of, inter alia, work duration (length of time of being settings
at work), start time of a work shift, number of con-
secutive work shifts, type of occupation, and prior Advantages of laboratory studies
sleep duration (Folkard et al., 2005), even when
the work tasks were comparable across shifts What is the benefit of performing studies in shift
(Folkard, 1997). Such findings are supported by workers under strictly controlled laboratory
controlled circadian laboratory studies on alertness, conditions? To answer that question, one needs
vigilance, sleepiness, and fatigue, which show a to consider all the confounding factors in real life,
clear 24-h variation in these parameters (Åkerstedt, which are very difficult to control in field studies.
2007; Aschoff, 1965; Cajochen et al., 1999; Dijk Such confounders encompass, inter alia interindi-
et al., 1992; Graw et al., 2004) and deterioration in vidual and intraindividual differences in shift-work
these parameters after sleep deprivation (Axelsson history (e.g., work schedule, work tasks, number of
et al., 2008; Dinges et al., 1997; Doran et al., 2001; work hours, etc.), the lighting environment both at
405

and away from work, stress both mental and physi- response to one night of total sleep deprivation
cal (which is often dependent upon job type and (TSD) (as a proxy for the first night of shift work)
work load), diet, sleep duration and sleep depriva- and recovery sleep under controlled laboratory
tion prior to and after a shift, and individual inter- conditions (Wehrens et al., 2010, 2012a, 2012b).
nal time (chronotype). Epidemiological studies, The effect of sleep deprivation on the participants’
for example, often fail to take account of the exact metabolic and cardiovascular function as well as
time (especially relative to internal time) samples alertness, mood, and performance was assessed.
of biomarkers have been collected in relation to a Both groups followed a 7-day regular sleep-wake
shift worker’s actual work hours (Boggild and cycle prior to the laboratory session. This was done
Knutsson, 1999; Sookoian et al., 2007), whereas to ensure that shift-workers were either on day
other studies have carried out such a control shifts or days off in order to avoid sleep depriva-
(Esquirol et al., 2009). The lack of control of con- tion and circadian misalignment in the study
founders is not merely a matter of the number of subjects. The in-laboratory study protocol con-
study subjects, since there are also smaller field sisted of one night of adaptation sleep, followed
studies with fewer subjects that have failed to by one night of baseline sleep, and one night of
address one or more of these important points TSD (30.5-h wakefulness). A subsequent daytime
mentioned above (Adams et al., 1998; Amir et al., nap opportunity of 4 h and a full recovery sleep
2004; Rauchenzauner et al., 2009). Such con- period were provided. All measurements were
founding differences must be taken into account performed relative to wake up time and controlled
else substantial information that could explain for body posture, diet, and light exposure. Post hoc
group differences may be missed, for example, measurement of dim light salivary melatonin onset
when studying groups employed in different shift- prior to the baseline night confirmed that there
work rotations or when comparing shift workers were no significant differences in circadian phase
with non-shift-workers. To disentangle the differ- between shift-workers and non-shift-workers.
ent causes that may contribute to the effects of shift Compared to the non-shift-workers, however, shift
work on health and wellbeing, laboratory studies workers showed a lower HR variability (HRV)
are able to control many factors and hence are able variance and higher sympathetic activity, as well
to manipulate single, specific variables at pre- as a trend towards lower endothelial function
defined time points. Shift-work simulation studies (assessed by flow-mediated dilatation of the bra-
(Hampton et al., 1996; Ribeiro et al., 1998), for chial artery) throughout the study. In addition,
example, were able to demonstrate that glucose, after the recovery sleep period, the postprandial
insulin, and lipid levels following food intake were insulin response was significantly increased and
dependent on a person’s internal time. In addition, the non-esterified fatty acid (NEFA) response
so-called forced desynchrony experiments showed reduced compared to after sleep deprivation
that circadian misalignment negatively affects and/or baseline sleep in the non-shift workers
metabolism (Scheer et al., 2009). To our knowl- only. In this study, the shift workers also felt more
edge, there is only one study that has investigated alert, more cheerful, more elated and were
real shift workers under laboratory conditions calmer throughout the protocol compared to
(Simon et al., 2000), more commonly laboratory the non-shift workers (Wehrens et al., 2012a).
studies investigate shift work naïve subjects that In addition, shift workers showed a faster
are mostly healthy and young. median reaction time (RT) compared to non-shift
Therefore, we have compared experienced shift workers although four other psychovigilance
workers (n ¼ 11, with at least 5 years of shift-work test parameters did not differ significantly bet-
experience) to matched (for age, BMI, and cho- ween the two groups. These findings suggest
lesterol) non-shift-workers (n ¼ 14) in their that experienced shift workers cope better with
406

laboratory sleep deprivation than non-shift This latter aspect prevents detecting small
workers. These group differences could be differences in study parameters, especially as many
explained inter alia by a selection bias into and biological parameters show large intra- and interin-
out of shift work, the absence of actual shift work dividual variation. In addition, as laboratory stud-
that was inherent to the laboratory study and ies are obviously not performed under “real-life”
regular scheduled sleep prior to the laboratory conditions, information on many aspects of shift
study. The data on HRV and endothelial function work is not recorded, for example, the social and
(Wehrens et al., 2012b), however, suggest economic reasons for choosing a shift-work job.
increased cardiovascular risk in the shift workers, In addition, shift-work simulation studies often
although this was not significantly affected by use shift-work naïve subjects instead of experi-
sleep deprivation. The observed effects of recov- enced shift workers. Although simulation studies
ery sleep on metabolism in the non-shift workers on non-shift workers can show the acute effects of
suggest a tendency towards a state of insulin resis- a shift-work schedule, they cannot assess the
tance in non-shift workers compared to the shift chronic effects shift work poses on employees. By
workers. This may either be a direct effect of contrast, field studies can achieve this by studying
recovery sleep or a delayed response to sleep the same subjects repeatedly and prospectively at
deprivation. Such responses after recovery sleep regular intervals (presuming no worker leaves the
are similar to the findings of the delayed stress company). However, from controlled laboratory
response on days off in our study on truck drivers investigations, we can determine which parameters
in Brazil (Ulhôa et al., 2011). The results from to control when performing field studies and vice
both these laboratory and field studies emphasize versa—clearly a win-win collaboration.
that shift work schedules should be designed to
allow for optimal physiological recovery. From a
neurobehavioral point of view, shift workers Conclusion
may underestimate how much recovery time they
actually need. A similar mismatch between the With this overview of our laboratory and field
subjective perception of increased sleep depriva- studies on shift workers and non-shift workers
tion and objective measures in performance has we aimed to illustrate the importance of the fol-
been shown previously (Franzen et al., 2008; lowing aspects:
Galliaud et al., 2008; Van Dongen et al., 2003,
2004) and remains an intriguing aspect for future  Studying shift workers and non-shift workers in
shift-work studies. the same work environment is imperative to
best control for interindividual variation and
worksite-related effects (e.g., as we have done
Disadvantages of laboratory studies in a Belgian steel factory study and in the study
of Brazilian truck drivers).
Despite all the advantages listed in the previous  Strictly controlled laboratory studies (as we
section, there is one major limitation to laboratory have presented) on real shift-workers and non-
shift-work studies; only a small number of subjects shift workers are fundamental to understand
can be studied under strictly controlled conditions. the immediate and acute effects that shift-work
Often due to costs for research staff, consumables, and sleep deprivation pose on health, as these
and laboratory space, time constraints or recruit- studies allow for adequate matching of subjects.
ment issues (i.e., the more strict the inclusion  Findings from strictly controlled laboratory stud-
criteria, the more difficult recruitment becomes), ies must be translated into field investigations,
these studies can be statistically underpowered. first to validate these findings and second to
407

The shift-clock-work

Real-life Controlled
field laboratory
studies studies
Healthier
Individual shift
predictors schedules

Data Computer
mining simulations

Fig. 2. The “shift-clock-work” illustrates the interconnectivity between field studies performed in real-life and controlled laboratory
investigations helping to cross-validate data, which in turn drives databases for data mining and computer simulations. The outcome
is “individual predictors” helping to design better and healthier shift-work schedules. Figure modified from Kantermann et al.
(2010).

discover new aspects, which in turn can be factors into account (Fig. 2). Although there remain
translated back into laboratory studies—a clear some profound gaps in our knowledge, for example,
win-win circulation of scientific knowledge. how shift work puts one individual but not another
 The introduction of new, noninvasive (bio-) at risk, the existence of good research tools and
markers into shift-work research has the poten- our current knowledge makes this noisy and individ-
tial to elucidate the underlying mechanisms of ual shift-work research doable.
the pathophysiological pathways involved in
shift work. These new (bio-) markers, for exam-
ple, are: chronotype (phase of entrainment), Acknowledgments
SJL, ISL, PWV, HR, HRV, and flow-mediated
dilation. T. K. is supported by the DFG (German Research
Foundation), S. M. T. W. is supported by the
Only by combining laboratory and field studies BBSRC (BB/I008470/1), M. U. and C. M. are
and by comparing these data, will it be possible to supported by Santander, USP, FAPESP, and
identify the individual predictors that constitute Surrey (2011/50169-6), and D. J. S. is a Royal Soci-
important set points in the interaction between shift ety Wolfson Research Merit Award holder. The
work and health status. Further, although not research was supported by an EU Marie Curie
explicitly discussed in this chapter, these data can Research Training Network grant (CT-2004-
then be subjected to shift-work models. Such 512362), 6th Framework project EUCLOCK
models will help, not only to rapidly generate new (018741) and Stockgrand Ltd., University of
hypotheses for future studies (both in the laboratory Surrey, Guildford, UK.
and in the field), but also to develop better and We especially thank our participants at
healthier shift-work schedules that take individual ArcelorMittal Industeel Belgium in Charleroi,
408

the truck drivers in São Paulo, Brazil, and the of the National Academy of Sciences of the United States of
participants in our laboratory shift-work study. America, 105(5), 1602–1607.
Bruni, O., Russo, P. M., Ferri, R., Novelli, L., Galli, F., &
We also thank Shelagh Hampton, Françoise Guidetti, V. (2008). Relationships between headache and
Duboutay, Myriam Kerkhofs, Damien Haubruge, sleep in a non-clinical population of children and
and Arno Schmidt-Trucksäss for their valuable adolescents. Sleep Medicine, 9(5), 542–548.
help with the studies. Without their contribution, Burgess, H. J., Sharkey, K. M., & Eastman, C. I. (2002). Bright
these studies would not have been made possible. light, dark and melatonin can promote circadian adaptation
in night shift workers. Sleep Medicine Reviews, 6(5),
Conflict of interest: None to declare. 407–420.
Cajochen, C., Khalsa, S. B., Wyatt, J. K., Czeisler, C. A., &
Dijk, D. J. (1999). EEG and ocular correlates of circadian
References melatonin phase and human performance decrements dur-
ing sleep loss. American Journal of Physiology, 277(3 Pt
Adams, S. L., Roxe, D. M., Weiss, J., Zhang, F., & 2), R640–R649.
Rosenthal, J. E. (1998). Ambulatory blood pressure and Costa, G. (2003a). Factors influencing health of workers and
Holter monitoring of emergency physicians before, during, tolerance to shift work. Theoretical Issues in Ergonomics
and after a night shift. Academic Emergency Medicine, 5(9), Science, 4(3–4), 263–288.
871–877. Costa, G. (2003b). Shift work and occupational medicine: An
Åkerstedt, T. (2007). Altered sleep/wake patterns and mental overview. Occupational Medicine (London), 53(2), 83–88.
performance. Physiology and Behavior, 90(2–3), 209–218. Costa, G., Lievore, F., Casaletti, G., Gaffuri, E., & Folkard, S.
Åkerstedt, T., Kecklund, G., & Gillberg, M. (2007). Sleep and (1989). Circadian characteristics influencing interindividual
sleepiness in relation to stress and displaced work hours. differences in tolerance and adjustment to shiftwork. Ergo-
Physiology and Behavior, 92(1–2), 250–255. nomics, 32(4), 373–385.
Åkerstedt, T., & Torsvall, L. (1981). Shift work. Shift- Dahlgren, A., Kecklund, G., Theorell, T., & Åkerstedt, T.
dependent well-being and individual differences. Ergonom- (2009). Day-to-day variation in saliva cortisol—Relation
ics, 24(4), 265–273. with sleep, stress and self-rated health. Biological Psychol-
Åkerstedt, T., & Wright, K. P. Jr. (2009). Sleep loss and ogy, 82(2), 149–155.
fatigue in shift work and shift work disorder. Sleep Medicine Dijk, D. J., Duffy, J. F., & Czeisler, C. A. (1992). Circadian
Clinics, 4(2), 257–271. and sleep/wake dependent aspects of subjective alertness
Amir, O., Alroy, S., Schliamser, J. E., Asmir, I., Shiran, A., and cognitive performance. Journal of Sleep Research,
Flugelman, M. Y., et al. (2004). Brachial artery endothelial 1(2), 112–117.
function in residents and fellows working night shifts. The Dinges, D. F., Pack, F., Williams, K., Gillen, K. A.,
American Journal of Cardiology, 93(7), 947–949. Powell, J. W., Ott, G. E., et al. (1997). Cumulative sleepi-
Aschoff, J. (1965). Circadian rhythms in man. Science, 148, ness, mood disturbance, and psychomotor vigilance perfor-
1427–1432. mance decrements during a week of sleep restricted to 4-5
Axelsson, J., Kecklund, G., Åkerstedt, T., Donofrio, P., hours per night. Sleep, 20(4), 267–277.
Lekander, M., & Ingre, M. (2008). Sleepiness and perfor- Doran, S. M., Van Dongen, H. P., & Dinges, D. F. (2001).
mance in response to repeated sleep restriction and Sustained attention performance during sleep deprivation:
subsequent recovery during semi-laboratory conditions. Evidence of state instability. Archives Italiennes de Biologie,
Chronobiology International, 25(2), 297–308. 139(3), 253–267.
Barnes, R. G., Deacon, S. J., Forbes, M. J., & Arendt, J. Duffy, J. F., Dijk, D. J., Hall, E. F., & Czeisler, C. A. (1999).
(1998). Adaptation of the 6-sulphatoxymelatonin rhythm in Relationship of endogenous circadian melatonin and tem-
shiftworkers on offshore oil installations during a 2-week perature rhythms to self-reported preference for morning
12-h night shift. Neuroscience Letters, 241(1), 9–12. or evening activity in young and older people. Journal of
Boggild, H., & Knutsson, A. (1999). Shift work, risk factors Investigative Medicine, 47(3), 141–150.
and cardiovascular disease. Scandinavian Journal of Work, Erren, T. C., Pape, H. G., Reiter, R. J., & Piekarski, C. (2008).
Environment and Health, 25(2), 85–99. Chronodisruption and cancer. Die Naturwissenschaften, 95
Bohle, P., & Tilley, A. J. (1989). The impact of night work on (5), 367–382.
psychological well-being. Ergonomics, 32(9), 1089–1099. Esquirol, Y., Bongard, V., Mabile, L., Jonnier, B., Soulat, J.-M.,
Brown, S. A., Kunz, D., Dumas, A., Westermark, P. O., & Perret, B. (2009). Shift work and metabolic syndrome:
Vanselow, K., Tilmann-Wahnschaffe, A., et al. (2008). Respective impacts of job strain, physical activity, and dietary
Molecular insights into human daily behavior. Proceedings rhythms. Chronobiology International, 26(3), 544–559.
409

Folkard, S. (1997). Black times: Temporal determinants of Hublin, C., Partinen, M., Koskenvuo, K., Silventoinen, K.,
transport safety. Accident Analysis and Prevention, 29(4), Koskenvuo, M., & Kaprio, J. (2010). Shift-work and cardio-
417–430. vascular disease: A population-based 22-year follow-up
Folkard, S. (2008). Do permanent night workers show circadian study. European Journal of Epidemiology, 25(5), 315–323.
adjustment? A review based on the endogenous melatonin Juda, M. (2010). The importance of chronotype in shift work.
rhythm. Chronobiology International, 25(2), 215–224. Department of Psychology. Munich, Ludwigs-Maximilians-
Folkard, S., Lombardi, D. A., & Tucker, P. T. (2005). Universität, LMU.
Shiftwork: Safety, sleepiness and sleep. Industrial Health, Kantermann, T. (2008). Challenging the human internal clock by
43(1), 20–23. Daylight Saving Time and Shift-Work. Institute for Medical
Folkard, S., & Monk, T. H. (1981). Individual differences in Psychology. Munich, Ludwigs-Maximilians-Universität, LMU.
the circadian response to a weekly rotating shift system. In Kantermann, T., Duboutay, F., Haubruge, D., Kerkhofs, M.,
A. Reinberg, N. Vieux & P. Andlauer (Eds.), Night and shift Schmidt-Trucksäss, A., & Skene, D. J. (2012a). Atheroscle-
work: Biological and social aspects (pp. 367–374). Oxford: rotic risk and social jetlag in rotating shift-workers: First evi-
Pergamon Press. dence from a pilot study. WORK: A Journal of Prevention,
Franzen, P. L., Siegle, G. J., & Buysse, D. J. (2008). Assessment, & Rehabilitation, in press.
Relationships between affect, vigilance, and sleepiness fol- Kantermann, T., Haubruge, D., & Skene, D. J. (2012b). The shift-
lowing sleep deprivation. Journal of Sleep Research, 17(1), work accident rate is more related to the shift type than to shift
34–41. rotation. Human and Ecological Risk Assessment, in press.
Freedman, M. S., Lucas, R. J., Soni, B., Von Schantz, M., Kantermann, T., Juda, M., Merrow, M., & Roenneberg, T.
Muñoz, M., David-Gray, Z., et al. (1999). Regulation of (2007). The human circadian clock’s seasonal adjustment is
mammalian circadian behavior by non-rod, non-cone, ocular disrupted by daylight saving time. Current Biology, 17(22),
photoreceptors. Science, 284(5413), 502–504. 1996–2000.
Galliaud, E., Taillard, J., Sagaspe, P., Valtat, C., Bioulac, B., & Kantermann, T., Juda, M., Vetter, C., & Roenneberg, T.
Philip, P. (2008). Sharp and sleepy: Evidence for dissocia- (2010). Shift-work research—Where do we stand, where
tion between sleep pressure and nocturnal performance. should we go? Sleep and Biological Rhythms, 8(2), 95–105.
Journal of Sleep Research, 17(1), 11–15. Kantermann, T., & Roenneberg, T. (2009). Is light-at-night a
Gibbs, M., Hampton, S., Morgan, L., & Arendt, J. (2002). health risk factor or a health risk predictor? Chronobiology
Adaptation of the circadian rhythm of 6-sulphatoxymelatonin International, 26(6), 1069–1074.
to a shift schedule of seven nights followed by seven days in Kantermann, T., Theadom, A., Roenneberg, T., &
offshore oil installation workers. Neuroscience Letters, 325 Cropley, M. (2012). Fibromyalgia syndrome and chronotype:
(2), 91–94. Late chronotypes are more affected. Journal of Biological
Graw, P., Krauchi, K., Knoblauch, V., Wirz-Justice, A., & Rhythms, 27, 176–179.
Cajochen, C. (2004). Circadian and wake-dependent modu- Khaleque, A. (1999). Sleep deficiency and quality of life of
lation of fastest and slowest reaction times during the psy- shift workers. Social Indicators Research, 46, 181–189.
chomotor vigilance task. Physiology and Behavior, 80(5), Knauth, P. (1993). The design of shift systems. Ergonomics, 36
695–701. (1–3), 15–28.
Griefahn, B., Kunemund, C., Golka, K., Thier, R., & Knutsson, A. (2003). Health disorders of shift workers. Occu-
Degen, G. (2002). Melatonin synthesis: A possible indicator pational Medicine (London), 53(2), 103–108.
of intolerance to shiftwork. American Journal of Industrial Knutsson, A. (2004). Methodological aspects of shift-work
Medicine, 42(5), 427–436. research. Chronobiology International, 21(6), 1037–1047.
Hampton, S. M., Morgan, L. M., Lawrence, N., Lee, C., Smith, M. R., & Eastman, C. I. (2006). A compromise
Anastasiadou, T., Norris, F., Deacon, S., et al. (1996). Post- phase position for permanent night shift workers: Circadian
prandial hormone and metabolic responses in simulated phase after two night shifts with scheduled sleep and light/
shift work. Journal of Endocrinology, 151(2), 259–267. dark exposure. Chronobiology International, 23(4), 859–875.
Härmä, M. (1993). Individual differences in tolerance to Lemos, L. C., Marqueze, E. C., Sachi, F., Lorenzi-Filho, G., &
shiftwork: A review. Ergonomics, 36(1–3), 101–109. Moreno, C. R. (2009). Obstructive sleep apnea syndrome in
Hauke, P., Kittler, H., & Moog, R. (1979). Interindividual truck drivers. Jornal Brasileiro de Pneumologia, 35(6),
differences in tolerance to shift-work related to 500–506.
morningness—Eveningness. Chronobiologia, 6(2), 109. Levandovski, R., Dantas, G., Fernandes, L. C., Caumo, W.,
Hildebrandt, G., & Stratmann, I. (1979). Circadian system Torres, I., Roenneberg, T., et al. (2011). Depression
response to nightwork in relation to the circadian phase scores associate with chronotype and social jetlag in a rural
position. International Archives of Occupational and Envi- population. Chronobiology International, doi:10.3109/
ronmental Health, 43, 73–83. 07420528.2011.602445(epub September 6).
410

Lund, J., Arendt, J., Hampton, S. M., English, J., & Roenneberg, T., Kumar, C. J., & Merrow, M. (2007b). The
Morgan, L. M. (2001). Postprandial hormone and metabolic human circadian clock entrains to sun time. Current Biology,
responses amongst shift workers in Antarctica. Journal of 17, R44–R45.
Endocrinology, 171(3), 557–564. Ross, J. K., Arendt, J., Horne, J., & Haston, W. (1995). Night-
Midwinter, M. J., & Arendt, J. (1991). Adaptation of the mel- shift work in Antarctica: Sleep characteristics and bright
atonin rhythm in human subjects following night-shift work light treatment. Physiology and Behavior, 57(6), 1169–1174.
in Antarctica. Neuroscience Letters, 122(2), 195–198. Saksvik, I. B., Bjorvatn, B., Hetland, H., Sandal, G. M., &
Mitler, M. M., Carskadon, M. A., Czeisler, C. A., Pallesen, S. (2011). Individual differences in tolerance to
Dement, W. C., Dinges, D. F., & Graeber, R. C. (1988). shift work—A systematic review. Sleep Medicine Reviews,
Catastrophes, sleep, and public policy: Consensus report. 15(4), 221–235.
Sleep, 11(1), 100–109. Sallinen, M., Holm, A., Hiltunen, J., Hirvonen, K., Härmä, M.,
Moreno, C. R., Carvalho, F. A., Lorenzi, C., Matuzaki, L. S., Koskelo, J., et al. (2008). Recovery of cognitive perfor-
Prezotti, S., Bighetti, P., et al. (2004). High risk for obstruc- mance from sleep debt: Do a short rest pause and a single
tive sleep apnea in truck drivers estimated by the Berlin recovery night help? Chronobiology International, 25(2–3),
questionnaire: Prevalence and associated factors. Chronobi- 279–296.
ology International, 21(6), 871–879. Scheer, F. A., Hilton, M. F., Mantzoros, C. S., & Shea, S. A.
Moreno, C. R. C., Marqueze, E. C., Lemos, L. C., Soares, N., (2009). Adverse metabolic and cardiovascular consequences
& Lorenzi-Filho, G. (2012). Job satisfaction and dis- of circadian misalignment. Proceedings of the National Acad-
crepancies between social and biological timing. Biological emy of Sciences of the United States of America, 106(11),
Rhythm Research, 43(1), 73–80. 4453–4458.
Natale, V., Adan, A., & Scapellato, P. (2005). Are seasonality Simon, C., Weibel, L., & Brandenberger, G. (2000). Twenty-
of mood and eveningness closely associated? Psychiatry four-hour rhythms of plasma glucose and insulin secretion
Research, 136(1), 51–60. rate in regular night workers. American Journal of Physiol-
Östberg, O. (1973). Interindividual differences in circadian ogy, Endocrinology and Metabolism, 278(3), E413–E420.
fatigue patterns of shift workers. British Journal of Industrial Smith, M. R., Cullnan, E. E., & Eastman, C. I. (2008). Shaping
Medicine, 30(4), 341–351. the light/dark pattern for circadian adaptation to night shift
Panda, S., Sato, T. K., Castrucc, A. M., Rollag, M. D., work. Physiology and Behavior, 95, 449–456.
Degrip, W. J., Hogenesch, J. B., et al. (2002). Melanopsin Smith, M. R., & Eastman, C. I. (2008). Night shift performance
(Opn4) requirement for normal light-induced circadian is improved by a compromise circadian phase position:
phase shifting. Science, 298, 2213–2216. Study 3. Circadian phase after 7 night shifts with an
Puttonen, S., Härmä, M., & Hublin, C. (2010). Shift work and intervening weekend off. Sleep, 31(12), 1639–1645.
cardiovascular disease—Pathways from circadian stress to Smith, M. R., Fogg, L. F., & Eastman, C. I. (2009). Practical
morbidity. Scandinavian Journal of Work, Environment interventions to promote circadian adaptation to permanent
and Health, 36(2), 96–108. night shift work: Study 4. Journal of Biological Rhythms,
Rauchenzauner, M., Ernst, F., Hintringer, F., Ulmer, H., 24(2), 161–172.
Ebenbichler, C. F., Kasseroler, M. T., et al. (2009). Sookoian, S., Gemma, C., Gianotti, T. F., Burgueno, A.,
Arrhythmias and increased neuro-endocrine stress response Alvarez, A., Gonzalez, C. D., et al. (2007). Effects of rotat-
during physicians’ night shifts: A randomized cross-over ing shift work on biomarkers of metabolic syndrome and
trial. European Heart Journal, 30(21), 2606–2613. inflammation. Journal of Internal Medicine, 261(3), 285–292.
Ribeiro, D. C., Hampton, S. M., Morgan, L., Deacon, S., & Stevens, R. G. (2009). Light-at-night, circadian disruption and
Arendt, J. (1998). Altered postprandial hormone and breast cancer: Assessment of existing evidence. International
metabolic responses in a simulated shift work environment. Journal of Epidemiology, 38(4), 963–970.
Journal of Endocrinology, 158(3), 305–310. Thompson, E. (2009). Understanding how night work
Roenneberg, T., Daan, S., & Merrow, M. (2003). The art influences the everyday family lives of nurses, their husbands
of entrainment. Journal of Biological Rhythms, 18(3), and children, University of Surrey.
183–194. Ulhôa, M., Marqueze, E., Kantermann, T., Skene, D., &
Roenneberg, T., Hut, R., Daan, S., & Merrow, M. (2010). Moreno, C. (2011). When does stress end? Evidence of a
Entrainment concepts revisited. Journal of Biological prolonged stress reaction in shift-working truck drivers.
Rhythms, 25(5), 329–339. Chronobiology International, 28(9), 810–818.
Roenneberg, T., Kuehnle, T., Juda, M., Kantermann, T., Ulhôa, M. A., Marqueze, E. C., Lemos, L. C., Silva, L. G.,
Allebrandt, K., Gordijn, M., et al. (2007a). Epidemiology Silva, A. A., Nehme, P., et al. (2010). Minor psychiatric dis-
of the human circadian clock. Sleep Medicine Reviews, 11, orders and working conditions in truck drivers. Revista de
429–438. Saúde Pública, 44(6), 1130–1136.
411

Van Dongen, H. P., Maislin, G., & Dinges, D. F. (2004). Deal- Wehrens, S. M., Hampton, S. M., Finn, R. E., & Skene, D. J.
ing with inter-individual differences in the temporal dynam- (2010). Effect of total sleep deprivation on postprandial
ics of fatigue and performance: Importance and techniques. metabolic and insulin responses in shift workers and non-
Aviation, Space, and Environmental Medicine, 75(Suppl. 3), shift workers. Journal of Endocrinology, 206(2), 205–215.
A147–A154. Wehrens, S. M. T., Hampton, S. M., Kerkhofs, M., & Skene,
Van Dongen, H. P., Maislin, G., Mullington, J. M., & D. J., (2012a). Mood, alertness and performance in response
Dinges, D. F. (2003). The cumulative cost of additional to sleep deprivation and recovery sleep in experienced
wakefulness: Dose-response effects on neurobehavioral shiftworkers versus non-shiftworkers. Chronobiology Inter-
functions and sleep physiology from chronic sleep restriction national, 29(5), 537–548.
and total sleep deprivation. Sleep, 26(2), 117–126. Wehrens, S. M., Hampton, S. M., & Skene, D. J. (2012b). Heart
Vetter, C. (2010). Clocks in action—Exploring the impact of rate variability and endothelial function after sleep depriva-
internal time in real life. Faculty of Psychology and Educa- tion and recovery sleep among male shift and non-shift
tional Sciences. Munich, Ludwigs-Maximilians-Universität, workers. Scandinavian Journal of Work, Environment and
LMU. Health, 38(2), 171–181.
Vidacek, S., Kaliterna, L., Radosevic-Vidacek, B., & Wittmann, M., Dinich, J., Merrow, M., & Roenneberg, T.
Folkard, S. (1988). Personality differences in the phase of (2006). Social jetlag: Misalignment of biological and social
circadian rhythms: A comparison of morningness and extra- time. Chronobiology International, 23(1–2), 497–509.
version. Ergonomics, 31(6), 873–888. Wood, J., Birmaher, B., Axelson, D., Ehmann, M.,
Wang, X. S., Armstrong, M. E., Cairns, B. J., Key, T. J., & Kalas, C., Monk, K., et al. (2009). Replicable differences
Travis, R. C. (2011). Shift work and chronic disease: The in preferred circadian phase between bipolar disorder
epidemiological evidence. Occupational Medicine (London), patients and control individuals. Psychiatry Research, 166
61(2), 78–89. (2–3), 201–209.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 23

The evolutionary physiology of


photoperiodism in vertebrates

David Hazlerigg*

Institute of Biological & Environmental Sciences, University of Aberdeen, Aberdeen, Scotland, United Kingdom

Abstract: The capacity to measure day length (photoperiod) is a trait subject to intense evolutionary
pressure, and the circadian system has become an important part of the photoperiodic machinery. With
the exception of mammals, vertebrates possess multiple sites of photosensitivity within the central
nervous system through which light responses may be coordinated. Of these, deep brain photoreceptors
play a special role in photoperiodism in nonmammalian vertebrates, independent of either retinal or
pineal pathways. In mammals, the pineal hormone, melatonin has assumed a function orthologous to
that of deep brain photoreceptors in other vertebrates. Contrasting with this dichotomy in photoperiodic
input pathways, downstream signal processing to produce switch-like seasonal responses is largely
conserved and appears to center on photoperiodic control of thyroid hormone bioavailability within the
hypothalamus. Recent studies implicate the clock-controlled gene eyes absent 3 in linking the circadian
clock to photoperiodic responses, dictating the seasonal level of expression of thyroid-stimulating
hormone by a specialized population of photoperiodically sensitive cells in the pituitary stalk which in
turn govern thyroid hormone metabolism in the basal hypothalamus. This ancestral control pathway
appears to have originated prior to the divergence of the mammalian and avian vertebrate lineages.

Keywords: photoperiodism; melatonin; thyroid-stimulating hormone; TSH; deep brain photoreceptor;


circadian; pars tuberalis; deiodinase; eyes absent.

Energy budgets, life history strategies,


and seasonality Achieving breeding success depends on securing suf-
ficient energetic resources, through nutrient intake,
The laws of natural selection dictate that species sur- to reach reproductive maturity. This summarizes
vival depends on individual success in reproduction. the essential focus on bioenergetics as a fundamental
issue in evolutionary biology.
*Corresponding author.
In managing the energy equation, a key prob-
Tel.: (44) 1224 272871; Fax: (44) 1224 272396 lem is the variable nature of the environment,
E-mail: d.hazlerigg@abdn.ac.uk with periods that are energetically favorable

http://dx.doi.org/10.1016/B978-0-444-59427-3.00023-X 413
414

interspersed with unfavorable episodes (drought, fundamental conservation can be found in elements
famine, winter). Furthermore, evolution leads spe- involved in photoperiodic time measurement. This
cies down avenues in which adaptive preparations review considers this question from the perspective
for breeding success (e.g., growth, gametogenesis, of seasonal timing in vertebrates focusing on recent
development of reproductive ornaments, and other developments in our understanding of neuroendo-
secondary sexual characteristics) demand periods crine control of seasonal reproduction.
of energetic favorability preceding the reproduc-
tive act itself. Offspring survival also depends crit-
ically on the environment, providing a further Thyroid hormone signaling as an ancestral
temporal constraint on the timing of reproduction. pathway controlling reproductive activation
Hence, we can see that the phasing of reproductive
activity to windows of energetic favorability is a It has been known since the 1930s (Benoit, 1936;
basic evolutionary challenge for life on Earth. Woitkevitsch, 1940) that thyroidectomy interferes
Environmental variability is to a large degree of with the seasonal cycle of reproductive activation
predictable geophysical origin, derived from the in birds, a finding that received renewed attention
Earth’s periodic motion, both in rotation about with the work of Follett and colleagues in the Jap-
its polar axis, giving rise to day and night, and in anese quail (Coturnix japonica) in the 1980s and
orbit around the Sun, giving rise to the seasons. 1990s (reviewed in Dawson et al., 2001), and
It is this latter annual–seasonal cycle that has a which was extended also to consider the role of
major bearing on the energetic considerations per- thyroid signaling in seasonal reproductive control
taining to the reproductive success for many spe- in sheep (e.g., Dahl et al., 1994; Follett and Potts,
cies, and as a result, mechanisms to synchronize 1990). The current picture, reviewed thoroughly
life history phases to the annual–seasonal cycle elsewhere (Hazlerigg and Loudon, 2008; Nakao
have evolved. Except at the Equator, the orbit of et al., 2008b), places light-dependent regulation of
the Earth around the Sun leads to an annual cycle type 2 and type 3 thyroid hormone deiodinases
of changing day length and total daily solar radia- (Dio2 and Dio3, respectively) in the basal hypothal-
tion exposure, and together, these provide the ulti- amus at the core of the seasonal photoperiodic
mate cause of the Earthly seasons and proximate response in birds and mammals (Yoshimura et al.,
cues that organisms use to determine the time of 2003). These enzymes activate and inactivate
year and hence to predict forthcoming energetic thyroxine, respectively, thereby controlling thyroid
conditions. The use of changing day length to syn- hormone-dependent biology in a tissue-specific
chronize life history cycles is a prevalent feature of manner (Lechan and Fekete, 2005; Fig. 1). The path-
life, and this suggests that evolution of photope- way linking light to deiodinase regulation in birds
riodic time measurement is an ancestral feature and mammals shows a high degree of conservation
of life of comparable importance to the evolution and centers on light-dependent regulation of thyro-
of circadian time keeping. tropin (thyroid-stimulating hormone, TSH) expres-
In the circadian realm, there is abundant evidence sion in cells in the pars tuberalis (PT) region of the
for evolutionary conservation of circadian rhythm pituitary stalk (Hanon et al., 2008; Nakao et al.,
generators among eukaryotes both in terms of 2008). TSH-expressing PT cells appear to constitute
organization of the core transcription–translation a distinctive population of thyrotrophic cells, which,
feedback loop and in terms of the molecular in contrast to classical thyrotrophs found in the pars
elements involved in photic synchronization (e.g., distalis (PD), are insensitive to hypothalamic thyro-
the cryptochrome and period genes) (Bell-Pedersen tropin-releasing hormone and to peripheral thyroid
et al., 2005; Dardente and Cermakian, 2007). And, hormone feedback (Bockmann et al., 1997). This
it is therefore reasonable to ask whether similar distinctive regulatory profile is probably crucial to
415

(a) Thyroid (b)


gland

T4 SP LP

Dio3 Dio3
OH- Outer Inner Ala

OH- Ala OH- Ala


Dio2 Dio2

T3 rT3

OH- Ala

T2

Fig. 1. Thyroid hormone metabolism in the hypothalamus. (a) Reaction scheme showing the enzymatic pathways leading to the
conversion of thyroxine (T4) produced by the thyroid gland into either triiodothyronine (T3) by outer ring deiodination
catalyzed by type 2 deiodinase (Dio2) or reverse T3 (rT3) by inner ring deiondination catalyzed by type 3 deiodinase (Dio3).
The former is an activating process leading to increased thyroid hormone actions through nuclear thyroid hormone receptors; the
latter is an inactivating process. Further metabolism of T3 or rT3 is possible, leading to diiodothyronine (T2) production, which
has no known signaling activity. (b) Reciprocal switching between states of high Dio2/low Dio3 expression and vice versa,
through photoperiodic control. Images are autoradiograms of radioactive in situ hybridization histochemistry for Dio2 or Dio3 in
coronal hypothalamic sections from sheep that were held on either SP (8 h light/24 h) or LP (16 h light/24 h) for 4 weeks prior to
sacrifice. Note strong labeling in the mediobasal hypothalamic area under LP for Dio2, but under SP for Dio3; note also that
Dio2 labeling extends into surrounding hypothalamic tissue, while Dio3 labeling is confined to the ependymal zone immediately
surrounding the third ventricle.

their role in photoperiodic time measurement. maturity (Heyland et al., 2005). Experiments to
Additionally, the current model states that PT- manipulate levels of thyroid hormone in modern
derived TSH acts locally on Dio2/Dio3-expressing day echinoderms (sand dollars) demonstrate that
cells in the hypothalamus or within the PT itself increased TH promotes reproductive activation,
(Hanon et al., 2008; Nakao et al., 2008; Ono et al., while reduced TH favors continued somatic
2008). Hence, recent work defines a fundamentally growth (Heyland and Hodin, 2004). It has been
novel role for TSH as seasonal photoperiodic signal proposed that there was a transition to endogenous
in neuroendocrine system for birds and mammals. thyroid hormone production in ancestral chordates
There are good reasons to consider this novel reg- with the establishment of a secretory structure
ulatory circuit as a reflection of ancestral function of known as the endostyle, which can be seen in
thyroid signaling in vertebrates. According to this modern day sea squirts (urochordates) and in
narrative, thyroid hormone signaling evolved from amphioxus (cephalochordates). The endostyle
an ancient nutrient-sensing pathway, reliant on appears to correspond to the thyroid gland of mod-
plant metabolites as indicators of environmental ern vertebrates, based upon orthologous develop-
favorability to cue the transition to reproductive mental gene expression. Comparative genomics
416

approaches have been used to explore how the subfunctionalization to have occurred in other ver-
additional ingredients seen in the thyroid axis of tebrate groups in which a distinct pituitary stalk
vertebrates appeared during the evolutionary PT region is present—which would include the
pathway originating with ancestral protochordates amphibians and reptiles. Contrastingly, modern
(Paris et al., 2008). According to this scheme, teleosts appear to lack a distinctive pituitary stalk
although a link between TH and reproduction or portal vasculature, relying instead on penetration
stems from the most primitive ancestors of the of neurosecretory fibers projecting from the hypo-
vertebrates, the establishment of a neuroendocrine thalamus into the PD itself, which is directly
pathway controlling TH production occurred apposed to the basal hypothalamus without an
with the agnathans (lampreys and hagfish), with intervening stalk region. Clearly, experiments to
further refinements appearing in subsequent bra- explore TSH and deiodinase gene expression in
nches leading to the current classes of modern the basal hypothalamus and pituitary of species
vertebrates (Fig. 2). from a wider range of vertebrate groups would
It remains unclear whether hypothalamic actions be desirable to clarify this issue and give a proper
of TSH, modulated by local deiodinase expression evolutionary perspective on this pathway.
are also at the center of the photoperiodic
response in modern vertebrate groups other than
birds and mammals. The bird/mammal story points The photoperiodic axis in mammals and its
to subfunctionalization of TSH signaling (light- coupling to brain thyroid signaling
dependent PT thyrotroph local actions and PD
thyrotroph actions on the thyroid) developing prior The described effects of photoperiod on hypotha-
to the divergence of the birds from the mammals. lamic deiodinase expression may be considered
On anatomical grounds, we might expect a similar as a (possibly the) key downstream target of the

Amphibians

TR duplication Mammals/sauropsids

Chondrichthyians
Neuroendocrine control
of TH production
Agnathans
TH-synthesizing
organ Cephalochordates/urochordates
Endogenous TH
production Echinoderms/hemichordates

Ancestral TR
Protostomes

Fig. 2. Evolutionary appearance of different elements of the thyroid signaling pathway during the evolution of modern chordates.
Sensitivity to TH-like molecules produced in the environment is thought to have occurred early in the deuterostomal lineage, prior
to the divergence of the protochordates. Endogenous TH synthesis may have preceded the cephalochordate and urochordate
lineages, but these see the first appearance of a TH-synthesizing organ, the endostyle. Neuroendocrine control of TH is a feature
of all vertebrates including the agnathans, with further refinements appearing in the jawed vertebrates. See Paris et al. (2008) for
further details.
417

so-called photoperiodic axis in birds and mammals. Recent studies of the molecular-transduction
This comprises the set of interconnected neural pathway linking melatonin to the seasonal photo-
and endocrine elements that collectively govern periodic response have focussed on melatonin’s
photoperiodic responses. In mammals, this axis is effects on rhythmical gene expression in the PT,
seemingly linear from light reception in the retina, and the way in which these might lead to a
through a core circadian pacemaker in the hypo- photoperiod-dependent change in levels of TSH
thalamus and its effects on pineal melatonin release (Dardente et al., 2010; Masumoto et al.,
production and culminating with melatonin recep- 2010). Current thinking on this suggests that rhyth-
tor-expressing cells, notably in TSH-expressing mical influences of melatonin on the expression of
cells in the PT (Fig. 2, left). Comprehensive the circadian clock-controlled gene thyrotroph
reviews on photoperiodic influences at different embryonic factor (Tef) and on the transcriptional
levels of the axis as a whole can be found else- coactivator eyes absent 3 (Eya3) are critical to pro-
where (Iuvone et al., 2005; Morgan et al., 1994; cessing of the melatonin signal in the PT (Fig. 3).
Schwartz et al., 2001; Simonneaux and Ribelayga, Of particular importance for this model is the con-
2003). From the perspective of seasonal photope- cept that melatonin has two influences on Eya3
riodism in mammals, the key property of this axis expression, a phase-resetting influence such that
is that it produces a light-modulated circadian the phase of peak Eya3 expression follows the
rhythm of melatonin release. Melatonin secretion evening rise in melatonin by a fixed interval of
is restricted to the night, continuing from after approximately 12 h and a direct inhibitory influ-
dark onset, through to dawn—and changes in ence of melatonin such that the amplitude of peak
the duration of this nocturnal melatonin signal Eya3 expression under conditions when melatonin
are both necessary and sufficient to drive seasonal is still at elevated nighttime levels is markedly
endocrine responses in mammals (Simonneaux lower than if the peak occurs when melatonin
and Ribelayga, 2003). is at low daytime levels. Together, these two

SP LP induction

Photoperiod

E-box
Melatonin
suppression Circadian signal

Ψ ≈ 12 h Ψ ≈ 12 h Transduction
Eya3 mechanism

EYA3 EYA3
TEF TEF
TSHβ Seasonal response
D D

Fig. 3. Photoperiodic switching in mammals through Eya3-dependent transcription of TSHb in the PT. Eya3 is a circadian clock-
controlled gene whose rhythmic expression is set by the phase of evening melatonin onset, and peak Eya3 expression
consistently occurs some 12 h after dark/melatonin onset independent of photoperiod. This means that under SP peak Eya3
expression occurs during the night, while the melatonin level is high and exerting a suppressive effect, and so the peak is small.
Contrastingly, under LP, the Eya3 peak occurs the following morning when the melatonin level is minimal, and so the peak is
large. This classic “external coincidence timer” mechanism, in which a circadian oscillation interacts with a light-dependent
stimulus, limits EYA3/TEF synergism to LP, dictating the onset of a summer phenotype.
418

influences appear to form what is known as a 2005). Deep brain photoreceptors are thought to
“coincidence timing” mechanism to determine the lie in the walls of the third cerebral ventricles,
photoperiodic response (Hazlerigg and Wagner, and three recent studies have argued for specific
2006). It is probably no accident that the phase opsin-like photoreceptors (VA opsin, opn4x/
lag between peak Eya3 expression in the PT, and melanopsin, and opn5) as mediators of the photo-
the evening rise in melatonin is approximately periodic response in birds (Halford et al., 2009;
12 h, leading to triggering of long day responses Kang et al., 2010; Nakane et al., 2010). In each
around the spring equinox, and approximating to case, evidence is largely based on the demonstration
the “critical day length” needed to trigger a sum- of immunoreactivity in the basal hypothalamus, and
mer phenotype in mammals (Elliott, 1980). projections from immunoreactive cells into the PT,
through which it is presumed that TSHb expression
is controlled. It is difficult to place these different
Deep brain photoreceptors and photoperiodic candidates in priority order in terms of likely contri-
signaling in birds bution to the photoperiodic response, but spectral
analysis of the quail photoperiodic response is
Eya3 is also prominent in the photoperiodic consistent with joint contributions from multiple
response of quail (Nakao et al., 2008), and there opsin-like photoreceptors in this region (Foster
is clear conservation in the proximal promoter et al., 1985; Nakane et al., 2010). Intriguingly, we
organization of the TSH beta subunit (TSHb) perhaps should not discount the possibility that mel-
gene between birds and mammals—including the atonin synthesis within the avian hypothalamus is
proximal D-element (RTTAYGTAAY) which is involved in the seasonal photoperiodic response of
key to the TEF/EYA3 control of mammalian birds, with evidence from turkeys for the expression
TSHb (Dardente et al., 2010). Given that Eya3, of two key enzymes in the melatonin synthesis
with TSHb is among the first genes to show pathway (tryptophan hydroxylase 1 and arylamine
increased expression following a transfer to LP N-acetyl transferase) in dopaminergic neurons in
in the quail (Nakao et al., 2008), it is most likely the premammillary hypothalamus of turkeys (Kang
that the TEF/EYA3-inductive mechanism for et al., 2007). Whether this colocalization story
control of TSHb in the PT is essentially conserved reflects an underlying link between melatonin and
between birds and mammals; nonetheless, a photoperiodic timing in birds or a more general link
direct demonstration of this in quail is lacking. between melatonin synthesis and opsin-like photo-
Where birds diverge intriguingly from the receptor expression, as seen in the retina and pineal
mammalian scheme is upstream of Eya3 and gland (Iuvone et al., 2005), remains to be clarified.
TSH control. The prevailing view is that con- I am unaware of any data demonstrating melatonin
trasting with mammals, melatonin is not required receptor expression in the avian PT, but this may be
for the seasonal reproductive response of birds, an avenue that should be reexplored.
and that instead deep encephalic photoreceptors
within the mediobasal hypothalamus relay the
effects of light to the photoperiodic response sys- The evolutionary origins of divergent melatonin
tem. The evidence for this position derives from a actions in different vertebrate lineages
complex series of experiments, in bird species
including ducks and song sparrows as well as the Extra-retinal photoreception appears to be the
Japanese quail, and involving blinding, pinealec- norm among vertebrates, with the mammals
tomy, hypothalamic lesioning, and use of light forming an exception. The reasons for a shift to a
guides to directly deliver light to the basal hypo- purely retinal route of photoreception in mammals
thalamus (reviewed in Dawson et al., 2001; Sharp, have been speculated upon, and a popular
419

hypothesis is that this is a further manifestation of essential appears to be the seasonal photoperiodic
the “nocturnal bottleneck,” thought to have led response. Why this is so remains unclear, but I am
to reduced color vision among the mammals tempted to suggest that it reflects an evolutionary/
(Menaker et al., 1997). developmental constraint within the photoperiodic
Sites of extra-retinal photoreception in the response circuit: it may be noteworthy that the
brains of nonmammals include the pineal gland cell types producing TSH and Dio2/Dio3 are non-
as well as periventricular deep brain sites. Evi- neural in nature and possibly addressable only via
dence from birds and reptiles makes clear that a humoral signal as opposed to a neural efferent
pineal production of melatonin plays a major role pathway.
in circadian coordination in nonmammalian
vertebrates (Underwood, 1990) which appears to
have become greatly attenuated in mammals (cir- Divergent evolution of melatonin receptors
cadian effects of melatonin can be observed in
mammals, particularly in the fetal and neonatal The loss of extra-retinal photoreception appears
period (Davis, 1997), but in adults are largely also to have been associated with a major change
redundant, and loss of pineal melatonin secretion in the structural evolution of melatonin receptor
does not lead to gross deficits in circadian func- subtypes. Three melatonin receptor subtypes
tion). Instead, the restriction of photoreception have been characterized and designated MT1
to the retina in mammals has led to the (previously Mel1a), MT2 (previously Mel1b),
suprachiasmatic nucleus (SCN) assuming a domi- and Mel1c. Of these, MT1 and MT2 are present
nant role in coordinating circadian biology, and a in all vertebrate groups, while Mel1c is found only
subordination of pineal gland as one output path- in nonmammalian vertebrates (Reppert et al.,
way from the retinally entrained SCN pacemaker. 1996a). Contrastingly, peculiar to mammals is
This framework predicts a decline in the extent of the orphan melatonin-related receptor (gpr50),
direct control of circadian physiology through which does not bind melatonin in in vitro expres-
melatonin responsiveness, and this is borne out sion assays and remains a puzzle from a func-
by our current understanding of melatonin recep- tional perspective (Reppert et al., 1996b).
tor biology in different vertebrate groups. Recently by synteny mapping, it was shown that
In the brains of birds (e.g., Siuciak et al., 1991), gpr50 is the mammalian orthologue of Mel1c
fishes (e.g., Gaildrat and Falcón, 2000), and reptiles (Dufourny et al., 2008). Rapid evolutionary diver-
(e.g., Larson-Prior et al., 1996), analysis of melato- gence from the ancestral Mel1c appears to have
nin receptor distribution by in vitro autoradiogra- occurred with the branching of the mammalian
phy, using the radioligand [125I]-2-iodo-melatonin lineage, and in the duck-billed platypus, a Mel1c-
(IMEL) reveals very widespread expression of like Mel1c/gpr50 is found. The underlying causes
binding, whereas in mammals, IMEL binding in a for this divergence are unclear but are most simply
given species shows a highly restricted distribution, interpreted as a consequence of a major chromo-
and in adult individuals, high levels of binding are somal rearrangement, leading to gpr50 residing
only observed consistently in the PT region of the near the end of the long arm of the X-chromosome
pituitary. Analysis of melatonin receptor RNA in eutherian mammals. X-chromosomal genes are
expression by in situ hybridization serves to further typically associated with sequence instability, and
emphasize this point (e.g., compare Reppert et al., the syntenic neighbors of gpr50 show a similar
1995 with Reppert et al., 1994). Hence, in mammals, rapid evolution (Dufourny et al., 2008). Despite
the reorganization of light sensing in the brain has this, gpr50 shows highly localized expression in
led to a restriction of melatonin action to the point the mammalian brain, consistent with maintained
where the only biology for which melatonin is function, and evolutionary sequence analysis
420

suggests that gpr50 has been subject to directional appear to link grp50 to metabolic regulation,
selection in mammals, rather than just neutral drift rather than directly to the photoperiodic
on an evolutionary pathway to a loss of function response, and this insight may help with the
(Dufourny et al., 2008). continuing search for ligands for gpr50.

A link between gpr50 and energy homeostasis Concluding remarks


These reflections get us little closer to establishing Photoperiodic seasonal responsiveness is an
a function for gpr50 in mammals or for identifying ancestral attribute of vertebrates, and its mecha-
possible ligands through which it may exert a nistic origins appear to lie in a thyroid signaling
signaling function. One school of thought, pro- pathway reflecting seasonal feast or famine,
pounded by Levoye et al. (2006), is that gpr50 acts coupled to a light-sensing pathway involving
as a dimerization partner for mammalian melato- retinal and extra-retinal photoreceptors, the cir-
nin receptors, thereby modulating melatonin signal cadian clock and melatonin. We are getting
transduction in a ligand-independent manner. A close to an integrated framework of understand-
difficulty with this hypothesis is the rather limited ing how these elements fit together, and how
overlap between melatonin receptor and gpr50 variant arrangements have evolved among the
expression, the former being concentrated in the vertebrates. Proper understanding of the role
PT, the latter showing the strongest expression in played by gpr50 in mammals will be necessary
brain sites in which melatonin receptor expression to complete this framework.
is weak or absent (Drew et al., 2001; but see also
Sidibe et al., 2010).
In the hypothalamus, gpr50 is expressed in
sites that have been linked to energy homeosta- Abbreviations
sis (e.g., dorsomedial hypothalamus and para-
ventricular hypothalamus) (Drew et al., 2001). Dio2, Dio3 type 2, type 3 deiodinase
Intriguingly, in rodents, gpr50 is also expressed Eya3 eyes absent 3
in the ependymal cell layer lining the wall of gpr50 orphan G-protein-coupled recep-
the third ventricle, overlapping the expression tor 50/melatonin-related receptor
distribution for Dio2 and Dio3 RNA. In Siberian IMEL 2-iodo-melatonin
hamsters (Phodopus sungorus), exposure to Mel1c type Mel1c melatonin receptor
short photoperiods leads to a loss of body fat as MT1 type 1 melatonin receptor
well as a reproductive switch off. At the same MT2 type 2 melatonin receptor
time, levels of grp50 are reduced in the medi- opn4x xenopus-type melanopsin
obasal hypothalamus (Barrett et al., 2006), while opn5 opsin 5
gpr50 null mice show a lean phenotype, with an PD pars distalis
unusual tendency to express torpor in response PT pars tuberalis
to food restriction (Ivanova et al., 2008). Inter- SCN suprachiasmatic nucleus
estingly, in sheep, a photoperiodic species in T4 thyroxine
which dramatic changes in adiposity are not a Tef thyrotroph-embryonic factor
feature, we have not observed photoperiodic TRH thyrotropin-releasing hormone
control of hypothalamic gpr50 expression TSH thyroid-stimulating hormone
(E. Krol, C.A. Wyse and D.G. Hazlerigg unpub- TSHb TSH beta subunit
lished observations). Collectively, these findings VA opsin vertebrate ancient opsin
421

References mediating the photoperiodic response in quail. Nature, 313,


50–52.
Barrett, P., Ivanova, E., Graham, E. S., Ross, A. W., Wilson, D., Gaildrat, P., & Falcón, J. (2000). Melatonin receptors in the
Ple, H., Mercer, J. G., Ebling, F. J., Schuhler, S., Dupré, S. M., pituitary of a teleost fish: mRNA expression, 2-[125I]
Loudon, A., & Morgan, P. J. (2006). Photoperiodic regulation Iodomelatonin binding and cyclic AMP response. Neuroen-
of cellular retinoic acid-binding protein 1, GPR50 and nestin in docrinology, 72(1), 57–66.
tanycytes of the third ventricle ependymal layer of the Halford, S., Pires, S. S., Turton, M., Zheng, L., González-
Siberian hamster. Journal of Endocrinology, 191, 687–698. Menéndez, I., Davies, W. L., et al. (2009). VA opsin-based
Bell-Pedersen, D., Cassone, V., Earnest, D. J., Golden, S. S., photoreceptors in the hypothalamus of birds. Current Biol-
Hardin, P. E., Thomas, T. L., & Zoran, M. J. (2005). Circa- ogy, 19, 1396–1402.
dian rhythms from multiple oscillators: lessons from diverse Hanon, E., Lincoln, G. A., Fustin, J.-M., Dardente, H.,
organisms. Nature Reviews Genetics, 6, 544–556. Masson-Pevet, M., Morgan, P. J., & Hazlerigg, D. G.
Benoit, J. (1936). Role of the thyroid in the gonado-stimulation (2008). Ancestral TSH mechanism signals summer in a sea-
by artificial light in the domestic duck. Comptes rendus de la sonal mammal. Current Biology, 18, 1147–1152.
Société de Biologie, Paris, 123, 243–246. Hazlerigg, D. G., & Loudon, A. S. I. (2008). New insights into
Bockmann, J., Bockers, T. M., Winter, C., Wittkowski, W., ancient seasonal life-timers. Current Biology, 18,
Winterhoff, H., Deufel, T., & Kreutz, M. R. (1997). Thyro- R795–R804.
tropin expression in hypophyseal pars tuberalis-specific cells Hazlerigg, D. G., & Wagner, G. C. (2006). seasonal photoperi-
is 3,5,3'-triiodothyronine, thyrotropin-releasing hormone, odism in vertebrates: from coincidence to amplitude. Trends
and pit-1 independent. Endocrinology, 138, 1019–1028. in Endocrinology & Metabolism, 17, 83–91.
Dahl, G. E., Evans, N. P., & Karsch, F. J. (1994). The thyroid Heyland, A., & Hodin, J. (2004). Heterochronic develop-
gland is required for neuroendocrine responses to photpperiod mental shift caused by thyroid hormone in larval sand
in the ewe. Endocrinology, 135, 10–15. dollars and its implications for phenotypic plasticity and
Dardente, H., & Cermakian, N. (2007). Molecular circadian the evolution of nonfeeding development. Evolution, 58,
rhythms in central and peripheral clocks in mammals. Chro- 524–538.
nobiology International, 24, 195–213. Heyland, A., Hodin, J., & Reitzel, A. M. (2005). Hormone sig-
Dardente, H., Wyse, C. A., Birnie, M. J., Dupre, S., naling in evolution and development: A non-model system
Loudon, A. S. I., Lincoln, G. A., et al. (2010). A molecular approach. Bioessays, 27, 64–75.
switch for photoperiod responsiveness in mammals. Current Iuvone, P. M., Tosini, G., Pozdeyeva, N., Haquea, R.,
Biology, 20, 2193–2198. David, C., Klein, D. C., et al. (2005). Circadian clocks, clock
Davis, F. C. (1997). Melatonin: Role in development. Journal networks, arylalkylamine N-acetyltransferase, and melato-
of Biological Rhythms, 12, 498–508. nin in the retina. Progress in Retinal and Eye Research, 24,
Dawson, A., King, V. M., Bentley, G. E., & Ball, G. F. (2001). 433–456.
Photoperiodic control of seasonality in birds. Journal of Kang, S. W., Leclerc, B., Kosonsiriluk, S., Mauro, L. J.,
Biological Rhythms, 16, 365–380. Iwasawa, A., & El Halawani, M. E. (2010). Melanopsin expres-
Drew, J. E., Barrett, P., Mercer, J. G., Moar, K., Canet, E., sion in dopamine-melatonin neurons of the premammillary
Delagrange, P., et al. (2001). Localization of the melatonin- nucleus of the hypothalamus and seasonal reproduction in
related receptor in the rodent brain and peripheral tissues. birds. Neuroscience, 170, 200–213.
Journal of Neuroendocrinology, 13, 453–458. Kang, S. W., Thayananuphat, A., Bakken, T., & El
Dufourny, L., Levasseur, A., Migaud, M., Callebaut, I., Halawani, M. E. (2007). Dopamine-melatonin neurons in
Pontarotti, P., Malpaux, B., et al. (2008). GPR50 is the the avian hypothalamus controlling seasonal reproduction.
mammalian ortholog of Mel1c: Evidence of rapid evolution Neuroscience, 150(1), 223–233.
in mammals. BMC Evolutionary Biology, 8, 105. Larson-Prior, L. J., Siuciak, J. A., & Dubocovich, M. L. (1996).
Elliott, J. (1980). Circadian rhythms, entrainment and photo- Localization of 2-[125I] iodomelatonin binding sites in visual
periodism in the Syrian hamster. In B. K. Follett & D. E. areas of the turtle brain. European Journal of Pharmacol-
Follett (Eds.), Biological clocks in seasonal reproductive ogy, 297, 181–185.
cycles. Bristol, UK: Scientechnica. Lechan, R. M., & Fekete, C. (2005). Role of thyroid hormone
Follett, B. K., & Potts, C. (1990). Hypothyroidism affects deiodination in the hypothalamus. Thyroid, 15, 883–897.
reproductive refractoriness and the seasonal anoestrus Levoye, A., Dam, J., Ayoub, M. A., Guillaume, J. L.,
period in Welsh mountain ewes. Journal of Endocrinology, Couturier, C., Delagrange, P., et al. (2006). The orphan
127, 103–109. GPR50 receptor specifically inhibits MT1 melatonin recep-
Foster, R. G., Follett, B. K., & Lythgoe, J. N. (1985). tor function through heterodimerization. The EMBO Jour-
Rhodopsin-like sensitivity of extra-retinal photoreceptors nal, 25, 3012–3023.
422

Masumoto, K., Ukai-Tadenuma, M., Kasukawa, T., that mediates reproductive and circadian responses. Neuron,
Nagano, M., Uno, K. D., Tsujino, K., et al. (2010). Acute 13, 1177–1185.
induction of Eya3 by late-night light stimulation triggers Reppert, S. M., Weaver, D. R., Ebisawa, T., Mahle, C. D., &
TSH beta expression in photoperiodism. Current Biology, Kolakowski, L. F. Jr. (1996). Cloning of a melatonin-related
20, 2199–2206. receptor from human pituitary. FEBS Letters, 386, 219–224.
Menaker, M., Moreira, L. F., & Tosini, G. (1997). Evolution of Reppert, S. M., Weaver, D. R., & Godson, C. (1996). Melato-
circadian organisation in vertebrates. Brazilian Journal of nin receptors step into the light: Cloning and classification
Medical and Biological Research, 30, 305–313. of subtypes. Trends in Pharmacological Sciences, 17,
Morgan, P. J., Barrett, P., Howell, H. E., & Helliwell, R. 100–102.
(1994). Melatonin receptors—Localization, molecular phar- Schwartz, W. J., de la Iglesia, H. O., Zlomanczuk, P., &
macology and physiological significance. Neurochemistry Illnerová, H. (2001). Encoding Le Quattro Stagioni within
International, 24, 101–146. the mammalian brain: Photoperiodic orchestration through
Nakane, Y., Ikegami, K., Ono, H., Yamamoto, N., Yoshida, S., the suprachiasmatic nucleus. Journal of Biological Rhythms,
Hirunagi, K., et al. (2010). A mammalian neural tissue opsin 16, 302–311.
(Opsin 5) is a deep brain photoreceptor in birds. Pro- Sharp, P. (2005). Photoperiodic regulation of seasonal breed-
ceedings of the National Academy of Sciences of the United ing in birds. Annals of the New York Academy of Sciences,
States of America, 107, 15264–15268. 1040, 189–199.
Nakao, N., Ono, H., Yamamura, T., Anraku, T., Takagi, T., Sidibe, A., Mullier, A., Chen, P., Baroncini, M., Boutin, J. A.,
Higashi, T., Yasuo, S., Katou, Y., Kageyama, S., Uno, Y., Delagrange, P., et al. (2010). Expression of the orphan
Kasukawa, T., Iigo, M., Sharp, P. J., Iwasawa, A., GPR50 protein in rodent and human dorsomedial hypothala-
Suzuki, Y., Sugano, S., Niimi, T., Mizutani, M., mus, tanycytes and median eminence. Journal of Pineal
Namikawa, T., Ebihara, S., Ueda, H. R., & Yoshimura, T. Research, 48, 263–269.
(2008). Thyrotrophin in the pars tuberalis triggers photope- Simonneaux, V., & Ribelayga, C. (2003). Generation of the
riodic response. Nature, 452, 317–322. melatonin endocrine message in mammals: A review of the
Nakao, N., Ono, H., & Yoshimura, T. (2008b). Thyroid hor- complex regulation of melatonin synthesis by norepineph-
mones and seasonal reproductive neuroendocrine inter- rine, peptides, and other pineal transmitters. Pharmacologi-
actions. Reproduction, 136, 1–8. cal Reviews, 55, 325–395.
Ono, H., Hoshinoa, Y., Yasuo, S., Watanabea, M., Siuciak, J. A., Krause, D. N., & Dubocovich, M. L. (1991).
Nakanea, Y., Muraic, A., et al. (2008). Involvement of thy- Quantitative pharmacological analysis of 2-125I-lodomelatonin
rotropin in photoperiodic signal transduction in mice. Pro- binding sites in discrete areas of the chicken brain. Journal of
ceedings of the National Academy of Sciences of the United Neuroscience, 11, 2855–2864.
States of America, 105, 18238–18242. Underwood, H. (1990). The pineal and melatonin: Regulators
Paris, M., Brunet, F., Markov, G. V., Schubert, M., & of circadian function in lower vertebrates. Cellular and
Laudet, V. (2008). The amphioxus genome enlightens the Molecular Life Sciences, 45, 914–922.
evolution of the thyroid hormone signaling pathway. Devel- Woitkevitsch, A. A. (1940). Dependence of seasonal periodic-
opment Genes and Evolution, 218, 667–680. ity in gonadal changes in the thyroid gland in Sturnus
Reppert, S. M., Weaver, D. R., Cassone, V. M., Godson, C., vulgaris. L.C.R. (Doklady) Acad. Sci. URSS stands for:
& Kolakowski, L. F. Jr. (1995). Melatonin receptors are Transactions (Doklady) of the USSR Academy of Sciences,
for the birds: Molecular analysis of two receptor subtypes 27, 741–745.
differentially expressed in chick brain. Neuron, 15, Yoshimura, T., Yasuo, S., Watanabe, M., Iigo, M., Yamamura, T.,
1003–1015. Hirunagi, K., & Ebihara, S. (2003). Light-induced hormone
Reppert, S. M., Weaver, D. R., & Ebisawa, T. (1994). Cloning conversion of T4 to T3 regulates photoperiodic response of
and characterization of a mammalian melatonin receptor gonads in birds. Nature, 426, 178–181.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 24

A kiss for daily and seasonal reproduction

Valérie Simonneaux*, Isabelle Bur, Caroline Ancel, Laura Ansel and Paul Klosen

Département de Neurobiologie des Rythmes, Institut des Neurosciences Cellulaires et Integratives, Strasbourg, France

Abstract: Reproduction is a fundamental biological function ensuring individual descendant survival


and species perpetuity. It is an energy-consuming process, and therefore, all underlying mechanisms
have to work in synchrony to ensure reproductive success. Synchronization of reproductive activity
with the best time of the day or the year is part of such adaptive processes. Recently, a neuropeptide
named kisspeptin, synthesized in two discrete hypothalamic nuclei, the anteroventral periventricular
nucleus and the arcuate nucleus, has been demonstrated to be a potent stimulator operating upstream
of the gonadotropic axis. In this review, we show how kisspeptinergic neurons integrate daily and
seasonal time cues to synchronize reproductive activity with the cycling environment.

Keywords: circadian clock; season; melatonin; reproduction; kisspeptin; RFRP; thyroid hormone;
hamster.

Hypothalamic kisspeptinergic neurons act drives the secretion of gonadotropins, luteinizing


upstream of the gonadotropic axis hormone (LH) and folliculo-stimulating hormone
(FSH), by the pituitary, which in turn controls
In mammals, early studies on the regulation of the gametogenesis and the production of sex steroids
reproductive axis have emphasized the pivotal role by the gonads. Finally, sex steroids feed back on
of the pulsatile release of gonadotropin-releasing the gonadotropic axis to modulate reproductive
hormone (GnRH), a decapeptide synthesized by activity (Witkin et al., 1982).
a few neurons (about 1000 in mice) located in the Recently, it has been discovered that GnRH syn-
rostral hypothalamus, mainly the preoptic area, thesis and release require the stimulatory action of
and released principally into the pericapillary the hypothalamic peptide kisspeptin (Kp). Indeed
space of the pituitary portal system at the median loss-of-function mutations of the Kp receptor
eminence (Hahn and Coen, 2006). GnRH further (KiSS1R/GPR54) in humans and rodents (de Roux
et al., 2003; Seminara et al., 2003) or the Kiss1 gene
*Corresponding author. coding for Kp (d’Anglemont de Tassigny et al.,
Tel.: þ33-3-88-45-66-71; Fax: þ33-3-88-45-66-54 2007) prevent pubertal development and cause
E-mail: simonneaux@inci-cnrs.unistra.fr infertility. Interestingly, genetic ablation of Kiss1

http://dx.doi.org/10.1016/B978-0-444-59427-3.00024-1 423
424

neurons or Kp-responsive neurons does not pre- AVPV ARC


vent puberty onset or reproductive ability, Kiss1 Kiss1
indicating that Kp signaling is only required when neuron neuron

this neuronal circuitry is present (Mayer and + -


Boehm, 2011). In mammals, the Kiss1 gene is
translated into a 145-amino acid (aa) precursor fur-
ther processed into Kp fragments of smaller sizes GnRH
(from 54 aa in humans—also named metastin—or neuron

52 aa in rodents down to 10 aa). All fragments

Sex steroids
Sex steroids
share the same 10 C-terminus aa sequence with a
final RF-amidated motif, a hallmark of the Median
eminence
RFamide super family of peptides (Kotani et al.,
2001). The common 10 C-terminus aa sequence is Gonadotrope
sufficient to induce a full activation of the KiSS1R cell

(Mikkelsen and Simonneaux, 2009); however, it is


not yet known what is(are) the naturally occurring
endogenous form(s) of the peptide. Gonads
The Kiss1 gene is expressed mainly in two
hypothalamic nuclei, the arcuate (ARC) and the
Fig. 1. Schematic representation of the mammalian
anteroventro-periventricular (AVPV) nuclei
gonadotropic axis as currently accepted. Kiss1-expressing
(Fig. 1), the latter displaying a marked sexual neurons, located in two hypothalamic nuclei, AVPV
dimorphism, being larger in females compared (anteroventro-paraventricular) and ARC (arcuate), project to
to males (Ansel et al., 2010; Clarkson and the GnRH neurons to induce the release of GnRH which
Herbison, 2006; Gottsch et al., 2004; Smith et al., in turn stimulates the production of gonadotropins from
the pituitary. Sex steroids exert a positive (AVPV) or
2005a,b). Kp neurons project to a few brain areas,
negative (ARC) feedback on Kiss1-expressing neurons.
specifically to the preoptic area onto the GnRH
neurons (Clarkson and Herbison, 2006; Clarkson
et al., 2009; Desroziers et al., 2010; Kinoshita
et al., 2005; Mikkelsen and Simonneaux, 2009). central feedback effects of sex steroids underlying
Low doses (from 1 fM) of Kp can potently induce the tightly controlled positive/negative feedback
GnRH (Messager et al., 2005) and thus LH/FSH loops of the reproductive axis (Ansel et al.,
(Gottsch et al., 2004) release. Indeed 80–90% of 2010; Revel et al., 2006a; Smith et al., 2005a,b).
GnRH neurons express c-FOS and are In addition, recent studies have reported that Kp
depolarized after Kp administration (Han et al., neurons integrate other factors, especially those
2005; Irwig et al., 2004). The stimulatory effect known to impact on reproductive activity, like
of Kp on GnRH release occurs via its binding to metabolic factors (Castellano et al., 2010).
KiSS1R which is highly expressed in these Recently, we and others have reported that Kp
neurons. Some experiments suggest that Kp may neurons integrate time of the day and time of
also act at the level of GnRH nerve terminals the year information to adjust reproduction to
(Ansel et al., 2011; d’Anglemont de Tassigny the cycling environment (Clarke et al., 2009;
et al., 2008; Han et al., 2005); however, the cellu- Khan and Kauffman, 2012; Revel et al., 2007;
lar localization of the KiSS1R awaits the produc- Simonneaux et al., 2009).
tion of a specific antibody. Living on earth indeed imposes daily and annual
Strikingly, Kp neurons are the main target for changes in light, temperature, and hygrometry,
the positive (in AVPV) and negative (in ARC) which in turn cause cyclic changes in food quantity
425

and quality. Therefore, reproductive activity further stimulates ovulation at the estrus stage.
should anticipate and adapt to these predictable During the first part of the estrous cycle, cir-
changes such that pups are born when food culating levels of estrogen are low and exert a
resources are optimal for their survival. Mammals negative feedback effect (Shupnik, 1996). On the
have a dedicated photoneuroendocrine pathway day of proestrus, the circulating levels of estrogen
converting the physical light information into increase and the sex hormone exerts a positive
nervous and endocrine signals which synchronize feedback. This stage is a prerequisite to trigger
biological functions with the environment ovulation. Strikingly, the preovulatory LH surge
(Simonneaux and Ribelayga, 2003). Briefly, this occurs only at a specific time window on the day
pathway involves a master biological clock located of proestrus, typically at the end of the afternoon.
in the suprachiasmatic nuclei (SCN). Circadian It is generally acepted that the occurrence of the
clock activity is synchronized to the 24 h light/ LH surge on the late afternoon of the proestrus
dark cycle by the retino-hypothalamic pathway requires the combination of two signals, a daily
originating in a specific set of melanopsin- signal arising from the hypothalamic clock and a
expressing ganglion cells of the retina. Impor- sufficient circulating level of estrogen. Recent
tantly, the circadian clock also integrates annual studies suggest that the Kp neurons of the AVPV
changes in photoperiod since the clock’s daily are the cellular site integrating these two signals
activity is modulated according to the light/dark for the GnRH-induced LH surge (Fig. 2).
ratio. Timing information built in the hypothalamic
biological clock is mainly forwarded to other hypo-
thalamic nuclei, which in turn set various biological The estrogen signal
activities in phase with the time of the day and year
(Buijs and Kalsbeek, 2001; Kalsbeek et al., 2006). Estrogen binds specific receptors acting via genomic
As far as reproduction is concerned, two clock- (ERa and ERb) or nongenomic (GPR30) pat-
driven pathways have been reported to impact on hways, ERa mediating the positive retroactive
gonadotropic activity: a direct SCN nervous output effect of estrogen in the brain (Wintermantel et al.,
involving vasopressin (VP)- and vasoactive intesti- 2006). ERa, however, are not located on GnRH
nal peptide (VIP)-containing neurons and an indi- neurons but rather on the Kp neurons of the AVPV
rect SCN-driven endocrine pathway involving the (Smith et al., 2006), and various experiments of
rhythmic release of the pineal hormone melatonin. gonadectomy or hormone administration have
Recent reports have demonstrated that Kp demonstrated that Kiss1 expression in female
neurons are set at the interface of these clock out- AVPV is strongly upregulated by estrogen (Ansel
puts and the GnRH neurons to synchronize repro- et al., 2010; Smith et al., 2005a). In addition, central
ductive activity with the time of the day and year. administration of Kp antibody or KiSS1R antago-
nist prevents the preovulatory LH surge (Kinoshita
et al., 2005; Pineda et al., 2010a). Altogether these
The kisspeptinergic neurons of the AVPV observations demonstrate that estrogen exerts its
integrate circadian information to time the positive feedback via the Kiss1 neurons of the
preovulatory LH surge AVPV to trigger the LH surge (Fig. 2).

In female rodents, the estrous cycle comprises


four stages of activity: metestrus, diestrus, proes- The circadian signal
trus, and estrus (the stage when ovulation occurs),
and lasts 4–5 days. Proestrus is characterized by a There is strong evidence that the circadian clock
marked and transient increase in LH, which governs the estrous cycle and more specifically
426

VP
neuron AVPV E2
Kiss1
VP neuron

SCN

Kp
POA
VIP
VIP GnRH
neuron neuron

GnRH

Gonadotrope E2
cell
LH secretion

LH surge

Metestrus Diestrus Proestrus Estrus


Ovarian E2 secretion

Fig. 2. Schematic representation of the circadian and estrogen signals controlling the LH surge in female rodents. Vasopressin (VP)
and vasoactive intestinal peptide (VIP) neurons located in the suprachiasmatic nuclei (SCN) signal the circadian time to the Kiss1
neurons of the anteroventro-periventricular nuclei (AVPV) and the GnRH neurons of the preoptic area (POA), respectively. The
coincidence of high circulating levels of estrogen (E2) and the circadian signal induces a surge of LH in the afternoon of the
proestrus.

the timing of the preovulatory LH surge in Third, SCN outputs toward GnRH and Kp
rodents. First, female mice with mutated circadian neurons have been described (de la Iglesia and
clock genes that are involved in encoding daily Schwartz, 2006; Williams et al., 2011), and two
biological time present alterations in reproductive SCN transmitters, VIP and VP, have been
capacities (Miller et al., 2004; Pilorz and reported as putative circadian signals to the
Steinlechner, 2008; Ratajczak et al., 2009). For reproductive system (Khan and Kauffman,
example, the Clock/Clock mutant female mice 2012). The current concept is that VIP can
present irregular estrous cycles, and this alter- directly stimulate GnRH neurons’ firing in a
ation is associated with a suppression of the LH time-dependent manner, and VP neurons project
surge and ovulation (Miller et al., 2004). Second, directly to the AVPV Kp neurons. About 50%
neuroanatomical studies have reported that SCN of the AVPV Kp neurons receive SCN VP fibers
ablation suppresses the LH surge (Antunes- (Vida et al., 2010) and the daily pattern of VP
Rodrigues and McCann, 1967), impairs the synthesis and release matches Kp neuron activity
estrous cycle, and leads to an anovulatory condi- with a peak in the late afternoon (Shinohara
tion characterized by a persistent estrous stage et al., 1994). Moreover, VP injection in the
(Brown-Grant and Raisman, 1977; Stetson and AVPV area induces an LH surge, even in SCN-
Watson-Whitmyre, 1976; Wiegand et al., 1980). lesioned animals (Palm et al., 1999), and
427

conversely, infusion of a VP antagonist on the day secretion in response to Kp and VIP (Gillespie
of proestrus prevents the LH surge (Funabashi et al., 2003; Zhao and Kriegsfeld, 2009). Interest-
et al., 1999). Although no experiments have yet ingly, a recent study reported that a circadian
formally proven that VP regulates Kp synthesis transcription factor, albumin D-site binding pro-
and/or release in the AVPV neurons, it is tempt- tein (Dbp), is able to induce Kiss1 transcription
ing to speculate that the Kp neurons of the in AVPV Kp neurons (Xu et al., 2011). However,
AVPV represent a key interface to integrate the further experiments are necessary to clarify the
circadian signal, in addition to the estrogen signal, existence of a circadian machinery in the AVPV
to time the LH surge in the late afternoon of the Kp neurons, which may govern the rhythmic
proestrus (Fig. 2). expression of Kiss1 and other rhythmic cues.

A circadian oscillator in Kp neurons? RFRP-3, another member of the RFamide family


of peptides, in the game
Interestingly, recent findings demonstrate that
AVPV Kp neurons also exhibit a robust circadian Recently, another member of the RFamide family
pattern of gene expression and neuronal activa- of peptides, RFRP-3, was proposed to be
tion coincident with the circadian pattern of LH involved in the circadian regulation of the female
(Robertson et al., 2009). Indeed, ovariectomized reproductive cycle (Gibson et al., 2008). The Rfrp
female mice implanted with an estrogen implant gene, the mammalian ortholog of avian gonado-
(OVX þ E2) showed a robust surge of LH 1 h tropin inhibitory hormone (GnIH), encodes a pre-
before the subjective night (CT11 and 12), cursor that produces two peptides, RFRP-1 and
whereas low levels of LH were detected at the RFRP-3 (Clarke et al., 2008; Fukusumi et al.,
other time points. In synchrony with this LH 2001; Hinuma et al., 2000; Kriegsfeld et al., 2006;
secretion, a strong circadian pattern with higher Ukena and Tsutsui, 2005; Ukena et al., 2002;
Kiss1 and c-Fos mRNA levels at CT11 and 12 Yoshida et al., 2003). RFRP neurons are located
was observed in the AVPV. Moreover, in intact in the dorsomedial nucleus of the hypothalamus
female mice, c-Fos and Kiss1 mRNA levels are (DMH) and fiber networks are found in multiple
significantly increased in Kp neurons in the late brain regions and particularly on GnRH neurons
afternoon of the proestrus day coincident with (Johnson et al., 2007; Mason et al., 2010; Ukena
the LH surge (Bur et al., unpublished; Smith and Tsutsui, 2001). A large body of evidence
et al., 2006). These findings demonstrate that cir- now indicates that both central and peripheral
cadian-timed neuronal activation of Kp neurons administration of RFRP-3 inhibits LH secretion
in the AVPV is involved in the late afternoon (Anderson et al., 2009; Clarke et al., 2008; John-
preovulatory LH surge. The circadian signal may son et al., 2007; Kadokawa et al., 2009; Kriegsfeld
arise from VP-containing SCN outputs; however, et al., 2006; Pineda et al., 2010b). RFRP-3
it may also come from an intrinsic autonomous neurons express ERa, suggesting that they may
circadian clock located in AVPV Kp neurons. mediate estrogen feedback (Kriegsfeld et al.,
Recent studies revealed that circadian clock gene 2006). Estrogen downregulates Rfrp mRNA
expression in neuroendocrine structures interferes expression in ovariectomized mice (Molnar et al.,
with their rhythmic behavior (Bonnefont, 2010). 2011) and RFRP-3 cell numbers as well as c-FOS/
Thus, clock genes and proteins are rhythmically RFRP-3 coexpression are reduced in proestrus
expressed in the immortalized GnRH-secreting Syrian hamsters, relative to diestrus (Gibson
GT1-7 cell line, and these cells exhibit daily et al., 2008). These results indicate that estrogen
changes in peptide expression and GnRH downregulation of the RFRP-3 system alleviates
428

inhibition of GnRH release during the LH surge. marked involution of the gonads and accessory
Interestingly, in rats and ewes, there are no organs resulting in low levels of circulating sex ster-
changes in Rfrp expression following alterations oids. Surgical removal of the pineal gland before
in estradiol levels (Quennell et al., 2010; Smith exposure to SD prevents hamsters from undergoing
et al., 2008), indicating that there may be species sexual inactivation. Conversely, exogenous melato-
differences in the involvement of RFRP neurons nin injections mimicking SD conditions in hamsters
in estrogen-feedback mechanisms. Little is known raised in LD induce sexual inactivation. Finally,
about the interaction between the SCN and the when SD exposure is prolonged over 25–30 weeks,
RFRP-3 system. In rodents, both VP- and VIP- hamsters become refractory to the inhibitory
positive fibers from the SCN project to the melatoninergic signal and undergo a complete
DMH; however, it is not known which of these, endogenous reactivation of the reproductive system.
if any, innervate RFRP-3 cells (Abrahamson and Strikingly, not all seasonal species respond to the
Moore, 2001; Gibson et al., 2008; Kalsbeek photoperiodic changes the same way, and this is
et al., 1993). Taken together, these data indicate related to the length of gestation time. Thus, in
that, in some species, RFRP-3 may play a role in sheep, where gestation lasts approximately 5
the regulation of the estrus cycle, notably in months, reproduction is stimulated in SD conditions
the estrogen-dependent feedback mechanisms. and becomes quiescent as days become longer. As a
However, whether and how this action is con- consequence, they are regarded as SD breeders,
trolled by the circadian system remains to be while hamsters are LD breeders. However, in
investigated. both photoperiodic models, the seasonal rhythm
of reproduction is synchronized by changes in
nocturnal melatonin peak duration. In the Syrian
The kisspeptinergic neurons of the ARC nucleus hamster, an increase in melatonin peak duration
integrate photoperiod/melatoninergic information inhibits reproductive activity, whereas in sheep,
to time seasonal reproduction the same melatoninergic information stimulates
reproduction.
Most animals living in the wild restrict fertility to a Although the role of melatonin in synchronizing
particular time of the year to ensure the birth and seasonal reproduction is indisputable, the cellular
weaning of the offspring during the most favorable and molecular sites of action of melatonin are still
season. Although various environmental factors not well understood. It is clear that melatonin
are important to take into consideration, a majority does not act directly on GnRH neurons and
of species use the highly predictable annual responsiveness to GnRH does not change with pho-
variations of light duration (or photoperiod) to toperiod (Urbanski et al., 1991). Melatonin-binding
establish the time of year. Since the pioneer studies sites have been identified in a number of brain
of Reiter and Hoffman in the 1960s, it is well structures but with considerable species differences
established that the annual/photoperiodic rhythm (Masson-Pevet et al., 1994). Lesion or melatonin-
in nocturnal melatonin is critical to synchronize infusion studies at various anatomical sites
reproduction with the seasons (Bartness et al., suggested that the mediobasal hypothalamus is
1993; Hoffman and Reiter, 1965; Pitrosky and Pevet, involved in melatonin action (Malpaux et al., 1993;
1997). Thus, Syrian hamsters raised in long day Maywood and Hastings, 1995). Besides, a high den-
(LD, 14 light/10 h dark) conditions are sexually sity of melatonin receptors has been identified in the
active and, upon exposure to short day (SD, 10 pars tuberalis of the adenohypophysis of a number
light/14 h dark) conditions for 8–10 weeks, undergo of mammalian species (Masson-Pevet and Gauer,
a dramatic inhibition of reproductive activity with 1994), pointing to this structure as a critical photope-
low levels of circulating gonadotropins and a riodic target for melatonin action.
429

Kp neurons as gatekeepers of seasonal (a)


reproduction

Total number of Kiss1 neurons


250

The potent stimulatory effect of Kp on GnRH 200 *

% of control
and gonadotropin release, the observation that
150
the sexual phenotype of KiSS1R-mutated mice
resembles that of hamsters exposed to SD, and 100
the localization of Kiss1-expressing neurons in
50
the mediobasal hypothalamic area were all strong
arguments to test, in the Syrian hamster, the 0
hypothesis that photoperiod/melatonin could reg- SD SD + PinX
ulate seasonal reproduction via an action on (b)
Kiss1-expressing neurons (Ansel et al., 2010,

Total number of Kiss1 neurons


2011; Revel et al., 2006a, 2007). 150
We first observed that Kiss1 mRNA and Kp-
immunoreactivity are found in neurons of the

% of control
100
ARC and AVPV of the adult male Syrian
hamsters, similar to other reports in rats and *
mice. When Syrian hamsters were kept for 8–10 50
weeks in SD, in parallel to the expected impor-
tant reduction in testis size and circulating levels
0
of testosterone, the number of Kiss1-expressing LD LD + Mel
neurons was markedly reduced in both nuclei. In
the ARC, the SD-induced reduction in Kiss1 Fig. 3. Melatonin regulation of Kiss1-expressing neurons in
the arcuate nucleus of male Syrian hamsters. (a) Hamsters
expression was not the result of decreased levels were pinealectomized (PinX) or sham-operated and kept in
of testosterone since castration of LD Syrian short day conditions (SD) for 10 weeks. The number of
hamsters led to a significant increase in Kiss1 Kiss1-expressing neurons (1) was higher in PinX, sexually
mRNA levels in the ARC. This is in agreement active, hamsters. (b) Hamsters were raised in long day (LD)
with an inhibitory effect of sex hormones on Kiss1 conditions and injected daily with ringer or melatonin (50 mg,
100 ml/animals; 2 h before lights off) for 8 weeks. The
expression in the ARC, as already reported in number of Kiss1-expressing neurons (measured by in situ
other rodent species. In contrast, we observed hybridization) was lower in melatonin-injected, sexually
that the removal of endogenous melatonin by inactive hamsters. Values are expressed as percent of control;
pinealectomy in SD-exposed Syrian hamsters *p < 0.05 as compared to control hamsters.
increased ARC Kiss1 mRNA and reactivated
reproductive activity within a few weeks (Fig. 3a). stimulates Kiss1 gene expression; therefore, the
Conversely, late afternoon melatonin injections in SD-induced decrease in Kiss1 mRNA is a conse-
sexually active Syrian hamsters kept in LD signif- quence of melatonin/SD-induced reduction in
icantly reduced ARC Kiss1 mRNA levels together the circulating level of testosterone. These
with a marked reduction in testicular activity findings strongly suggest that the long nocturnal
(Fig. 3b). It is important to note that melatonin peak of melatonin in SD inhibits reproductive
treatment had to be carried out for at least 5 activity via a sex steroid-independent (ARC) or
weeks to be effective. In the AVPV, the level of a sex steroid-related (AVPV) inhibition of Kiss1
Kiss1 mRNA was also reduced in SD compared expression in the male Syrian hamster. Strikingly,
to LD, but the underlying mechanisms are differ- when Syrian hamsters became refractory to the
ent. Indeed, in this structure, testosterone inhibitory melatoninergic signal (kept over 25
430

weeks in SD), the spontaneous sexual reactiva- appears unable to reactivate the reproductive axis
tion was associated with an increase in Kiss1 (Greives et al., 2008). Whether this discrepancy is
mRNA in the ARC, indicating that Kiss1 expres- due to differences in dose, route of administration
sion also escapes melatonin’s inhibitory signal in or species is being explored in our laboratory.
photorefractory animals. We further tested the The effect of photoperiod/season on Kiss1
hypothesis that the SD-driven reduction in hypo- expression has also been examined in the ewe
thalamic Kp may be responsible for the inhibition (Caraty et al., 2007; Franceschini et al., 2006; Smith
of reproductive activity (Ansel et al., 2011; Revel et al., 2007, 2009). Kiss1-expressing neurons are
et al., 2006a). Thus, sexually inactive Syrian found in the ARC, and Kiss1 mRNA levels are
hamsters, kept in SD, received a chronic intra- increased during the breeding season in SD. Kp
cerebroventricular infusion of 0.25 nmol/h Kp10 infusion during the anestrous season caused ovula-
or vehicle for 4 weeks. Remarkably, the Kp- tion in more than 80% of treated animals, whereas
treated animals underwent a strong reactivation less than 20% of control animals ovulated.
of testicular activity and increased levels of cir- This suggests that, similar to hamsters, a steroid-
culating testosterone, in marked contrast to the independent seasonal modulation of Kiss1 expres-
vehicle-treated hamsters. This gonadal response sion occurs in the ARC of sheep to drive seasonal
was of similar amplitude to that observed in SD reproduction. However, whether this effect on Kp
hamsters transferred back to LD on the day of neurons is also controlled by melatonin has not been
surgery. Acute peripheral (intraperitoneal, intra- demonstrated. If melatonin is indeed driving sea-
venous, or subcutaneous) administration of Kp is sonal changes in Kiss1 expression in the sheep as it
known to stimulate reproductive activity almost does in the Syrian hamster, it will be a major chal-
as efficiently as an acute central infusion lenge to determine why a long SD melatonin peak
(Mikkelsen et al., 2009; Thompson et al., 2004). inhibits Kiss1 expression in Syrian hamsters and
In line with this observation, we found that twice increases it in sheep. This may help to identify the
daily intraperitoneal injections of Kp54 in SD- long-sought-after mechanisms underlying the differ-
inhibited hamsters could restore testicular activity ence between LD and SD breeders. Most probably,
as well. Altogether, our data demonstrate that, in melatonin is not acting directly on ARC neurons (Li
Syrian hamsters, melatonin in SD inhibits Kp et al., 2011; Maywood et al., 1996), and therefore,
expression in the ARC and AVPV which in turn melatonin-sensitive system(s), operating upstream
downregulates reproductive activity (Fig. 4). of the Kiss1 neurons, have to be discovered. Two
Photoperiodic variation of Kp expression was candidates, the RFRP neurons, located in the
reported in other seasonal mammals. In male DMH, and the thyroid stimulating hormone
Siberian hamsters kept in SD, Kiss1 gene expression (TSH)-regulated deiodinases, located in the
is reduced in the AVPV but, unexpectedly, higher in tanycytes of the basal hypothalamus, may be the
the ARC compared to LD (Mason et al., 2007; missing link(s) between melatonin and the Kp-
Simonneaux et al., 2009). In Siberian, like in Syrian regulated seasonal reproduction.
hamsters, testosterone inhibits Kiss1 expression in
the ARC; thus it is possible that the reduced levels
of testosterone in SD are sufficient to explain the RFRP neurons: A missing link between
increase in Kiss1 expression in the Siberian hamster, melatonin and Kiss1 neurons?
whereas in the Syrian hamster, this effect would be
blunted by the strong inhibitory effect of melatonin. Because melatonin receptors have been identified
Acute Kp injection in the male Siberian hamster in the Syrian hamster DMH, and because lesions
triggers LH release, but chronic subcutaneous of this same area prevent SD-induced testicular
administration of Kp in photoinhibited hamsters regression (Maywood and Hastings, 1995), we
431

RFRP neuron

DMH

T3 TSHR
Dio2/3
Kiss neuron T4
Kiss neuron

tanycytes 3V

AVPV ARC
cap

TSH
GnRH neuron

cap

cap
Melatonin
CH3
TSH
CH3O
CH2 CH2 N
H
C
O
MeIR pars tuberalis
N

Fig. 4. Schematic representation of a putative hypothalamic T3/RFRP/Kp network controlling seasonal reproduction in mammals.
Melatonin binding to its receptors located in the pars tuberalis regulates TSH release. TSH in turn binds to its receptors located on
the tanycyte cell bodies lining the basal part of the third ventricle to regulate diodinases 2/3 activities and therefore local metabolism
of the thyroid hormones T4 and T3. Kiss1 neurons located in the anteroventro-periventricular (AVPV) and arcuate (ARC) nuclei
project to GnRH cell bodies and nerve terminals to stimulate GnRH release in the portal blood at the median eminence. RFRP-
expressing neurons located in the dorsomedial hypothalamus (DMH) also project to the GnRH neurons, and possibly to the
Kiss1 neurons. Notably, melatonin in short photoperiod downregulates expression of Dio2, Rfrp, and Kiss1 genes in the Syrian
hamster. Additionally, melatonin regulates the plasticity of the tanycytes’ end feet in the median eminence.

hypothesized that RFRP neurons might be the pinealectomized hamsters, whereas injections of
missing link between melatonin and Kiss1. We LD hamsters with melatonin reduced Rfrp expres-
subsequently demonstrated that, in Syrian and sion down to SD levels (Revel et al., 2008),
Siberian hamsters, the level of Rfrp mRNA indicating that the regulation is dependent on mel-
and the number of RFRP-immunoreactive cell atonin. We also found that central administration
bodies were reduced in sexually quiescent animals of RFRP-3 in sexually active male Syrian hamsters
acclimated to SD compared with sexually active produced a significant increase in both LH and tes-
animals maintained under LD (Revel et al., tosterone plasma concentrations (Ancel et al.,
2008). The reduction in Rfrp expression in 2012). Whereas this result is in complete contra-
SD was independent from secondary changes diction with the available data in the literature, it
in gonadal steroids and was abolished in is in accordance with the observation that Rfrp
432

expression is higher in sexually active Syrian located on the tanycytes (Hanon et al., 2008; Nakao
hamsters relative to sexually quiescent hamsters. et al., 2008). Fourth, chronic central infusion of T3
We also observed that a 5-week-long central infu- reactivates photoinhibited reproductive activity in
sion of RFRP-3 to sexually quiescent SD male the quail (Yoshimura et al., 2003) and Siberian ham-
Syrian hamsters restored testicular activity (Ancel ster (Barrett et al., 2007). Altogether, these recent
et al., 2012), similarly to Kp (Revel et al., 2006a). findings point to the pars tuberalis TSH/tanycyte
Remarkably, we found that the administration of T3 model as a pivotal system for the photoperiodic
RFRP-3 induced a significant increase in Kiss1 control of reproduction. In mammals, the
mRNA levels in the ARC (Ancel et al., 2012). mechanisms involved in the T3-driven regulation
Taken together, these data indicate that RFRP-3 of reproductive activity are yet unknown. Interest-
stimulates the reproductive function in the male ingly, we recently reported in the Siberian hamster
Syrian hamster, likely via the Kiss1 neurons of that melatonin displays a profound remodeling of
the ARC (Fig. 4). the tanycyte structure, especially in the median emi-
nence area where the GnRH nerve fibers end
(Bolborea et al., 2011). Because GnRH secretion
TSH-driven regulation of tanycyte deiodinases is is tightly controlled by the tanycytes at their nerve
central for seasonal processes terminals (Prevot et al., 2003, 2007), it can be
hypothesized that the local production of T3 by
Thyroid hormones have long been known to be Dio2 may regulate reproductive activity via a modu-
involved in seasonal reproduction (Karsch et al., lation of GnRH release at the median eminence
1995), but only quite recently the underlying (Yamamura et al., 2006).
mechanisms and their relationship with melatonin
started to be understood (Fig. 4). First, in European
hamsters, cells of the pars tuberalis containing mela- Toward a seasonal T3/RFRP/Kp network in the
tonin receptors synthesize TSH in a photoperiod- mediobasal hypothalamus?
dependent manner with lower levels in SD
conditions (Dardente et al., 2003) and the molecular In the recent years, the findings that (1) expres-
signaling mechanism of melatonin on TSH expres- sion of Kp and RFRP in the mediobasal hypothal-
sion was recently described in sheep pars tuberalis amus and local thyroid hormone metabolism by
(Dardente et al., 2010). Second, deiodinases the tanycytes are markedly regulated by melato-
2 (Dio2) and deiodinases 3 (Dio3), the enzymes nin/photoperiod in seasonal mammals and (2)
involved in the local metabolism of the thyroid hor- each of these transmitters can modulate repro-
mones T4 and T3, are expressed specifically in ductive activity have led us to reconsider the neu-
tanycytes, a specific subset of glial cells lining the roendocrine control of seasonal reproduction
basal part of the third ventricle, in a photoperiod- (Fig. 4). Obviously, the three transmitters play a
dependent manner in seasonal breeders (quail critical role in this regulation, but whether they
(Yoshimura et al., 2003), Syrian hamster (Revel act independently or not is yet to be established.
et al., 2006b), Siberian hamster (Barrett et al., Remarkably, Kp is always stimulatory of repro-
2007; Watanabe et al., 2004), European hamster ductive activity (regardless of the species tested),
(Hanon et al., 2010), sheep (Hanon et al., 2008), but the lack of melatonin receptors in the rodent
and goat (Yasuo et al., 2006)). Notably, this photo- and sheep ARC indicates that Kp neurons are
periodic regulation is also present in some mice located downstream of the melatoninergic signal.
strains, which are not seasonal breeders (Yasuo By contrast, the RFRP-3 effect on gonadotropin
et al., 2009). Third, TSH secreted by the pars and sex steroid production is species dependent,
tuberalis stimulates Dio2 activity via TSH receptors being stimulatory in Syrian and Siberian hamsters
433

but inhibitory in sheep. RFRP-3 neurons may References


therefore be located at a crucial node to translate
the photoperiodic melatoninergic message as Abrahamson, E. E., & Moore, R. Y. (2001). Suprachiasmatic
nucleus in the mouse: Retinal innervation, intrinsic organi-
inhibitory or stimulatory of reproductive activity.
zation and efferent projections. Brain Research, 916,
Although melatonin-binding sites have been 172–191.
reported in the DMH area of the Syrian hamster, Ancel, C., Bentsen, A. H., Sebert, M. E., Tena-Sempere, M.,
whether melatonin receptors are localized on Mikkelsen, J. D., & Simonneaux, V. (2012). Stimulatory
RFRP-3 neurons remains to be established. Inter- effect of RFRP-3 on the gonadotrophic axis in the male
Syrian hamster: The exception proves the rule.
estingly, we have observed that the testicular
Endocrinology, 153, 1352–1363.
reactivation of photoinhibited Syrian hamsters Anderson, G. M., Relf, H. L., Rizwan, M. Z., & Evans, J. J.
by a chronic central infusion of RFRP-3 is coinci- (2009). Central and peripheral effects of RFamide-related
dent with a significant increase in ARC Kiss1 peptide-3 on luteinizing hormone and prolactin secretion
expression, indicating that RFRP neurons could in rats. Endocrinology, 150, 1834–1840.
Ansel, L., Bentsen, A. H., Ancel, C., Bolborea, M., Klosen, P.,
be upstream of the Kp neurons for the mela-
Mikkelsen, J. D., et al. (2011). Peripheral kisspeptin reverses
toninergic regulation of reproductive activity in short photoperiod-induced gonadal regression in Syrian
these hamsters. Finally, the pars tuberalis/TSH hamsters by promoting GnRH release. Reproduction, 142,
control of tanycyte T3 production is set at the first 417–425.
position to translate the melatoninergic signal Ansel, L., Bolborea, M., Bentsen, A. H., Klosen, P.,
Mikkelsen, J. D., & Simonneaux, V. (2010). Differential reg-
toward the central control of reproduction, but it
ulation of kiss1 expression by melatonin and gonadal hor-
is yet unknown where and how T3 exerts this con- mones in male and female Syrian hamsters. Journal of
trol. In spite of the great advances made recently Biological Rhythms, 25, 81–91.
in the understanding of the seasonal control of Antunes-Rodrigues, J., & McCann, S. M. (1967). Effect of
reproduction, particularly in identifying the critical suprachiasmatic lesions on the regulation of luteinizing hor-
mone secretion in the female rat. Endocrinology, 81,
transmitters involved, some pieces of the puzzle
666–670.
are still lacking to fully understand how photope- Barrett, P., Ebling, F. J., Schuhler, S., Wilson, D., Ross, A. W.,
riodic changes in melatonin production synchro- Warner, A., et al. (2007). Hypothalamic thyroid hormone
nize reproduction in seasonal mammals. catabolism acts as a gatekeeper for the seasonal control of
body weight and reproduction. Endocrinology, 148,
3608–3617.
Conclusion Bartness, T. J., Powers, J. B., Hastings, M. H., Bittman, E. L.,
& Goldman, B. D. (1993). The timed infusion paradigm for
melatonin delivery: What has it taught us about the melato-
The 2003 milestone discovery of the pivotal role
nin signal, its reception, and the photoperiodic control of
of Kp as a potent stimulator of GnRH release seasonal responses? Journal of Pineal Research, 15, 161–190.
has helped to understand how the time of the Bolborea, M., Laran-Chich, M. P., Rasri, K., Hildebrandt, H.,
day and year is integrated in brain structures to Govitrapong, P., Simonneaux, V., et al. (2011). Melatonin
synchronize reproductive activity with the cycling controls photoperiodic changes in tanycyte vimentin and
neural cell adhesion molecule expression in the Djungarian
environment. The latest findings indicate that the
hamster (Phodopus sungorus). Endocrinology, 152,
SCN VP output connection with AVPV Kp 3871–3883.
neurons is critical to time the LH surge, at least Bonnefont, X. (2010). Circadian timekeeping and multiple
in rodents. In seasonal species, the ARC Kp timescale neuroendocrine rhythms. Journal of Neuroendo-
neurons appear to be the final gatekeeper to syn- crinology, 22, 209–216.
Brown-Grant, K., & Raisman, G. (1977). Abnormalities in
chronize reproduction with seasons. It remains to
reproductive function associated with the destruction of
be discovered what is(are) the system(s) acting the suprachiasmatic nuclei in female rats. Proceedings of
upstream to control the photoperiodic synthesis the Royal Society of London - Series B: Biological Sciences,
of Kp. 198, 279–296.
434

Buijs, R. M., & Kalsbeek, A. (2001). Hypothalamic integration Desroziers, E., Mikkelsen, J., Simonneaux, V., Keller, M.,
of central and peripheral clocks. Nature Reviews Neurosci- Tillet, Y., Caraty, A., et al. (2010). Mapping of kisspeptin
ence, 2, 521–526. fibres in the brain of the pro-oestrous rat. Journal of Neuro-
Caraty, A., Smith, J. T., Lomet, D., Ben Said, S., Morrissey, A., endocrinology, 22, 1101–1112.
Cognie, J., et al. (2007). Kisspeptin synchronizes preovulatory Franceschini, I., Lomet, D., Cateau, M., Delsol, G., Tillet, Y.,
surges in cyclical ewes and causes ovulation in seasonally & Caraty, A. (2006). Kisspeptin immunoreactive cells of
acyclic ewes. Endocrinology, 148, 5258–5267. the ovine preoptic area and arcuate nucleus co-express
Castellano, J. M., Bentsen, A. H., Mikkelsen, J. D., & Tena- estrogen receptor alpha. Neuroscience Letters, 401, 225–230.
Sempere, M. (2010). Kisspeptins: Bridging energy homeo- Fukusumi, S., Habata, Y., Yoshida, H., Iijima, N.,
stasis and reproduction. Brain Research, 1364, 129–138. Kawamata, Y., Hosoya, M., et al. (2001). Characteristics
Clarke, I. J., Sari, I. P., Qi, Y., Smith, J. T., Parkington, H. C., and distribution of endogenous RFamide-related peptide-1.
Ubuka, T., et al. (2008). Potent action of RFamide-related Biochimica et Biophysica Acta (BBA). Molecular Cell
peptide-3 on pituitary gonadotropes indicative of a Research, 1540, 221–232.
hypophysiotropic role in the negative regulation of gonado- Funabashi, T., Aiba, S., Sano, A., Shinohara, K., & Kimura, F.
tropin secretion. Endocrinology, 149, 5811–5821. (1999). Intracerebroventricular injection of arginine-vaso-
Clarke, I. J., Smith, J. T., Caraty, A., Goodman, R. L., & pressin V1 receptor antagonist attenuates the surge of
Lehman, M. N. (2009). Kisspeptin and seasonality in sheep. luteinizing hormone and prolactin secretion in proestrous
Peptides, 30, 154–163. rats. Neuroscience Letters, 260, 37–40.
Clarkson, J., d’Anglemont de Tassigny, X., Colledge, W. H., Gibson, E. M., Humber, S. A., Jain, S., Williams, W. P., 3rd,
Caraty, A., & Herbison, A. E. (2009). Distribution of Zhao, S., Bentley, G. E., et al. (2008). Alterations in
kisspeptin neurones in the adult female mouse brain. Jour- RFamide-related peptide expression are coordinated with
nal of Neuroendocrinology, 21, 673–682. the preovulatory luteinizing hormone surge. Endocrinology,
Clarkson, J., & Herbison, A. E. (2006). Postnatal development 149, 4958–4969.
of kisspeptin neurons in mouse hypothalamus; sexual dimor- Gillespie, J. M., Chan, B. P., Roy, D., Cai, F., &
phism and projections to gonadotropin-releasing hormone Belsham, D. D. (2003). Expression of circadian rhythm
neurons. Endocrinology, 147, 5817–5825. genes in gonadotropin-releasing hormone-secreting GT1-7
d’Anglemont de Tassigny, X., Fagg, L. A., Carlton, M. B., & neurons. Endocrinology, 144, 5285–5292.
Colledge, W. H. (2008). Kisspeptin can stimulate gonadotro- Gottsch, M. L., Cunningham, M. J., Smith, J. T., Popa, S. M.,
pin-releasing hormone (GnRH) release by a direct action at Acohido, B. V., Crowley, W. F., et al. (2004). A role for
GnRH nerve terminals. Endocrinology, 149, 3926–3932. kisspeptins in the regulation of gonadotropin secretion in
d’Anglemont de Tassigny, X., Fagg, L. A., Dixon, J. P., the mouse. Endocrinology, 145, 4073–4077.
Day, K., Leitch, H. G., Hendrick, A. G., et al. (2007). Greives, T. J., Kriegsfeld, L. J., & Demas, G. E. (2008). Exog-
Hypogonadotropic hypogonadism in mice lacking a func- enous kisspeptin does not alter photoperiod-induced
tional Kiss1 gene. Proceedings of the National Academy of gonadal regression in Siberian hamsters (Phodopus sun-
Sciences of the United States of America, 104, 10714–10719. gorus). General and Comparative Endocrinology, 156,
Dardente, H., Klosen, P., Pevet, P., & Masson-Pevet, M. 552–558.
(2003). MT1 melatonin receptor mRNA expressing cells in Hahn, J. D., & Coen, C. W. (2006). Comparative study of the
the pars tuberalis of the European hamster: Effect of photo- sources of neuronal projections to the site of gonadotrophin-
period. Journal of Neuroendocrinology, 15, 778–786. releasing hormone perikarya and to the anteroventral per-
Dardente, H., Wyse, C. A., Birnie, M. J., Dupre, S. M., iventricular nucleus in female rats. The Journal of Compar-
Loudon, A. S., Lincoln, G. A., et al. (2010). A molecular ative Neurology, 494, 190–214.
switch for photoperiod responsiveness in mammals. Current Han, S. K., Gottsch, M. L., Lee, K. J., Popa, S. M., Smith, J. T.,
Biology, 20, 2193–2198. Jakawich, S. K., et al. (2005). Activation of gonadotropin-
de la Iglesia, H. O., & Schwartz, W. J. (2006). Minireview: releasing hormone neurons by kisspeptin as a neuroendo-
Timely ovulation: Circadian regulation of the female crine switch for the onset of puberty. Journal of Neurosci-
hypothalamo-pituitary-gonadal axis. Endocrinology, 147, ence, 25, 11349–11356.
1148–1153. Hanon, E. A., Lincoln, G. A., Fustin, J. M., Dardente, H.,
de Roux, N., Genin, E., Carel, J. C., Matsuda, F., Masson-Pevet, M., Morgan, P. J., et al. (2008). Ancestral
Chaussain, J. L., & Milgrom, E. (2003). Hypogonadotropic TSH mechanism signals summer in a photoperiodic mam-
hypogonadism due to loss of function of the KiSS1-derived mal. Current Biology, 18, 1147–1152.
peptide receptor GPR54. Proceedings of the National Acad- Hanon, E. A., Routledge, K., Dardente, H., Masson-Pevet, M.,
emy of Sciences of the United States of America, 100, Morgan, P. J., & Hazlerigg, D. G. (2010). Effect of photope-
10972–10976. riod on the thyroid-stimulating hormone neuroendocrine
435

system in the European hamster (Cricetus cricetus). Journal the brains of mammals. Proceedings of the National Acad-
of Neuroendocrinology, 22, 51–55. emy of Sciences of the United States of America, 103,
Hinuma, S., Shintani, Y., Fukusumi, S., Iijima, N., 2410–2415.
Matsumoto, Y., Hosoya, M., et al. (2000). New neuropeptides Li, Q., Rao, A., Pereira, A., Clarke, I. J., & Smith, J. T. (2011).
containing carboxy-terminal RFamide and their receptor in Kisspeptin cells in the ovine arcuate nucleus express prolac-
mammals. Nature Cell Biology, 2, 703–708. tin receptor but not melatonin receptor. Journal of Neuroen-
Hoffman, R. A., & Reiter, R. J. (1965). Pineal gland: Influence docrinology, 23, 871–882.
on gonads of male hamsters. Science, 148, 1609–1611. Malpaux, B., Daveau, A., Maurice, F., Gayrard, V., &
Irwig, M. S., Fraley, G. S., Smith, J. T., Acohido, B. V., Thiery, J. C. (1993). Short-day effects of melatonin on
Popa, S. M., Cunningham, M. J., et al. (2004). Kisspeptin luteinizing hormone secretion in the ewe: Evidence for cen-
activation of gonadotropin releasing hormone neurons and tral sites of action in the mediobasal hypothalamus. Biology
regulation of KiSS-1 mRNA in the male rat. Neuroendocri- of Reproduction, 48, 752–760.
nology, 80, 264–272. Mason, A. O., Duffy, S., Zhao, S., Ubuka, T., Bentley, G. E.,
Johnson, M. A., Tsutsui, K., & Fraley, G. S. (2007). Rat Tsutsui, K., et al. (2010). Photoperiod and reproductive con-
RFamide-related peptide-3 stimulates GH secretion, inhibits dition are associated with changes in RFamide-related pep-
LH secretion, and has variable effects on sex behavior in the tide (RFRP) expression in Syrian hamsters (Mesocricetus
adult male rat. Hormones and Behavior, 51, 171–180. auratus). Journal of Biological Rhythms, 25, 176–185.
Kadokawa, H., Shibata, M., Tanaka, Y., Kojima, T., Mason, A. O., Greives, T. J., Scotti, M. A., Levine, J.,
Matsumoto, K., Oshima, K., et al. (2009). Bovine C-terminal Frommeyer, S., Ketterson, E. D., et al. (2007). Suppression
octapeptide of RFamide-related peptide-3 suppresses of kisspeptin expression and gonadotropic axis sensitivity
luteinizing hormone (LH) secretion from the pituitary as well following exposure to inhibitory day lengths in female
as pulsatile LH secretion in bovines. Domestic Animal Endo- Siberian hamsters. Hormones and Behavior, 52, 492–498.
crinology, 36, 219–224. Masson-Pevet, M., & Gauer, F. (1994). Seasonality and mela-
Kalsbeek, A., Teclemariam-Mesbah, R., Pévet, P. (1993). tonin receptors in the pars tuberalis in some long day
Efferent projections of the suprachiasmatic nucleus in the breeders. Biological Signals, 3, 63–70.
golden hamster (Mesocricetus auratus). The Journal of Com- Masson-Pevet, M., George, D., Kalsbeek, A., Saboureau, M.,
parative Neurology, 332, 293–314. Lakhdar-Ghazal, N., & Pevet, P. (1994). An attempt to cor-
Kalsbeek, A., Palm, I. F., La Fleur, S. E., Scheer, F. A., relate brain areas containing melatonin-binding sites with
Perreau-Lenz, S., Ruiter, M., et al. (2006). SCN outputs rhythmic functions: A study in five hibernator species. Cell
and the hypothalamic balance of life. Journal of Biological and Tissue Research, 278, 97–106.
Rhythms, 21, 458–469. Mayer, C., & Boehm, U. (2011). Female reproductive matura-
Karsch, F. J., Dahl, G. E., Hachigian, T. M., & Thrun, L. A. tion in the absence of kisspeptin/GPR54 signaling. Nature
(1995). Involvement of thyroid hormones in seasonal repro- Neuroscience, 14, 704–710.
duction. Journal of Reproduction and Fertility. Supplement, Maywood, E. S., Bittman, E. L., & Hastings, M. H. (1996).
49, 409–422. Lesions of the melatonin- and androgen-responsive tissue
Khan, A. R., & Kauffman, A. S. (2012). The role of kisspeptin of the dorsomedial nucleus of the hypothalamus block the
and RFRP-3 neurons in the circadian-timed preovulatory gonadal response of male Syrian hamsters to programmed
luteinizing hormone surge. Journal of Neuroendocrinology, infusions of melatonin. Biology of Reproduction, 54,
24, 131–143. 470–477.
Kinoshita, M., Tsukamura, H., Adachi, S., Matsui, H., Maywood, E. S., & Hastings, M. H. (1995). Lesions of the
Uenoyama, Y., Iwata, K., et al. (2005). Involvement of cen- iodomelatonin-binding sites of the mediobasal hypothala-
tral metastin in the regulation of preovulatory luteinizing mus spare the lactotropic, but block the gonadotropic
hormone surge and estrous cyclicity in female rats. Endocri- response of male Syrian hamsters to short photoperiod and
nology, 146, 4431–4436. to melatonin. Endocrinology, 136, 144–153.
Kotani, M., Detheux, M., Vandenbogaerde, A., Messager, S., Chatzidaki, E. E., Ma, D., Hendrick, A. G.,
Communi, D., Vanderwinden, J. M., Le Poul, E., et al. Zahn, D., Dixon, J., et al. (2005). Kisspeptin directly
(2001). The metastasis suppressor gene KiSS-1 encodes stimulates gonadotropin-releasing hormone release via G
kisspeptins, the natural ligands of the orphan G protein- protein-coupled receptor 54. Proceedings of the National
coupled receptor GPR54. Journal of Biological Chemistry, Academy of Sciences of the United States of America, 102,
276, 34631–34636. 1761–1766.
Kriegsfeld, L. J., Mei, D. F., Bentley, G. E., Ubuka, T., Mikkelsen, J. D., Bentsen, A. H., Ansel, L., Simonneaux, V.,
Mason, A. O., Inoue, K., et al. (2006). Identification & Juul, A. (2009). Comparison of the effects of peripherally
and characterization of a gonadotropin-inhibitory system in administered kisspeptins. Regulatory Peptides, 152, 95–100.
436

Mikkelsen, J. D., & Simonneaux, V. (2009). The neuroanat- Ratajczak, C. K., Boehle, K. L., & Muglia, L. J. (2009).
omy of the kisspeptin system in the mammalian brain. Impaired steroidogenesis and implantation failure in
Peptides, 30, 26–33. Bmal1/ mice. Endocrinology, 150, 1879–1885.
Miller, B. H., Olson, S. L., Turek, F. W., Levine, J. E., Revel, F. G., Ansel, L., Klosen, P., Saboureau, M., Pevet, P.,
Horton, T. H., & Takahashi, J. S. (2004). Circadian clock Mikkelsen, J. D., et al. (2007). Kisspeptin: A key link to sea-
mutation disrupts estrous cyclicity and maintenance of preg- sonal breeding. Reviews in Endocrine & Metabolic Dis-
nancy. Current Biology, 14, 1367–1373. orders, 8, 57–65.
Molnar, C. S., Kallo, I., Liposits, Z., & Hrabovszky, E. (2011). Revel, F. G., Saboureau, M., Masson-Pevet, M., Pevet, P.,
Estradiol down-regulates RF-amide-related peptide (RFRP) Mikkelsen, J. D., & Simonneaux, V. (2006a). Kisspeptin
expression in the mouse hypothalamus. Endocrinology, 152, mediates the photoperiodic control of reproduction in
1684–1690. hamsters. Current Biology, 16, 1730–1735.
Nakao, N., Ono, H., Yamamura, T., Anraku, T., Takagi, T., Revel, F. G., Saboureau, M., Pevet, P., Mikkelsen, J. D., &
Higashi, K., et al. (2008). Thyrotrophin in the pars tuberalis Simonneaux, V. (2006b). Melatonin regulates type 2 deiodinase
triggers photoperiodic response. Nature, 452, 317–322. gene expression in the Syrian hamster. Endocrinology, 147,
Palm, I. F., Van Der Beek, E. M., Wiegant, V. M., 4680–4687.
Buijs, R. M., & Kalsbeek, A. (1999). Vasopressin induces Revel, F. G., Saboureau, M., Pevet, P., Simonneaux, V., &
a luteinizing hormone surge in ovariectomized, estradiol- Mikkelsen, J. D. (2008). RFamide-related peptide gene is a
treated rats with lesions of the suprachiasmatic nucleus. melatonin-driven photoperiodic gene. Endocrinology, 149,
Neuroscience, 93, 659–666. 902–912.
Pilorz, V., & Steinlechner, S. (2008). Low reproductive success Robertson, J. L., Clifton, D. K., de la Iglesia, H. O.,
in Per1 and Per2 mutant mouse females due to accelerated Steiner, R. A., & Kauffman, A. S. (2009). Circadian regula-
ageing? Reproduction, 135, 559–568. tion of Kiss1 neurons: Implications for timing the preovula-
Pineda, R., Aguilar, E., Pinilla, L., & Tena-Sempere, M. tory gonadotropin-releasing hormone/luteinizing hormone
(2010a). Physiological roles of the kisspeptin/GPR54 system surge. Endocrinology, 150, 3664–3671.
in the neuroendocrine control of reproduction. Progress in Seminara, S. B., Messager, S., Chatzidaki, E. E.,
Brain Research, 181, 55–77. Thresher, R. R., Acierno, J. S., Jr., Shagoury, J. K., et al.
Pineda, R., Garcia-Galiano, D., Sanchez-Garrido, M. A., (2003). The GPR54 gene as a regulator of puberty. The
Romero, M., Ruiz-Pino, F., Aguilar, E., et al. (2010b). New England Journal of Medicine, 349, 1614–1627.
Characterization of the inhibitory roles of RFRP3, the mam- Shinohara, K., Honma, S., Katsuno, Y., Abe, H., & Honma, K.
malian ortholog of GnIH, in the control of gonadotropin (1994). Circadian rhythms in the release of vasoactive intes-
secretion in the rat: In vivo and in vitro studies. American tinal polypeptide and arginine-vasopressin in organotypic
Journal of Physiology, Endocrinology and Metabolism, slice culture of rat suprachiasmatic nucleus. Neuroscience
299, E39–E46. Letters, 170, 183–186.
Pitrosky, B., & Pevet, P. (1997). The photoperiodic response Shupnik, M. A. (1996). Gonadal hormone feedback on pitui-
in Syrian hamsters depends upon a melatonin-driven rhythm tary gonadotropin genes. Trends in Endocrinology and
of sensitivity to melatonin. Biological Signals, 6, 264–271. Metabolism, 7, 272–276.
Prevot, V., Cornea, A., Mungenast, A., Smiley, G., & Simonneaux, V., Ansel, L., Revel, F. G., Klosen, P., Pevet, P.,
Ojeda, S. R. (2003). Activation of erbB-1 signaling in & Mikkelsen, J. D. (2009). Kisspeptin and the seasonal con-
tanycytes of the median eminence stimulates transforming trol of reproduction in hamsters. Peptides, 30, 146–153.
growth factor beta1 release via prostaglandin E2 production Simonneaux, V., & Ribelayga, C. (2003). Generation of the
and induces cell plasticity. Journal of Neuroscience, 23, melatonin endocrine message in mammals: A review of
10622–10632. the complex regulation of melatonin synthesis by norepi-
Prevot, V., Dehouck, B., Poulain, P., Beauvillain, J. C., nephrine, peptides, and other pineal transmitters. Pharma-
Buee-Scherrer, V., & Bouret, S. (2007). Neuronal- cological Reviews, 55, 325–395.
glial-endothelial interactions and cell plasticity in the post- Smith, J. T., Clay, C. M., Caraty, A., & Clarke, I. J. (2007).
natal hypothalamus: Implications for the neuroendocrine KiSS-1 messenger ribonucleic acid expression in the hypo-
control of reproduction. Psychoneuroendocrinology, 32 thalamus of the ewe is regulated by sex steroids and season.
(Suppl 1), S46–S51. Endocrinology, 148, 1150–1157.
Quennell, J. H., Rizwan, M. Z., Relf, H. L., & Smith, J. T., Clifton, D. K., & Steiner, R. A. (2006). Regulation
Anderson, G. M. (2010). Developmental and steroidogenic of the neuroendocrine reproductive axis by kisspeptin-
effects on the gene expression of RFamide related peptides GPR54 signaling. Reproduction, 131, 623–630.
and their receptor in the rat brain and pituitary gland. Jour- Smith, J. T., Coolen, L. M., Kriegsfeld, L. J., Sari, I. P.,
nal of Neuroendocrinology, 22, 309–316. Jaafarzadehshirazi, M. R., Maltby, M., et al. (2008).
437

Variation in kisspeptin and RFamide-related peptide photoperiodic regulation of reproduction. Endocrinology,


(RFRP) expression and terminal connections to gonadotro- 145, 1546–1549.
pin-releasing hormone neurons in the brain: A novel Wiegand, S. J., Terasawa, E., Bridson, W. E., & Goy, R. W.
medium for seasonal breeding in the sheep. Endocrinology, (1980). Effects of discrete lesions of preoptic and
149, 5770–5782. suprachiasmatic structures in the female rat. Alterations in
Smith, J. T., Cunningham, M. J., Rissman, E. F., Clifton, D. K., the feedback regulation of gonadotropin secretion. Neuro-
& Steiner, R. A. (2005a). Regulation of Kiss1 gene expres- endocrinology, 31, 147–157.
sion in the brain of the female mouse. Endocrinology, 146, Williams, W. P., 3rd, Jarjisian, S. G., Mikkelsen, J. D., &
3686–3692. Kriegsfeld, L. J. (2011). Circadian control of kisspeptin and
Smith, J. T., Dungan, H. M., Stoll, E. A., Gottsch, M. L., a gated GnRH response mediate the preovulatory
Braun, R. E., Eacker, S. M., et al. (2005b). Differential reg- luteinizing hormone surge. Endocrinology, 152, 595–606.
ulation of KiSS-1 mRNA expression by sex steroids in the Wintermantel, T. M., Campbell, R. E., Porteous, R., Bock, D.,
brain of the male mouse. Endocrinology, 146, 2976–2984. Grone, H. J., Todman, M. G., et al. (2006). Definition of
Smith, J. T., Saleh, S. N., & Clarke, I. J. (2009). Seasonal and estrogen receptor pathway critical for estrogen positive
cyclical change in the luteinizing hormone response to feedback to gonadotropin-releasing hormone neurons and
kisspeptin in the ewe. Neuroendocrinology, 90, 283–291. fertility. Neuron, 52, 271–280.
Stetson, M. H., & Watson-Whitmyre, M. (1976). Nucleus Witkin, J. W., Paden, C. M., & Silverman, A. J. (1982). The
suprachiasmaticus: The biological clock in the hamster? Sci- luteinizing hormone-releasing hormone (LHRH) systems
ence, 191, 197–199. in the rat brain. Neuroendocrinology, 35, 429–438.
Thompson, E. L., Patterson, M., Murphy, K. G., Smith, K. L., Xu, Z., Kaga, S., Tsubomizu, J., Fujisaki, J., Mochiduki, A.,
Dhillo, W. S., Todd, J. F., et al. (2004). Central and periph- Sakai, T., et al. (2011). Circadian transcriptional factor
eral administration of kisspeptin-10 stimulates the hypotha- DBP regulates expression of Kiss1 in the anteroventral per-
lamic-pituitary-gonadal axis. Journal of Neuroendocrinology, iventricular nucleus. Molecular and Cellular Endocrinology,
16, 850–858. 339, 90–97.
Ukena, K., Iwakoshi, E., Minakata, H., & Tsutsui, K. (2002). Yamamura, T., Yasuo, S., Hirunagi, K., Ebihara, S., &
A novel rat hypothalamic RFamide-related peptide Yoshimura, T. (2006). T(3) implantation mimics photoperi-
identified by immunoaffinity chromatography and mass odically reduced encasement of nerve terminals by glial pro-
spectrometry. FEBS Letters, 512, 255–258. cesses in the median eminence of Japanese quail. Cell and
Ukena, K., & Tsutsui, K. (2001). Distribution of novel Tissue Research, 324, 175–179.
RFamide-related peptide-like immunoreactivity in the Yasuo, S., Nakao, N., Ohkura, S., Iigo, M., Hagiwara, S.,
mouse central nervous system. Neuroscience Letters, 300, Goto, A., et al. (2006). Long-day suppressed expression of
153–156. type 2 deiodinase gene in the mediobasal hypothalamus of
Ukena, K., & Tsutsui, K. (2005). A new member of the hypo- the Saanen goat, a short-day breeder: Implication for sea-
thalamic RF-amide peptide family, LPXRF-amide peptides: sonal window of thyroid hormone action on reproductive
Structure, localization, and function. Mass Spectrometry neuroendocrine axis. Endocrinology, 147, 432–440.
Reviews, 24, 469–486. Yasuo, S., Yoshimura, T., Ebihara, S., & Korf, H. W. (2009).
Urbanski, H. F., Doan, A., & Pierce, M. (1991). Immunocyto- Melatonin transmits photoperiodic signals through the MT1
chemical investigation of luteinizing hormone-releasing hor- melatonin receptor. Journal of Neuroscience, 29, 2885–2889.
mone neurons in Syrian hamsters maintained under long or Yoshida, H., Habata, Y., Hosoya, M., Kawamata, Y.,
short days. Biology of Reproduction, 44, 687–692. Kitada, C., & Hinuma, S. (2003). Molecular properties of
Vida, B., Deli, L., Hrabovszky, E., Kalamatianos, T., endogenous RFamide-related peptide-3 and its interaction
Caraty, A., Coen, C. W., et al. (2010). Evidence for with receptors. Biochimica et Biophysica Acta (BBA).
suprachiasmatic vasopressin neurones innervating kisspeptin Molecular Cell Research, 1593, 151–157.
neurones in the rostral periventricular area of the mouse Yoshimura, T., Yasuo, S., Watanabe, M., Iigo, M.,
brain: Regulation by oestrogen. Journal of Neuroendocrinol- Yamamura, T., Hirunagi, K., et al. (2003). Light-induced
ogy, 22, 1032–1039. hormone conversion of T4 to T3 regulates photoperiodic
Watanabe, M., Yasuo, S., Watanabe, T., Yamamura, T., response of gonads in birds. Nature, 426, 178–181.
Nakao, N., Ebihara, S., et al. (2004). Photoperiodic Zhao, S., & Kriegsfeld, L. J. (2009). Daily changes in GT1-7
regulation of type 2 deiodinase gene in Djungarian hamster: cell sensitivity to GnRH secretagogues that trigger ovula-
Possible homologies between avian and mammalian tion. Neuroendocrinology, 89, 448–457.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 25

Circannual rhythm in the varied carpet beetle,


Anthrenus verbasci

Yosuke Miyazaki{, Tomoyosi Nisimura{ and Hideharu Numata},*

{
Faculty of Clinical Education, Ashiya University, Hyogo, Japan
{
College of Bioresource Sciences, Nihon University, Fujisawa, Japan
}
Graduate School of Science, Kyoto University, Kyoto, Japan

Abstract: Although circannual rhythms controlling different physiological processes and various aspects
of behavior have been reported in numerous organisms, our understanding of the underlying biological
mechanisms is still quite limited. We examined the mechanisms controlling the circannual pupation
rhythm of the varied carpet beetle, Anthrenus verbasci. This rhythm is self-sustainable, exhibits
temperature compensation of the periodicity, and is entrainable to environmental changes. In addition,
the circannual phase response curves to a photoperiod pulse display Type 0 or Type 1 resetting,
depending on the duration of the pulse. Thus, we infer that this rhythm is derived from a self-sustaining
biological oscillator with a period of about a year, that is, a circannual clock, analogous to the circadian
clock. Further, a circadian clock appears to mediate photoperiodic time measurement for phase resetting
of the circannual clock. Based on these results and previous research performed in other organisms, we
discuss the general characteristics of the physiological mechanisms underpinning circannual rhythmicity.

Keywords: Anthrenus verbasci; circannual clock; circannual rhythm; circadian clock; entrainment; phase
response curve; photoperiodism; temperature compensation; zeitgeber.

Introduction to enhance the rate of survival and reproduction.


When isolated from external temporal cues,
Most organisms have evolved endogenous rhythms these biological rhythms free-run with an inherent
corresponding to various environmental cycles, period close to, but significantly different from,
including daily cycles, tidal cycles, and annual cycles, that of the environmental cycle. Appropriate envi-
ronmental cues, that is, zeitgebers, can entrain
*Corresponding author. these endogenous rhythms to the environmental
Tel.: þ81-75-753-4073; Fax: þ81-75-753-4113 cycle (DeCoursey, 2004; Pittendrigh, 1981a;
E-mail: numata@ethol.zool.kyoto-u.ac.jp Saunders, 2002). The circadian rhythm, which has

http://dx.doi.org/10.1016/B978-0-444-59427-3.00025-3 439
440

an endogenous period close to 24 h and entrains to environmental cycle and advance and delay phase
daily cycles of light and temperature, has been stud- shifts induced by the stimulus (see also Gwinner,
ied by many researchers in a variety of organisms. 1986; Saunders, 2002). The importance of Blake’s
Chronobiologists have started to uncover the work is comparable to the many studies of
mechanisms behind the biological clocks that gener- circannual rhythms carried out later in other
ate rhythmicity, and today, the basic principles of cir- animals but has not received the attention it
cadian clocks are understood in detail (Johnson deserves (Saunders, 2002).
et al., 2004; Rosbash, 2009; Saunders, 2002; Stillman Since 1996, we have been examining the pupa-
et al., 2007). However, compared to circadian tion rhythm in a Japanese population of A. verbasci.
rhythms, the sources of other endogenous rhythms In Japan, the duration of the life cycle in this species
are only partly understood, owing to a considerable is generally considered to be 1 year. In some cases,
lack of research. In particular, it is unclear whether however, 2 or 3 years are required to complete
the mechanism that generates internal periodicity is the larval development (Kiritani, 1958; Miyazaki
a self-sustaining biological oscillator similar to the et al., 2009a). We reevaluated Blake’s results and
circadian clock. introduced various new approaches based on our
A biological rhythm with an endogenous period current knowledge of circadian and circannual
close to a year is called a circannual rhythm. rhythms and of photoperiodism. We found that
The circannual rhythms of various physiological the circannual rhythm of A. verbasci is generated
processes and behavioral traits have been reported by a circannual clock analogous to the circadian
in a variety of organisms, including dinoflagellates, clock in its phase response (Miyazaki et al., 2005,
brown algae, higher plants, coelenterates, annelids, 2007; Nisimura and Numata, 2001) and that the
mollusks, crustaceans, insects, fish, reptiles, birds, circannual rhythm, with a period of about 40 weeks,
and mammals (Goldman et al., 2004; Gwinner, 1986; is entrained to an exact year by naturally changing
Olive and Garwood, 1983). In vertebrates, Pengelley photoperiods (Miyazaki et al., 2006; Nisimura
and Fisher (1957, 1963) were the first to conclusively and Numata, 2003). Moreover, the circannual clock
demonstrate the circannual rhythm of the golden- is independent of the circadian clock (Nisimura
mantled ground squirrel, Spermophilus lateralis. and Numata, 2002), although a circadian clock is
The seasonal changes in body weight and food con- involved in the time measurement system for phase
sumption, and the occurrence of hibernation, were resetting of the circannual clock in response to
shown to have a period shorter than 12 months when photoperiodic changes (Miyazaki et al., 2009b).
squirrels were maintained for 2 years under constant Menaker (1974) stated that the major difficulty
photoperiod (12-h light and 12-h darkness, LD 12:12) in the study of circannual rhythms is the ratio of
and temperature. the period length of a single circannual cycle to
In invertebrates, Blake (1958, 1959) first the length of the productive life of the biologist.
reported that a circannual rhythm regulates pupa- Despite this difficulty, a number of researchers
tion of the varied carpet beetle, Anthrenus verbasci have studied circannual rhythms in a variety
(Insecta, Coleoptera, Dermestidae), which nor- of species (Goldman et al., 2004; Gwinner,
mally has a 2-year life cycle in southern England. 1986, 2003; Lincoln and Hazlerigg, 2010; Paul
Moreover, she was the first to show the following et al., 2008; Wikelski et al., 2008). In this review,
major properties of the circannual rhythm (Blake, based on our results obtained in A. verbasci and
1958, 1959, 1960, 1963): (1) a self-sustaining oscilla- those from studies on various other organisms,
tion with a period deviating significantly from that we discuss the general characteristics of the
of the environmental cycle, (2) temperature com- physiological mechanism producing circannual
pensation of the period length, and (3) the exis- rhythmicity and its responsiveness to photope-
tence of a zeitgeber entraining the rhythm to the riodic changes.
441

Characteristics of the circannual rhythm A third pupation group was observed 89–109
weeks after hatching. The interval between the
Self-sustainability medians of the first and second groups was 37
weeks and that of the second and third groups
Biological rhythms exhibit self-sustainability was 37.5 weeks. This periodicity can be explained
when examined under constant conditions with- by the gate concept, which was originally
out input from environmental cues (Johnson introduced for the allowed time for eclosion in
et al., 2004; Pittendrigh, 1981a; Saunders, 2002). the circadian rhythm of flies (Pittendrigh, 1966).
A constant 24-h light–dark cycle has been used Thus, the existence of an endogenous circannual
routinely to detect circannual rhythms because rhythm in the process of pupation was demonstra-
unchanging photoperiods contain no information ted in a Japanese population of A. verbasci
about the duration of a year (Gwinner, 1986). (Nisimura and Numata, 2001). We repeated the
Figure 1a shows the frequency of pupation times experiments under LD 12:12 at 20  C from various
of A. verbasci when larvae were reared under times of the year and observed a similar rhythmic-
LD 12:12 at 20  C. These insects exhibited rhyth- ity, although the third pupation group has not been
mic pupation. Larvae pupated 23–31 weeks after clearly observed in many cases (Miyazaki et al.,
hatching as the first pupation group. The second 2005, 2009b; Nisimura and Numata, 2001, 2003).
group pupated 49–74 weeks after hatching. The timing of pupation in the first and second
cycles was negligibly affected by slowing larval
development with a diet of low-nutrient food,
80 (a)
although the percentage of pupae in each cycle
changed remarkably (Miyazaki et al., 2009a).
60 It has been reported that, for some species,
circannual rhythms persist under constant light
or darkness. For example, the circannual hiberna-
40
tion rhythm of S. lateralis persists with a period
shorter than 12 months, irrespective of whether
Number of pupae

20 the lighting conditions are constant darkness, con-


stant light, or LD 12:12 (Pengelley et al., 1976; see
also Gwinner, 1986). Figure 1b shows the fre-
0
60 (b) quency of pupation times of A. verbasci when lar-
vae were continuously maintained under constant
dim light. The periodic pattern was less clear than
40
that under LD 12:12 (Nisimura and Numata,
2001). When we first obtained these results, we
20 thought that this indistinct rhythm resulted from
the damping of the circannual rhythm, as observed
for the circadian eclosion rhythm under constant
0 light in flies (Pittendrigh, 1966; Saunders, 2002).
0 20 40 60 80 100 120 However, later experiments revealed that the
Weeks after hatching circannual rhythm of A. verbasci clearly persists
under constant light after preexposure to LD 12:12
Fig. 1. Pupation of Anthrenus verbasci under LD 12:12 (a) or
constant dim light of 0.006 Wm 2 intensity (b) at 20  C. The
for 8 weeks. The pattern of pupation was similar to
triangle indicates the median of each pupation group. that under continuous LD 12:12 (Fig. 2; Miyazaki
Modified from Nisimura and Numata (2001). and Numata, 2010). The circannual rhythmicity
442

50 (a) The circannual period under constant dim light in


the dark-eyed junco, Junco hyemalis, ranges from
25
6 to 20 months (mean, 13.7 months) in testicular
0 development and from 6 to 21 months (mean,
Number of pupae

50 (b) 15.0 months) in the onset of postnuptial molts


(Holberton and Able, 1992). These deviations
25 from 12 months, of about  50% and þ 75% in
0
extreme cases, represent considerable variations,
50 (c)
compared to deviations from 24 h of approxi-
mately  13% (21 h) and þ 17% (28 h) in circadian
25 free-running periods (Aschoff, 1979). In addition,
some species display circannual rhythms only
0
under certain photoperiodic conditions (Gwinner,
0 20 40 60 80 100 1986, 1996; Paul et al., 2008; Schaffelke and
Weeks after hatching Lüning, 1994). For example, the European star-
ling, Sturnus vulgaris, shows a clear circannual
Fig. 2. The effect of preexposure to LD 12:12 on pupation of
Anthrenus verbasci under constant light of 0.9 Wm 2
rhythm in testicular development under LD 12:12
intensity at 20  C. Larvae were maintained continuously but not under LD 11:13 or LD 13:11 (Schwab,
under LD 12:12 (a) or constant light (b), or transferred from 1971). The circannual rhythm of A. verbasci was
LD 12:12 to constant light at 8 weeks after hatching (c). obscure under LD 15:9 and LD 16:8 and arrhyth-
Filled bars indicate the duration of LD 12:12 exposure. The mic under LD 14:10 (Nisimura and Numata,
triangle indicates the median of each pupation group.
Modified from Miyazaki and Numata (2010). 2003). The influence of photoperiod on the
circannual period has also been reported
(Gwinner, 1986 for review). In the rainbow trout,
observed under constant light indicates that daily Oncorhynchus mykiss, the circannual period is sig-
light–dark cycles are not necessary to sustain this nificantly shorter under LD 18:6 and constant light
endogenous rhythm, as observed in a British popu- than under LD 6:18 (Duston and Bromage, 1986).
lation under constant darkness (Blake, 1958, The circannual period of A. verbasci was 5 weeks
1959). As described later, the change in the photo- longer under LD 13:11 than under LD 12:12
period is a zeitgeber for this circannual rhythm. (Nisimura and Numata, 2003). Under LD 10:14 in
It is likely that the circannual rhythm of A. verbasci one experiment, the rhythm persisted with a much
oscillates under constant light, but the initial longer period of 52 weeks (Nisimura and Numata,
circannual phase of newly hatched larvae varies 2002), but in another experiment, pupation after
among individuals. Therefore, for a circannual the first peak was asynchronous and the period
rhythm to be clearly observed under constant light, length could not be determined (Miyazaki and
it is necessary to synchronize the phase of the larvae Numata, 2009). Although circannual rhythms in
by subjecting them to 24-h photoperiodic cycles. some species, including the African stonechat,
The circannual period of A. verbasci is 8–10 Saxicola torquatus, and the Siberian chipmunk,
months (Blake, 1959; Miyazaki et al., 2009b; Tamias sibiricus, have been reported to persist for
Nisimura and Numata, 2001). The circannual per- approximately 10 years (Gwinner, 1996; Kondo
iods under constant conditions are shorter than 1 et al., 2006), circannual rhythms in most organisms
year in many species, although longer periods tend to dampen within several cycles under constant
have also been reported (Gwinner, 1986). Individ- conditions in the laboratory (Gwinner, 1986).
ual circannual periods often show great variability, In general, both inter- and intraindividual variability
even within a single species (Gwinner, 1986). in the endogenous period in the same species is
443

considerably larger for most circannual rhythms These results suggest that the circannual period
than for circadian rhythms. Thus, Gwinner (1986) is, at least to a certain extent, temperature com-
concluded that circannual rhythms have weaker pensated (Gwinner, 1986).
self-sustainability than circadian rhythms. More conspicuous examples of temperature-
compensated circannual rhythms are present in
poikilotherms (Gwinner, 1986). In our studies
Temperature compensation of A. verbasci, when larvae were exposed under
LD 12:12, pupation began 21–27 weeks after hatch-
Biological oscillators are based on intricate bio- ing at constant temperatures between 17.5  C and
chemical mechanisms. Although most biochemical 27.5  C, and the second group pupated about 40
processes depend on temperature, temperature weeks after the first. High temperatures did not
compensation is required for accurate time result in earlier pupation as might be expected
measurement in biological rhythms (Johnson from observations of normal developmental pro-
et al., 2004; Pittendrigh, 1981a; Saunders, 2002). cesses. In fact, the first pupation peaks were slightly
Aperiodic fluctuations in temperature perturb delayed by higher temperature (Fig. 3; Nisimura
time measurement and need to be eliminated by and Numata, 2001). We concluded that the timing
buffering mechanisms, that is, compensation of pupation displayed temperature compensation,
mechanisms. In fact, period length is almost fully in line with the results of Blake (1958, 1959).
temperature compensated in circadian rhythms
under steady-state conditions at various constant
temperatures (Balzer and Hardeland, 1988; 20
27.5 °C
Saunders, 2002; Sweeney and Hastings, 1960). In 10
biological rhythms with longer periods, however,
0
the demonstration of temperature compensation 20
is sometimes difficult. First, it is difficult to main- 25 °C
tain strict constant conditions for a long duration, 10
and second, a desired constant temperature under
Number of pupae

0
experimental conditions often interferes with the 20
22.5 °C
regulation of seasonal physiological events (Balzer
10
and Hardeland, 1988; Gwinner, 1986). Neverthe-
less, temperature compensation has been reported 0
in some circannual rhythms (Gwinner, 1986). 20
20 °C
Temperature compensation was shown in the 10
circannual rhythm period of a hibernating homeo-
0
therm, S. lateralis, which spends many months 20
each year in deep torpor with a body temperature 17.5 °C
close to that of its environment. In this animal, 10
although the period of the first circannual cycle 0
during hibernation is slightly longer at 3  C than
at 12  C, this difference disappears in the second 0 20 40 60 80 100

and third cycles (Pengelley and Asmundson, Weeks after hatching


1969). Much larger differences would have been Fig. 3. Pupation of Anthrenus verbasci under LD 12:12 at
expected if the processes underlying circannual various constant temperatures. The triangle indicates the
rhythms showed a temperature dependence simi- median of each pupation group. Modified from Nisimura and
lar to that of many other physiological processes. Numata (2001).
444

Temperature compensation of the circannual When we exposed A. verbasci larvae to natural


period has also been shown in various other annual changes in photoperiod in Japan, at a con-
invertebrates, for example, oviposition of the slug, stant 20  C, pupation peaks were observed in Jan-
Limax flavus (10–20  C; Segal, 1960; see also uary or February in both the first and second
Gwinner, 1986); growth and development of cycles (Fig. 4a; Miyazaki et al., 2006; Nisimura
the hydroid, Campanularia flexuosa (10–24  C;
Brock, 1975a,b); and sexual maturation of the
polychaete, Nereis diversicolor (5–10  C; Olive and (a)
16
Garwood, 1983). Thus, in invertebrates, compensa- 10
40
tion mechanisms appear to modulate time measure-
ment in circannual rhythms under conditions of 0
(b)
changing environmental temperature. 16
40 10
Entrainability
0
(c)
Biological rhythms can be entrained by zeitgebers 16

Number of pupae
from environmental cycles, thereby maintaining a 40 10

Day length (h)


stable phase relationship to environmental cycles
0
(Johnson et al., 2004; Pittendrigh, 1981a; Saunders, (d)
2002). In many species, a change in photoperiod is 16
the dominant zeitgeber for entraining the circannual 40 10
rhythm to an annual cycle (Goldman et al., 2004; 0
Gwinner, 1986; Paul et al., 2008). For example, the (e)
16
circannual rhythms were entrained to sinusoidal
40 10
photoperiodic cycles when the zeitgeber periods
were 4–24 months and 2–12 months in antler 0
replacement of the sika deer, Cervus nippon, and (f)
16
in testicular development of S. vulgaris, respectively
40 10
(Goss, 1969; Gwinner, 1977; see also Gwinner,
1986). The circannual rhythms of growth in two kelp 0
species, Pterygophora californica and Laminaria J J A S ON D J FM A M J J A S ON D J FM A M

hyperborea, were also entrained to sinusoidal pho-


Fig. 4. The effect of a 4-week long-day pulse on pupation of
toperiodic cycles with a period of 3–12 months Anthrenus verbasci under naturally changing day length at
(Lüning and Kadel, 1993; Schaffelke and Lüning, 20  C. Larvae maintained under naturally changing day
1994). Although the endogenous period of cir- length were exposed to LD 16:8 for 4 weeks starting on
cannual rhythms in many species often deviates August 4 (b), September 1 (c), September 29 (d), October 27
(e), or November 24 (f). Nonexposed control larvae are
considerably from 12 months, entrainment to the
shown in (a). The triangle indicates the median of each
natural year is uneventfully accomplished by a pupation group. A vertical dotted line crosses through the
range of entrainment proportionally much larger median of each pupation group in the control experiment
than that for circadian rhythms (Goldman et al., (a). The solid curve in each panel indicates the natural day
2004; Gwinner, 1986). Gwinner (1986) suggested length, including 1 h of twilight in Osaka, Japan (35 N), for
that at least some circannual rhythms with weaker most of the experiment, and the artificial day length
produced by white fluorescent lamps during the LD 16:8
self-sustainability were more dependent on environ- pulse. The shaded bar in (a) shows the critical photoperiodic
mental zeitgebers and more easily affected by them range for determining whether the day length is long or
than circadian rhythms. short. Modified from Miyazaki et al. (2006).
445

and Numata, 2003), similar to the observations changes act as an effective zeitgeber for the
by Blake (1960) of A. verbasci populations in circannual rhythm of A. verbasci, as suggested
England. Therefore, a circannual pupation by Blake (1960).
rhythm in A. verbasci, displaying a periodicity of
about 40 weeks under LD 12:12, is entrained to
exactly 1 year by exposure to natural changes in Three different hypotheses for circannual rhythm
photoperiod. generation
Blake (1963) suggested that the first and sec-
ond cycles of the pupation rhythm in A. verbasci Circannual clock hypothesis
were controlled differently. She argued that, in
the first cycle, pupation was advanced by increas- A simple hypothesis to explain the generation of
ing day length but that, in the second cycle, the circannual rhythm assumes the existence of a
pupation was inhibited by decreasing day length. self-sustaining biological oscillator with a period
However, our results obtained by altering contin- of about a year, that is, a circannual clock,
uous short-day (LD 12:12) and long-day (LD analogous to the circadian clock (Gwinner, 1981a;
16:8) cycles at various time points indicate that Mrosovsky, 1978; Pengelley and Asmundson,
whether pupation is advanced or delayed depends 1974). However, the most troublesome feature of
on the phase at which the photoperiodic stimulus this hypothesis is the long-term nature of the oscil-
is applied, rather than the cycle of the circannual lator. Many animals exhibiting circannual rhythms,
rhythm (Miyazaki et al., 2005, 2007; Nisimura including hibernating mammals and migratory
and Numata, 2001). Phase-dependent shifts in birds, have short life spans of a few years
rhythm are important factors in entrainment and (Gwinner, 1986), and therefore it is difficult to
have been observed in many different circadian examine the circannual clock hypothesis in these
rhythms. This topic will be explored in detail organisms. Then, the other two hypotheses that
later. do not involve a circannual clock, as mentioned
Compared to the photoperiod, the role of non- below, may have arisen. However, for A. verbasci,
photoperiodic cues has not been sufficiently our data suggest that the pupation rhythm is gener-
examined, although seasonal changes in tempera- ated by a circannual clock similar to the circadian
ture, daytime light intensity, and social stimuli clock, because the phase response curves (PRCs)
have been suggested as potential zeitgebers in obtained in this circannual rhythm share many
various species (Goldman et al., 2004; Gwinner, similar features with PRCs obtained in circadian
1986, 2003; Paul et al., 2008). In many cases, rhythms (Miyazaki et al., 2005, 2007; and see later
nonphotoperiodic cues act by directly driving sections for detail).
or suppressing expression of the trait without
entrainment of the endogenous rhythm (Paul
et al., 2008). When A. verbasci larvae are exposed Frequency demultiplication hypothesis
to natural fluctuations in light and temperature,
they pupate each year in April, rather than in Gwinner (1973) proposed the frequency demulti-
January and February (Miyazaki et al., 2009a; plication hypothesis, which states that circannual
Nisimura and Numata, 2003). It is likely that rhythms are derived from circadian rhythms
low temperatures between December and March through a process of frequency demultiplication.
suppress pupation until April and that this In other words, organisms generate circannual
suppression also contributes to the synchrony rhythms by transforming the circadian periodicity
of pupation in the field. However, there of about a day to one cycle of about a year. The
remains the possibility that natural temperature possibility that circadian rhythms are involved in
446

the production of circannual rhythms is appeal- 30


ing, because circadian rhythms are a ubiquitous

Larval duration (weeks)


feature of animal physiology. Moreover, Gwinner
(1973) observed a positive correlation between
circadian and circannual periods under constant 25
dim light in S. vulgaris.
The frequency demultiplication hypothesis
requires that the period of the circannual rhythm
is proportional to the period of the entrained circa- 20
dian rhythms. Therefore, rigorous tests of this
hypothesis can be performed by exposing 0
20 22 24 26
individuals to light–dark cycles of different periods Period of the light–dark
(T). We reared A. verbasci larvae under different cycle (h)
constant photoperiods, ranging from T ¼ 20 to
26 h (Nisimura and Numata, 2002). Because the Fig. 5. Evaluation of the frequency demultiplication
hypothesis for the circannual pupation rhythm in Anthrenus
range of entrainment in circadian rhythms is usu- verbasci. Larvae were kept under LD 10:10 (T ¼ 20), 10:14
ally between 18 and 30 h (Aschoff and Pohl, (T ¼ 24), and 10:16 (T ¼ 26) at 20  C. Open circles represent
1978), the circadian rhythms of the larvae should medians of the first pupation group. Vertical lines represent
entrain to the LD 10:10, LD 10:14, and LD 10:16 interquartiles. Closed circles and a broken line represent the
photoperiods. Based on the result under LD timing of the first pupation group predicted by the frequency
demultiplication hypothesis and the results under LD 10:14.
10:14 (T ¼ 24), in which the first group of larvae Based on the data of Nisimura and Numata (2002).
pupated 25 weeks after hatching, the frequency
demultiplication hypothesis predicts that pupation
should occur 21 and 27 weeks after hatching under Sequence of linked stages hypothesis
LD 10:10 (T ¼ 20) and LD 10:16 (T ¼ 26), respec-
tively (Fig. 5). However, the first group of larvae There is the hypothesis that circannual rhythms
pupated 27 weeks after hatching under LD 10:10 merely result from a sequence of linked physio-
and 25 weeks after hatching under LD 10:16. logical stages, with each stage normally taking a
Thus, there is no positive correlation between given amount of time to complete and then lead-
the period of the light–dark cycle and the ing to the next, and so on (i.e., a chain of interval
timing of pupation in A. verbasci (Nisimura and timers). The last stage is linked back to the first
Numata, 2002). To date, similar experiments (Mrosovsky, 1978). Circannual rhythms in some
have been performed in three species of birds hibernators, for example, the edible dormouse,
and one mammal, but results supporting the fre- Glis glis; the little pocket mouse, Perognathus
quency demultiplication hypothesis have not been longimembris; and the 13-lined ground squirrel,
obtained (Carmichael and Zucker, 1986; Gwinner, Spermophilus tridecemlineatus, can be explained
1981b; Wikelski et al., 2008). by this hypothesis, but those in others cannot
Further, circannual rhythms are exhibited by (Mrosovsky, 1978).
most individuals even after disruption of circadian Mrosovsky (1974) suggested that a circannual
activity patterns by ablating the suprachiasmatic clock and a sequence of stages can be characterized
nucleus in mammals or the pineal gland in birds, as a pendulum-type oscillator and a relaxation
which are the locations of the circadian pacemaker oscillator, respectively. However, because the lat-
(e.g., Pant and Chandola-Saklani, 1992; Zucker ter is one-dimensional, the sequence of linked
et al., 1983). These results also do not support the stages hypothesis cannot explain some features of
frequency demultiplication hypothesis. phase resetting of rhythms (see Lakin-Thomas,
447

1995 for details; see also Oda et al., 2000), for larvae under LD 12:12 and exposed the larvae to
example, Type 0 resetting and arrhythmicity LD 16:8 for 4 weeks (4-week long-day pulse) dur-
induced by a zeitgeber pulse in the circannual ing various phases of the rhythm (Miyazaki et al.,
rhythm of A. verbasci (Miyazaki et al., 2005, 2005). When larvae were exposed to a 4-week
2007), which we discuss further in the following long-day pulse, 4 weeks after hatching, pupation
section. of the first and second groups was delayed by
9–9.5 weeks, compared to control insects (Fig. 7b).
However, when a 4-week long-day pulse was given
PRCs and the biological oscillator 16 weeks after hatching, pupation was advanced
by 2–3 weeks in both groups (Fig. 7c). Thus,
In circadian rhythms, the response to a zeitgeber whether a 4-week long-day pulse advances or delays
depends on the phase subjected to the zeitgeber. the phase of the circannual rhythm depends on the
PRCs are constructed to represent these phase- phase in which the pulse is given. Based on these
dependent responses. A circadian PRC is a plot observations, we constructed a circannual PRC
of the magnitude of phase shifts (phase advances (Fig. 6c) where the period of the rhythm under con-
and delays) obtained by giving a light pulse at vari- tinuous LD 12:12 (37 weeks; Fig. 7a) is shown in
ous phases of a free-running rhythm under con- terms of angle degrees (0–360 ). The initial phase
stant darkness (Johnson, 1999; Johnson et al., under LD 12:12, that is, the beginning of this exper-
2004; Pittendrigh, 1981b). A light pulse in early iment, is represented as 180 . The range 0–180 is
subjective night generally delays a phase, whereas considered subjective summer and 180–360 is con-
a light pulse in late subjective night advances a sidered subjective winter. This circannual PRC to
phase. A pulse applied in subjective day has a little 4-week long-day pulses closely resembles the Type
or no effect on the phase. The strength of a light 0 PRC of circadian rhythms (Fig. 6a and c).
pulse, that is, intensity, duration, or both, often We also constructed a circannual PRC to
changes the amplitude of phase shifts in the PRC. 2-week long-day pulses (Fig. 6d). The phase shifts
Strong pulses produce Type 0 PRCs with large are relatively smaller than those caused by a
phase shifts and a distinct break point at the transi- 4-week long-day pulse, and a PRC to 2-week
tion between delays and advances, whereas weak long-day pulses has a continuous transition
pulses produce Type 1 PRCs with small phase between delays and advances in the middle of
shifts and a continuous transition between phase the subjective winter. Therefore, this curve is cat-
delays and advances (Fig. 6a and b). egorized as Type 1 (Miyazaki et al., 2007).
In contrast to circadian rhythms, reports of In circadian rhythms, the two types of phase
PRCs to zeitgeber pulses in circannual rhythms resetting, dependent on the strength of the zeitge-
are uncommon (Paul et al., 2008). In the ber stimulus, can be theoretically explained by
circannual spawning rhythm of the rainbow trout, the concept that a circadian clock is a biological
O. mykiss, a pulse of constant light for 2 months oscillator that has two or more state variables.
induced a phase-dependent phase shift. However, A number of different oscillator types have been
because the treatment was performed on the proposed, including an adjustable-amplitude oscil-
rhythm entrained to naturally changing day lator, a limit cycle oscillator, and a multioscillator
length, the response curve obtained did not satisfy system (Johnson et al., 2004; Lakin-Thomas,
the definition of the PRC, in which a free-running 1995; Winfree, 2000). It is known that arrhy-
rhythm is perturbed under constant conditions thmicity of a circadian rhythm can be evoked by
(Randall et al., 1998). a single light pulse of a certain strength delivered
To construct a complete PRC to photoperiod near the middle of the subjective night. This phe-
pulses for a circannual rhythm, we kept A. verbasci nomenon can be explained by the notion that the
448

(a) (b)
Subjective Subjective Subjective Subjective
day night day night
180

Phase shift (degree)


90

-90

-180

(c) (d)
Subjective Subjective Subjective Subjective
summer winter summer winter
180
Phase shift (degree)

90

-90

-180
0 90 180 270 360 0 90 180 270 360
Phase of pulse onset (degree) Phase of pulse onset (degree)

Fig. 6. Comparison of phase response curves for circadian and circannual rhythms. (a, b) Phase response curves in circadian
rhythms, Type 0 (a) and Type 1 (b). (c, d) Phase response curves in the circannual pupation rhythm of Anthrenus verbasci, a
curve to 4-week long-day pulses (c) and a curve to 2-week long-day pulses (d). Larvae were kept under LD 12:12 at 20  C and
exposed to LD 16:8 for 4 (c) or 2 (d) weeks at various phases in the circannual rhythm. The circannual period under continuous
LD 12:12 is shown in terms of angle degrees (0–360 ). Open and closed circles represent the phase shifts in the first and second
pupation group after pulse perturbation, respectively. Broken lines in (c) show the split into advanced and delayed groups. (c)
From Miyazaki et al. (2005). (d) From Miyazaki et al. (2007).

oscillator is driven to the phaseless singularity in be a biological oscillator with two or more state
the phase space or that the phases of individual variables with circannual variation (Johnson
oscillators of the multioscillator system are et al., 2004; Lakin-Thomas, 1995). This implies that
scattered (Johnson et al., 2004; Lakin-Thomas, the circannual rhythm of A. verbasci is generated
1995; Winfree, 1970, 2000). In the circannual by a circannual clock specialized for adaptation
rhythm of A. verbasci, arrhythmicity of pupation to annual cycles, much like the circadian clock
can also be evoked by a 4-week long-day pulse and its adaptation to daily cycles.
administered in the middle of the subjective winter Recently, a circannual PRC for reproduction in
(Miyazaki et al., 2007). According to the theoreti- the European hamster, Cricetus cricetus, was con-
cal explanations for Type 1 and Type 0 phase structed by exposing hamsters, normally maintained
resetting and the loss of rhythmicity induced under long days (LD 16:8), to short days (LD 10:14)
by a zeitgeber pulse, the mechanism behind for 1 month (Monecke et al., 2009). The shape of the
the circannual rhythm of A. verbasci is thought to PRC was nearly identical to the circannual PRC to
449

100 cycles seems to be applicable to many diverse


(a)
organisms.
80

60
Photoperiodism for entrainment
40
Entrainment under natural day length
20
The circannual rhythm of A. verbasci shows endog-
0 enous periodicity of about 40 weeks under continu-
100
(b) ous LD 12:12 but is entrained to a strictly annual
80 period by changes in photoperiod. This entrain-
Number of pupae

ment can be explained as a phase delay induced


60 by a decrease in day length in late summer or early
autumn (Nisimura and Numata, 2001, 2003). How-
40 ever, it is necessary to know the exact temporal
changes in the phase of the circannual clock under
20
natural day length to understand the biological
0
entrainment mechanisms operating under natural
80 conditions.
(c)
If we compare the phase shifts induced by the
60 photoperiod pulse under natural photoperiod
with those shown in the circannual PRC, we can
40
describe the temporal change in phase under nat-
20
ural day length in terms of the phase in the PRC.
Therefore, we examined the phase responses to
0 4-week long-day pulses given between August 4
0 20 40 60 80 100 and November 24 under natural day length at a
Weeks after hatching
constant temperature of 20  C (Miyazaki et al.,
2006). These pulses caused a phase shift in the
Fig. 7. The effect of a 4-week long-day pulse on pupation of first pupation group, as under LD 12:12. A long-
Anthrenus verbasci under LD 12:12 at 20  C. Larvae day pulse given on August 4, September 1, or
maintained under LD 12:12 (filled bars) were exposed to LD
September 29 caused a phase delay, and a pulse
16:8 (rectangle bars) for 4 weeks, commencing 4 weeks after
hatching (b) or 16 weeks after hatching (c). Nonexposed given on October 27 or November 24 caused a
control larvae are shown in (a). The triangle indicates the phase advance. Pupation was least synchronous
median of each pupation group. A vertical dotted line just before the transition from delay to advance
crosses through the median of each pupation group in the (Fig. 4). These phase responses were similar to
control experiment (a). Modified from Miyazaki et al. (2005).
those in the subjective winter of a circannual
PRC to 4-week long-day pulses under LD 12:12
LD 16:8 for 4 weeks in A. verbasci, a species taxo- (Fig. 6c). Therefore, the circannual clock changes
nomically quite distant from C. cricetus. Therefore, its phase at least from early August to late
these phase responses may signify a fundamental November under natural day length, as occurs in
and common characteristic of circannual rhythms, the subjective winter under LD 12:12. In the sec-
and the concept that the circannual rhythm is pro- ond pupation group, however, larvae pupated at
duced by a circannual clock specialized for annual the same time as control larvae not subjected to
450

a long-day pulse, regardless of the phase response hormone that mediates the effects of the photope-
of the first group (Fig. 4). This result can be riod. The infusion was designed to simulate sea-
explained by entrainment to the geophysical year son-specific melatonin secretion. Infusion during
by long days during spring and summer. the summer was most effective in entrainment to
We also examined the range of photoperiodic annual cycles and maintained the proper relation-
changes effective for circannual phase shifts in ship between the phase of the rhythm and the
A. verbasci by exposing larvae to various longer season during which the infusion was provided.
photophases for 4 weeks against a background The response and phase setting of circannual
of LD 12:12 or LD 10:14 (Miyazaki and Numata, rhythms to photoperiodic changes likely vary,
2009). Phase advances were smaller than depending on the species or subspecies (Gwinner,
phase delays, and the magnitude of the change 1986; Helm et al., 2009). To better understand the
depended on the amplitude of the photoperiodic full diversity of responses of circannual rhythms
changes rather than the absolute photophase to photoperiodic changes, further research will
duration. In contrast, a clear phase delay was be required in a wide variety of species.
induced when the photoperiodic change exceeded
a critical value in the photophase between 13 and
14 h, regardless of the amplitude of the change. Involvement of a circadian clock
It is likely that the response in phase delay is more
important in the seasonal timing of pupation under A wide variety of species have evolved photoperi-
natural conditions because the endogenous period odism without relation to circannual rhythms.
of this rhythm is considerably shorter than 1 year. In such photoperiodism, the photoperiodic time
In photoperiodic responses of many species, an measurement system generally involves a circa-
important feature is the existence of a critical day dian clock (Goldman et al., 2004; Nelson et al.,
length above or below which behavioral and phys- 2010; Saunders, 2002). In phase responses in the
iological responses are observed (Goldman et al., circannual rhythm of A. verbasci, photoperiodic
2004; Nelson et al., 2010; Saunders, 2002). To time measurement is also necessary, although
entrain a circannual period to the natural annual unlike regular photoperiodism, the effects of pho-
cycle, A. verbasci in central Japan probably uses toperiod are indirect and are due to resetting of
a critical value in the photophase between 13 and the circannual rhythm. The Nanda–Hamner pro-
14 h for determining whether the day length is long tocol has frequently been used to clarify the
or short. It is thought that stable entrainment to involvement of a circadian system in photope-
the natural year is established by giving long-term riodic responses. In this protocol, groups of
exposure to a photophase longer than 13 h during organisms are subjected to light–dark cycles with
spring and summer, or during late subjective sum- a fixed short photophase followed by a variable
mer or early subjective winter of the circannual scotophase to give cycle lengths (T) of up to
rhythm, where phase delay responses to long days 72 h or more. If organisms show short-day
are exhibited in the PRC (Figs. 4a and 6c). responses when T is a multiple of 24 h, but not
An important role of long days in summer has when T is not a multiple of 24 h, the response is
also been suggested in entrainment of the considered positive, thus indicating the involve-
circannual rhythm in luteinizing hormone secre- ment of a circadian system in photoperiodic time
tion of the sheep, Ovis aries (Woodfill et al., measurement.
1994). In pinealectomized ewes, which do not The Nanda–Hamner protocol has been used to
respond to changes of day length, photoperiodic examine the reproductive activities of several ver-
information was applied during one of the four tebrate species exhibiting circannual rhythms. In
seasons each year via infusion of melatonin, a O. mykiss, the effects of constant 48-h, 54-h, and
451

60-h photoperiods were examined over 2 years, LD 12:12


50
but the circadian nature of the photoperiodic
responses was obscure (Duston and Bromage, 0
1986). Positive responses to the Nanda–Hamner LD 12:24
50
protocol were demonstrated for birds and
mammals (e.g., Almeida and Lincoln, 1982; 0
Gwinner and Eriksson, 1977), although these LD 12:36
50
observations were conducted over a period of less
than a year.

Number of pupae
0
We examined the effects of exposure to LD 12:48
50
the Nanda–Hamner protocol (LD 12:12, LD
12:24, LD 12:36, LD 12:48, and LD 12:60) on 0
the circannual rhythm of A. verbasci to examine LD 12:60
50
whether the circadian system is involved in
photoperiodic time measurement for entrainment 0
(Miyazaki et al., 2009b). Figure 8 shows the Continuous darkness
50
results when larvae were exposed to various
photoperiods for 120 days and then transferred 0
to LD 12:12. When larvae were exposed to LD LD 16:8
50
12:36 or LD 12:60, for which T is a multiple of
24 h, the pupation pattern was similar to that 0
under continuous LD 12:12. In contrast, under
LD 12:24 (T ¼ 36 h), the pupation pattern was 0 20 40 60 80 100
similar to that under LD 16:8, being considerably Weeks after hatching
delayed compared to that under LD 12:12. Under
Fig. 8. The effect of exposure to the Nanda–Hamner protocol for
LD 12:48 (T ¼ 60 h), pupation was also delayed 120 days on pupation of Anthrenus verbasci at 20  C. Larvae were
and was less synchronous than that of the first transferred to LD 12:12 after exposure to various photoperiods.
group under LD 12:60 (T ¼ 72 h) or constant Shaded areas represent exposure to various photoperiods. The
darkness. Thus, photoperiods for which T was triangle indicates the median of each pupation group. Modified
from Miyazaki et al. (2009b).
not a multiple of 24 h (i.e., LD 12:24 and LD
12:48) produced a large phase delay compared
to LD 12:12. These results clearly indicate that temperature compensation of the period, and
the circannual rhythm of A. verbasci shows a pos- entrainability to a zeitgeber (Nisimura and
itive Nanda–Hamner response in phase resetting Numata, 2001), all of which are key characteristics
to photoperiod. Therefore, a circadian clock is of a clock mechanism linked to environmental
most likely involved in photoperiodic entrainment cycles (Johnson et al., 2004; Pittendrigh, 1981a;
of this rhythm, although it is not responsible for Saunders, 2002). The characteristics of this rhythm
generating the circannual rhythm, as described indicate that the frequency demultiplication
above. hypothesis does not accurately describe the under-
lying mechanism (Nisimura and Numata, 2002).
The circannual PRCs were obtained by giving
Conclusions and future directions long-day pulses, and the shapes of the curves
closely resembled those of circadian PRCs to light
In the circannual pupation rhythm of A. verbasci, pulses (Miyazaki et al., 2005, 2007). Similarities to
we confirmed self-sustainability (i.e., persistence), circadian rhythms in the resetting patterns indicate
452

that the mechanism producing circannual rhyth- organisms, for example, O. aries (Lincoln and
micity in A. verbasci is a circannual clock analogous Hazlerigg, 2010; Lincoln et al., 2006).
to the circadian clock (Miyazaki et al., 2007). The neural and molecular mechanisms that
The circannual clock has most likely evolved to generate circannual rhythmicity remain unknown,
increase the chances of survival and reproductive not only for A. verbasci but also for other species
success, by allowing the organism to better pre- as well. In S. lateralis, various neural tissues and
dict and adapt to seasonal changes (DeCoursey, endocrine glands, including the suprachiasmatic
2004; Gwinner, 1981a, 1986; Pittendrigh, 1981a). nucleus and the pineal gland, have been ablated,
To permit accurate- and consistent-seasonal but the anatomical location of the circannual
responses, A. verbasci refers not only to ambient pacemaker is still unclear (Zucker, 2001). In
photoperiod but also to the phase of a circannual O. aries, however, Lincoln et al. (2006) suggested
clock (Miyazaki et al., 2006), and it refers to a cir- that the adenohypophysis is the site of the pace-
cadian clock to judge photoperiod (Miyazaki maker for the circannual rhythm of prolactin
et al., 2009b). Figure 9 shows a schematic diagram secretion. The adenohypophysis secrets prolactin
of the physiological system behind the circannual with circannual periodicity, without neural inputs
rhythm of A. verbasci. Thus, in the field, an from the brain, and is a target of pineal melatonin
annual rhythm of A. verbasci is established by a signals transmitting photoperiodic information.
combination of exogenous factors and endoge- Recently, a few models have been proposed to
nous physiological mechanisms comprising circa- generate the circannual rhythm in mammals.
dian and circannual clocks. It is likely that this In T. sibiricus, the complex of hibernation-specific
schema is also applicable to certain other proteins is secreted by the liver. Before hibernation,

Light

Photoreceptor
Phase reset

Circannual Circadian Circadian


Photoperiodic
clock clock clock
time
measurement
system
Phase
reset
Time-of-year Time-of-day
information information

Effector Effector
system system

Seasonal Daily
activity activity

Fig. 9. Schematic diagram of the physiological components of the circannual rhythm in Anthrenus verbasci. The relationship
between the circadian clock involved in photoperiodic time measurement and a circadian clock controlling daily activities
(dashed arrow) is unclear.
453

this complex is transported to the brain where it photoperiodism (Goto et al., 2010). Another
mediates the onset of hibernation (Kondo et al., method is provided by studies in birds. In avian
2006). For circannual rhythm generation in this spe- circannual rhythms, there are clear differences in
cies, Kondo (2007) proposed a model in which the degree of rhythmicity, timing of behavior, and
levels of this hibernation-specific protein complex responsiveness to photoperiod among populations
in the brain and periphery form a negative-feed- living in different areas. This geographic variation
back loop. In the circannual rhythm of O. aries, is clearly attributable to genetic differences
MacGregor and Lincoln (2008) proposed a physio- because crossbreeding produces an intermediate
logical model consisting of a negative-feedback response in the hybrid progeny (Gwinner, 1986,
mechanism comprising the components of the pars 1996; Helm et al., 2009). A. verbasci is a geograph-
tuberalis and pars distalis within the adenohypoph- ically widespread species, and therefore, it proba-
ysis. These two models postulates a negative-feed- bly also has genetic variation. The genetic
back loop formed by a number of different tissues. differences may help in elucidating the molecular
In contrast, Lincoln and Hazlerigg (2010) proposed mechanisms of circannual rhythmicity in this
that circannual rhythmicity is autonomously gener- organism.
ated by cyclical histogenesis in specific sites in A direct way to identify molecular candidates
various tissues, such as the brain, pituitary, and involved in the generation of circannual rhythmic-
periphery. ity is to use DNA microarrays, similar to
Because circannual rhythms have also been approaches used to identify candidates that play
observed in coelenterates (Brock, 1975a), it is a role in circadian rhythm generation (Akhtar
possible that animals can exhibit circannual rhyth- et al., 2002; McDonald and Rosbash, 2001).
micity without the need for a central nervous sys- Subsequently, the precise involvement of the can-
tem. Further, circannual rhythms are observed didates in the circannual rhythm may be revealed
even in eukaryotic unicellular organisms, and by RNA interference (RNAi). In beetles, inhibi-
therefore, molecular-feedback loops within a sin- tion of gene expression in various physiological
gle cell, similar to those in a circadian clock, responses has been performed by RNAi (e.g.,
may be involved in generating the circannual Niimi et al., 2005; Tomoyasu and Denell, 2004).
rhythm (Anderson and Keafer, 1987; Matrai Therefore, we propose DNA microarrays and
et al., 2005). If circannual rhythms have evolved RNAi as the next step to clarify the biological
independently several times and are of heteroge- mechanism behind the circannual rhythm in
neous origin in different groups of organisms A. verbasci.
(Farner, 1970; Gwinner, 1981a, 1986), entirely dif- Since Konopka and Benzer (1971) first demon-
ferent mechanisms may drive the circannual strated that a gene, called period, regulates the cir-
rhythm in different species. cadian behavior of the fruit fly, Drosophila
Several approaches are available to uncover the melanogaster, enormous advances in our knowl-
mechanism behind circannual rhythmicity in edge of the genetic and molecular aspects of circa-
A. verbasci. Because a photoperiodic response dian rhythmicity have occurred. A clear picture
involving a circadian system resets the phase of a of the underlying molecular-feedback loops is
circannual clock in this species (Fig. 9), one way is emerging for numerous organisms (Rosbash,
to identify the anatomical location of a circannual 2009; Saunders, 2002). We anticipate that our
clock by reference to the centers for photoperiod- knowledge of the physiological and molecular
ism and circadian rhythmicity. One may proceed mechanisms behind circannual rhythmicity will
by studying the expression and regulation of circa- similarly advance greatly when researchers begin
dian clock genes, similar to the approach used to to isolate the molecular components of this funda-
unravel the molecular mechanisms of insect mental biological process.
454

References Goldman, B., Gwinner, E., Karsch, F. J., Saunders, D.,


Zucker, I., & Gall, G. F. (2004). Circannual rhythms and
Akhtar, R. A., Reddy, A. B., Maywood, E. S., Clayton, J. D., photoperiodism. In J. C. Dunlap, J. J. Loros & P. J.
King, V. M., Smith, A. G., et al. (2002). Circadian cycling of DeCoursey (Eds.), Chronobiology—Biological timekeeping
the mouse liver transcriptome, as revealed by cDNA micro- (pp. 107–142). Sunderland: Sinauer Associates.
array, is driven by the suprachiasmatic nucleus. Current Goss, R. J. (1969). Photoperiodic control of antler cycles in
Biology, 12, 540–550. deer. I. Phase shift and frequency changes. The Journal of
Almeida, O. F., & Lincoln, G. A. (1982). Photoperiodic regu- Experimental Zoology, 170, 311–324.
lation of reproductive activity in the ram: Evidence for the Goto, S. G., Shiga, S., & Numata, H. (2010). Photoperiodism
involvement of circadian rhythms in melatonin and prolactin in insects: Perception of light and the role of clock genes.
secretion. Biology of Reproduction, 27, 1062–1075. In R. J. Nelson, D. L. Denlinger & D. E. Somers (Eds.),
Anderson, D. M., & Keafer, B. A. (1987). An endogenous Photoperiodism: The biological calendar (pp. 258–286).
annual clock in the toxic marine dinoflagellate Gonyaulax New York: Oxford University Press.
tamarensis. Nature, 325, 616–617. Gwinner, E. (1973). Circannual rhythms in birds: Their interac-
Aschoff, J. (1979). Circadian rhythms: Influences of internal tion with circadian rhythms and environmental photoperiod.
and external factors on the period measured in constant Journal of Reproduction and Fertility. Supplement, 19, 51–65.
conditions. Zeitschrift für Tierpsychologie, 49, 225–249. Gwinner, E. (1977). Photoperiodic synchronization of
Aschoff, J., & Pohl, H. (1978). Phase relations between a circa- circannual rhythms in the European starling (Sturnus
dian rhythm and its zeitgeber within the range of entrain- vulgaris). Naturwissenschaften, 64, 44–45.
ment. Naturwissenschaften, 65, 80–84. Gwinner, E. (1981a). Circannual systems. In J. Aschoff (Ed.),
Balzer, I., & Hardeland, R. (1988). Influence of temperature Handbook of behavioral neurobiology. Biological rhythms
on biological rhythms. International Journal of Biometeorol- (Vol. 4), (pp. 391–410). New York: Plenum Press.
ogy, 32, 231–241. Gwinner, E. (1981b). Circannual rhythms: Their dependence
Blake, G. M. (1958). Diapause and the regulation of develop- on the circadian system. In B. K. Follett & D. E. Follett
ment in Anthrenus verbasci (L.) (Col., Dermestidae). Bulle- (Eds.), Biological clocks in seasonal reproductive cycles
tin of Entomological Research, 49, 751–775. (pp. 153–169). Bristol: John Wright & Sons.
Blake, G. M. (1959). Control of diapause by an ‘internal clock’ Gwinner, E. (1986). Circannual rhythms. Berlin: Springer-
in Anthrenus verbasci (L.) (Col., Dermestidae). Nature, 183, Verlag.
126–127. Gwinner, E. (1996). Circannual clocks in avian reproduction
Blake, G. M. (1960). Decreasing photoperiod inhibiting meta- and migration. Ibis, 138, 47–63.
morphosis in an insect. Nature, 188, 168–169. Gwinner, E. (2003). Circannual rhythms in birds. Current
Blake, G. M. (1963). Shortening of a diapause-controlled life cycle Opinion in Neurobiology, 13, 770–778.
by means of increasing photoperiod. Nature, 198, 462–463. Gwinner, E., & Eriksson, L. O. (1977). Circadiane Rhythmik
Brock, M. A. (1975a). Circannual rhythms-I. Free-running und photoperiodische Zeitmessung beim Star (Sturnus
rhythms in growth and development of the marine cnidar- vulgaris). Journal of Ornithology, 118, 60–67.
ian, Campanularia flexuosa. Comparative Biochemistry and Helm, B., Schwabl, I., & Gwinner, E. (2009). Circannual basis
Physiology, 51A, 377–383. of geographically distinct bird schedules. Journal of Experi-
Brock, M. A. (1975b). Circannual rhythms-II. Temperature- mental Biology, 212, 1259–1269.
compensated free-running rhythms in growth and develop- Holberton, R. L., & Able, K. P. (1992). Persistence of
ment of the marine cnidarian, Campanularia flexuosa. Com- circannual cycles in a migratory bird held in constant dim
parative Biochemistry and Physiology, 51A, 385–390. light. Journal of Comparative Physiology. A, 171, 477–481.
Carmichael, M. S., & Zucker, I. (1986). Circannual rhythms of Johnson, C. H. (1999). Forty years of PRCs—What have we
ground squirrels: A test of the frequency demultiplication learned? Chronobiology International, 16, 711–743.
hypothesis. Journal of Biological Rhythms, 1, 277–284. Johnson, C. H., Elliott, J., Foster, R., Honma, K. I., &
DeCoursey, P. J. (2004). The behavioral ecology and evolution Kronauer, R. (2004). Fundamental properties of circadian
of biological timing systems. In J. C. Dunlap, J. J. Loros & rhythms. In J. C. Dunlap, J. J. Loros & P. J. DeCoursey
P. J. DeCoursey (Eds.), Chronobiology—Biological time- (Eds.), Chronobiology—Biological timekeeping (pp. 67–105).
keeping (pp. 27–65). Sunderland: Sinauer Associates. Sunderland: Sinauer Associates.
Duston, J., & Bromage, N. (1986). Photoperiodic mechanisms Kiritani, K. (1958). Factors influencing the development of
and rhythms of reproduction in the female rainbow trout. Anthrenus verbasci L. Botyu-Kagaku, 23, 137–146 (In Japa-
Fish Physiology and Biochemistry, 2, 35–51. nese, abstract in English.).
Farner, D. S. (1970). Predictive functions in the control of Kondo, N. (2007). Endogenous circannual clock and
annual cycles. Environmental Research, 3, 119–131. HP complex in a hibernation control system. In: B. Stillman,
455

D. Stewart & T. Grodzicker (Eds.), Clocks and rhythms: Cold Miyazaki, Y., Nisimura, T., & Numata, H. (2009b). A circadian
Spring Harbor symposia on quantitative biology (Vol. 72), system is involved in photoperiodic entrainment of the
(pp. 607–613). Woodbury, NY: Cold Spring Harbor Labora- circannual rhythm of Anthrenus verbasci. Journal of Insect
tory Press. Physiology, 55, 494–498.
Kondo, N., Sekijima, T., Kondo, J., Takamatsu, N., Tohya, K., Miyazaki, Y., & Numata, H. (2009). Responsiveness to photo-
& Ohtsu, T. (2006). Circannual control of hibernation by periodic changes in the circannual rhythm of the varied car-
HP complex in the brain. Cell, 125, 161–172. pet beetle, Anthrenus verbasci. Journal of Comparative
Konopka, R. J., & Benzer, S. (1971). Clock mutants of Dro- Physiology. A, 195, 241–246.
sophila melanogaster. Proceedings of the National Academy Miyazaki, Y., & Numata, H. (2010). Exhibition of circannual
of Sciences of the United States of America, 68, 2112–2116. rhythm under constant light in the varied carpet beetle
Lakin-Thomas, P. L. (1995). A beginner’s guide to limit cycles, Anthrenus verbasci. Biological Rhythm Research, 41,
their uses and abuses. Biological Rhythm Research, 26, 441–448.
216–232. Monecke, S., Saboureau, M., Malan, A., Bonn, D., Masson-
Lincoln, G. A., Clarke, I. J., Hut, R. A., & Hazlerigg, D. G. Pévet, M., & Pévet, P. (2009). Circannual phase response
(2006). Characterizing a mammalian circannual pacemaker. curves to short and long photoperiod in the European ham-
Science, 314, 1941–1944. ster. Journal of Biological Rhythms, 24, 413–426.
Lincoln, G. A., & Hazlerigg, D. G. (2010). Mammalian Mrosovsky, N. (1974). Comment. In E. T. Pengelley (Ed.),
circannual pacemakers. In M. C. Lucy, J. L. Pate, M. F. Smith Circannual clocks: Annual biological rhythms
& T. E. Spencer (Eds.), Reproduction in domestic ruminants, (pp. 161–163). New York: Academic Press.
7, (pp. 171–186). Nottingham: Nottingham University Press. Mrosovsky, N. (1978). Circannual cycles in hibernators. In
Lüning, K., & Kadel, P. (1993). Daylength range for circannual L. C. H. Wang & J. W. Hudson (Eds.), Strategies in cold:
rhythmicity in Pterygophora californica (Alariaceae, Natural torpidity and thermogenesis (pp. 21–65). New York:
Phaeophyta) and synchronization of seasonal growth by Academic Press.
daylength cycles in several other brown algae. Phycologia, Nelson, R. J., Denlinger, D. L., & Somers, D. E. (Eds.),
32, 379–387. (2010). Photoperiodism: The biological calendar. New York:
MacGregor, D. J., & Lincoln, G. A. (2008). A physiological Oxford University Press.
model of a circannual oscillator. Journal of Biological Niimi, T., Kuwayama, H., & Yaginuma, T. (2005). Larval
Rhythms, 23, 252–264. RNAi applied to the analysis of postembryonic develop-
Matrai, P., Thompson, B., & Keller, M. (2005). Circannual ment in the ladybird beetle, Harmonia axyridis. Journal of
excystment of resting cysts of Alexandrium spp. from east- Insect Biotechnology and Sericology, 74, 95–102.
ern Gulf of Maine populations. Deep Sea Research Part II: Nisimura, T., & Numata, H. (2001). Endogenous timing mech-
Topical Studies in Oceanography, 52, 2560–2568. anism controlling the circannual pupation rhythm of the
McDonald, M. J., & Rosbash, M. (2001). Microarray analysis varied carpet beetle Anthrenus verbasci. Journal of Compar-
and organization of circadian gene expression in Drosoph- ative Physiology. A, 187, 433–440.
ila. Cell, 107, 567–578. Nisimura, T., & Numata, H. (2002). Evaluation of the fre-
Menaker, M. (1974). Circannual rhythms in circadian perspec- quency demultiplication hypothesis of circannual pupation
tive. In E. T. Pengelley (Ed.), Circannual clocks: Annual rhythm in the varied carpet beetle Antherenus verbasci
biological rhythms (pp. 507–518). New York: Academic Press. (Coleoptera: Dermestidae). Biological Rhythm Research,
Miyazaki, Y., Nisimura, T., & Numata, H. (2005). A phase 33, 255–260.
response curve for circannual rhythm in the varied carpet Nisimura, T., & Numata, H. (2003). Circannual control of the
beetle Anthrenus verbasci. Journal of Comparative Physiol- life cycle in the varied carpet beetle Anthrenus verbasci.
ogy. A, 191, 883–887. Functional Ecology, 17, 489–495.
Miyazaki, Y., Nisimura, T., & Numata, H. (2006). Phase Oda, G. A., Caldas, I. L., Piqueira, J. R. C., Waterhouse, J. M.,
responses in the circannual rhythm of the varied carpet bee- & Marques, M. D. (2000). Coupled biological oscillators in a
tle, Anthrenus verbasci, under naturally changing day length. cave insect. Journal of Theoretical Biology, 206, 515–524.
Zoological Science, 23, 1031–1037. Olive, P. J. W., & Garwood, P. R. (1983). The importance of
Miyazaki, Y., Nisimura, T., & Numata, H. (2007). Phase long term endogenous rhythms in the maintenance of repro-
resetting and phase singularity of an insect circannual oscilla- ductive cycles of marine invertebrates: A reappraisal. Inter-
tor. Journal of Comparative Physiology. A, 193, 1169–1176. national Journal of Invertebrate Reproduction, 6, 339–347.
Miyazaki, Y., Nisimura, T., & Numata, H. (2009a). Circannual Pant, K., & Chandola-Saklani, A. (1992). Pinealectomy and
pupation rhythm in the varied carpet beetle Anthrenus ver- LL abolished circadian perching rhythms but did not alter
basci under different nutrient conditions. Entomological Sci- circannual reproductive or fattening rhythms in finches.
ence, 12, 370–375. Chronobiology International, 9, 413–420.
456

Paul, M. J., Zucker, I., & Schwartz, W. J. (2008). Tracking the Schaffelke, B., & Lüning, K. (1994). A circannual rhythm con-
seasons: The internal calendars of vertebrates. Philosophical trols seasonal growth in the kelps Laminaria hyperborea and
Transactions of the Royal Society B, 363, 341–361. L. digitata from Helgoland (North Sea). European Journal
Pengelley, E. T., & Asmundson, S. M. (1969). Free-running of Phycology, 29, 49–56.
periods of endogenous circannian rhythms in the golden Schwab, R. G. (1971). Circannian testicular periodicity in the
mantled ground squirrel, Citellus lateralis. Comparative Bio- European starling in the absence of photoperiodic change.
chemistry and Physiology, 30, 177–183. In M. Menaker (Ed.), Biochronometry (pp. 428–447).
Pengelley, E. T., & Asmundson, S. J. (1974). Circannual rhyth- Washington DC: National Academy of Sciences.
micity in hibernating mammals. In E. T. Pengelley (Ed.), Segal, E. (1960). Discussion to the paper of A J Marshall. In: A.
Circannual clocks: Annual biological rhythms (pp. 95–160). Chovnick (Ed.), Biological clocks: Cold Spring Harbor
New York: Academic Press. symposia on quantitative biology (Vol. 25), (pp. 504–505).
Pengelley, E. T., Asmundson, S. J., Barnes, B., & Aloia, R. C. NY: Cold Spring Harbor: The Biological Laboratory.
(1976). Relationship of light intensity and photoperiod to Stillman, B., Stewart, D., & Grodzicker, T. (Eds.), (2007).
circannual rhythmicity in the hibernating ground squirrel, Clocks and rhythms: Cold Spring Harbor symposia on quan-
Citellus lateralis. Comparative Biochemistry and Physiology, titative biology (Vol. 72). Woodbury, NY: Cold Spring
53A, 273–277. Harbor Laboratory Press.
Pengelley, E. T., & Fisher, K. C. (1957). Onset and cessation Sweeney, B. M., & Hastings, J. W. (1960). Effects of tempera-
of hibernation under constant temperature and light in the ture upon diurnal rhythms. In: A. Chovnick (Ed.),
golden-mantled ground squirrel, Citellus lateralis. Nature, Biological clocks: Cold Spring Harbor symposia on quantita-
180, 1371–1372. tive biology (Vol. 25), (pp. 87–104). NY: Cold Spring
Pengelley, E. T., & Fisher, K. C. (1963). The effect of temper- Harbor: The Biological Laboratory.
ature and photoperiod on the yearly hibernating behavior of Tomoyasu, Y., & Denell, R. E. (2004). Larval RNAi in Tri-
captive golden-mantled ground squirrels (Citellus lateralis bolium (Coleoptera) for analyzing adult development.
tescorum). Canadian Journal of Zoology, 41, 1103–1120. Development Genes and Evolution, 214, 575–578.
Pittendrigh, C. S. (1966). The circadian oscillation in Drosoph- Wikelski, M., Martin, L. B., Scheuerlein, A., Robinson, M. T.,
ila pseudoobscura pupae: A model for the photoperiodic Robinson, N. D., Helm, B., et al. (2008). Avian circannual
clock. Zeitschrift für Pflanzenphysiologie, 54, 275–307. clocks: Adaptive significance and possible involvement of
Pittendrigh, C. S. (1981a). Circadian systems: General perspec- energy turnover in their proximate control. Philosophical
tive. In J. Aschoff (Ed.), Handbook of behavioral neurobiol- Transactions of the Royal Society B, 363, 411–423.
ogy. Biological rhythms (Vol. 4), (pp. 57–80). New York: Winfree, A. T. (1970). Integrated view of resetting a circadian
Plenum Press. clock. Journal of Theoretical Biology, 28, 327–374.
Pittendrigh, C. S. (1981b). Circadian systems: Entrainment. Winfree, A. T. (2000). The geometry of biological time (2nd
In J. Aschoff (Ed.), Handbook of behavioral neurobiology. ed.). New York: Springer-Verlag.
Biological rhythms (Vol. 4), (pp. 95–124). New York: Ple- Woodfill, C. J., Wayne, N. L., Moenter, S. M., & Karsch, F. J.
num Press. (1994). Photoperiodic synchronization of a circannual repro-
Randall, C. F., Bromage, N. R., Duston, J., & Symes, J. (1998). ductive rhythm in sheep: Identification of season-specific
Photoperiod-induced phase-shifts of the endogenous clock time cues. Biology of Reproduction, 50, 965–976.
controlling reproduction in the rainbow trout: A circannual Zucker, I. (2001). Circannual rhythms: Mammals. In J. S.
phase-response curve. Journal of Reproduction and Fertility, Takahashi, F. W. Turek & R. Y. Moore (Eds.), Handbook
112, 399–405. of behavioral neurobiology. Circadian Clocks (Vol. 12),
Rosbash, M. (2009). The implications of multiple circadian (pp. 509–528). New York: Kluwer Academic/Plenum.
clock origins. PLoS Biology, 7, e1000062. Zucker, I., Boshes, M., & Dark, J. (1983). Suprachiasmatic
Saunders, D. S. (2002). Insect clocks (3rd ed.). Amsterdam: nuclei influence circannual and circadian rhythms of ground
Elsevier. squirrels. American Journal of Physiology, 244, R472–R480.
A. Kalsbeek, M. Merrow, T. Roenneberg and R. G. Foster (Eds.)
Progress in Brain Research, Vol. 199
ISSN: 0079-6123
Copyright Ó 2012 Elsevier B.V. All rights reserved.

CHAPTER 26

Avian migration: Temporal multitasking and


a case study of melatonin cycles in waders

Barbara Helm{,{,},*, Eberhard Gwinner{,w, Anita Koolhaas}, Phil Battley||,


Ingrid Schwabl{, Anne Dekinga} and Theunis Piersma}

{
Max-Planck-Institut für Ornithologie, Andechs, Germany
{
Fachbereich Biologie, Universität Konstanz, Konstanz, Germany
}
Institute of Biodiversity, Animal Health and Comparative Medicine, University of
Glasgow, Glasgow, United Kingdom
}
Department of Marine Ecology, Royal Netherlands Institute for Sea Research (NIOZ), Den Burg, Texel,
The Netherlands
||
Ecology Group, Institute of Natural Resources, Massey University, Private Bag, Palmerston North, New Zealand

Abstract: Timing “in the real world” must cope with the temporal complexity of natural environments.
Extreme examples for the resultant “multitasking” are migratory birds, which precisely time movements
to remote areas. New field technologies highlight temporal accuracy, while captivity studies emphasize
underlying programs and plasticity of schedules. After reviewing these findings, we focus on waders,
which undertake spectacular long-distance migrations, have robust circannual clocks, and cope with diel,
tidal, and polar environments. To explore features that may facilitate such multitasking, we speculated
that melatonin amplitudes are low and damped during seasons when entrainment to subtle Zeitgebers
occurs. We measured melatonin profiles under European daylength in two species with different
ecologies and found low-amplitude melatonin cycles that changed over the year. Annual patterns neither
fully supported our hypothesis, nor simply reflected daylight availability. While migratory birds are
inspiring models for chronobiology, mechanistic understanding of their multitasking is still poor.

Keywords: clock; circannual; annual; tidal; circadian; moon; bird.

Temporal multitasking—Life in complex


*Corresponding author.
Tel.: þ497531884915; Fax: þ497531884916
temporal environments
E-mail: helm@orn.mpg.de
Chronobiology advances in great strides, as is
evident from this volume. Its predominant focus
w
Deceased

http://dx.doi.org/10.1016/B978-0-444-59427-3.00026-5 457
458

recently has been on circadian rhythms, owing to to accommodate temporal complexity. There
their near-ubiquitous occurrence and to rapidly is increasing evidence that, while clocks on differ-
developing molecular tools. While this work has ent time scales may rely on different mechanisms,
been immensely productive, we have gained com- there is cross talk, coordination, and mutual
paratively little insight on how biological time adjustment (e.g., Miyazaki et al., 2012; Tessmar-
keeping functions under the natural conditions in Raible et al., 2011). For example, fundamental
which it has evolved. This “real world” (Menaker, properties of circadian clocks, such as entrain-
2006) is structured by rhythmic processes on differ- ment mechanisms, deviation of the free-running
ent temporal scales, which are all of core interest to period from 24 h, and possibly the evolution of
chronobiology as a discipline concerned with “sys- its endogenous basis, are related to the annual
tematic study of adaptations to cope with regular change in daylength (Pittendrigh and Daan,
geophysical cycles” (Dunlap et al., 2004). Thus, its 1976). Further, circadian and circatidal rhythms
domains are based on rotations of earth and moon may interact in such a way that one suppresses
in relation to sun and comprise rhythms at the the other at certain phase angles (Palmer, 2000;
scales of solar year, solar day, lunar month, and Tessmar-Raible et al., 2011). Accordingly, study-
tidal cycle, as well as shorter rhythms contained ing clocks on different time scales could lead
within (e.g., ultradian rhythms). This wide scope to substantial advances in our understanding of
was emphasized early on (Chovnick, 1960; biological time keeping from an ecological as
Menaker, 1971), but time scales other than circa- well as mechanistic perspective. Based on such
dian have received little further attention. developments, there has been an increasing call
Organisms, however, cope with the different time to reactivate a broader perspective in chronobiol-
scales and their clocks must be constructed accord- ogy with the help of new tools from molecular
ingly. For example, daylength changes periodically and field biology (Hut et al., 2012; Miyazaki
over the year in most global habitats, and many et al., 2012; Tessmar-Raible et al., 2011).
species adjust daily activity patterns over the Our contribution uses a classical model system
annual cycle (Daan and Aschoff, 1975; Foster and of chronobiology, avian long-distance migration
Kreitzman, 2005, 2009; Foster and Roenneberg, (Aschoff, 1955; Gwinner, 1986; Hamner and Stock-
2008; Hut et al., 2012). Likewise, the periodic ing, 1970; Rowan, 1926), as an extreme example of
waxing and waning of the moon can have consider- temporal multitasking. We first provide an overview
able effects on activity patterns, in particular for of avian timing and then illustrate its complexity and
nocturnal species (Fernández-Duque et al., 2010) general characteristics in a taxon that stands out for
and in marine environments (Tessmar-Raible performance on different temporal scales, waders
et al., 2011). Further, cycles on different time scales (Order Charadriiformes). We complement the
can impart conflicting benefits and demands on review by empirical data from a study aimed at
scheduling. For example, in tidal habitats, bitidal identifying aspects of the circadian system that
cycles have a period length of 24.8 h, while a may enhance such multitasking. To this end, we
solar day takes 24 h. For some tidal organisms, examined diel profiles of the circadian hormone
therefore, important aspects of the environment, melatonin across the annual cycle in two species of
such as food availability or predation risk, may wader selected for contrasting temporal habits.
fluctuate with different periodicity (Palmer, 2000; Our overall intention is to promote thought about
Tessmar-Raible et al., 2011). timing as an integrated response to complex, tempo-
Organisms have internalized timing on differ- ral challenges. Although focused on birds, these
ent scales as endogenous circadian, circatidal, considerations are relevant for many kinds of
circalunar, or circannual rhythms (Gwinner, 1986; organisms, with the possible inclusion of our own
Tessmar-Raible et al., 2011), which must interact species for which cycles on other than daily scales
459

may be more relevant than sometimes acknowl- migrant experiences. In addition to light, other
edged (e.g., Foster and Kreitzman, 2009; Foster temporal aspects of habitats like daily patterns
and Roenneberg, 2008; Miyazaki et al., 2012; of food availability or predation risk may also
Tessmar-Raible et al., 2011; Wirz-Justice et al., 2001). be changed by migration, and in extreme cases,
timing can shift from diel to tidal schedules.
Migration also complicates the annual cycle by
Challenges of avian migration adding processes which birds undergo. Next to
migration periods per se, migrants add life-cycle
Birds as a taxon have much to offer with respect stages such as additional moult periods and pre-
to timing in the real world because their ecology paratory phases such as hyperphagia to build fuel
and evolutionary biology are relatively well stores (Newton, 2008; Wingfield, 2005). Further,
explored. Being conspicuous by sight and sound, migratory birds are a well-known model for close
the striking and sophisticated annual and daily links between time scales. Many species change
schedules can be readily observed and experi- their activity patterns during migration periods
mentally tested. On an annual scale, birds as a by assuming nocturnality in the wild and related,
long-lived taxon show regular, full cycles of spontaneous night activity (Zugunruhe) in captiv-
diverse activities (e.g., Gwinner, 1996b; Newton, ity (Gwinner, 1996a). This temporal plasticity is
2008). Within each annual cycle, almost all spe- of wide interest to chronobiology and sleep
cies undergo at least one bout of reproduction research (Mrosovsky and Hattar, 2005; Rattenborg
and regeneration (i.e., moult, the replacement of et al., 2004). Below, we give an overview of how
body plumage and flight feathers), respectively. birds meet the temporal challenges of migration.
There are countless variations of annual cycles.
For example, closely related taxa may differ
strikingly by complex patterns of additional, inter- Annual time keeping
mittent moults (Newton, 2008). In general, such
seasonal activities have to be well timed with The phenology (i.e., periodic seasonal recurrence)
respect to the environment and usually require of bird migration has been closely observed
anticipation of environmental fluctuations well in by humans from historical times until today
advance (Visser et al., 2010). Daily activities (Birkhead, 2008; Foster and Kreitzman, 2009).
are often “announced” by songs and calls so that For migrants, keeping track of the seasons is diffi-
species-specific timing (i.e., phase of activity) cult because of their exposure to widely separated
inspired the construction of “bird clocks” that global environments (Fig. 1). Their movements
illustrate its sequence. The outstanding mobility nonetheless often show such striking “punctuality”
of birds entails a need to keep track of time that they were used as calendrical information, for
of day, for example, for the use of a time- example, for farming practices (so-called calendar
compensated sun compass as a navigational tool birds). Observers also noted considerable variation
(Kramer, 1959). Although being mostly diurnal between species in punctuality. So-called weather
as a taxon, many species show occasional or fre- birds differ from calendar birds by large variation
quent nocturnal and crepuscular activities (Daan in spring arrival dates, often in response to exter-
and Aschoff, 1975; Martin, 1990) or keep track nal conditions (Gwinner and Helm, 2003; Newton,
of lunar and tidal cycles (Daan and Koene, 2008; Ramenofsky et al., 2012). The spectrum of
1981; Leyrer et al., 2006; Pinet et al., 2011). patterns found in nature indicates that long-term
Migration adds a layer of complexity to timing timing in birds involves diverse internal and exter-
on these different scales. Movements often have nal factors (Visser et al., 2010). For the purpose
substantial effects on the Zeitgeber conditions a of studying endogenous rhythms, the primary
460

Red Knot
Great Knot
Bar-Tailed Godwit
Black-Tailed Godwit
Ruff

Fig. 1. Map showing major southward migration pathways around the world of Red Knot, Great Knot, Ruff, Bar-Tailed Godwit,
and Black-Tailed Godwit. Different lines for Red Knots and Bar-Tailed Godwits represent subspecies with differing migration
patterns.

interest of chronobiologists has been in calendar (Hau et al., 1998; Helm et al., 2009) or follow
birds. Nonetheless, weather birds have a lot to offer, accelerated photoperiodic cycles by undergoing
as chronobiologists become increasingly interested several annual cycles within a single year (Gwinner,
in plasticity of time keeping (Hut et al., 2012; 1986). However, the use of photoperiod as calendri-
Ramenofsky et al., 2012). cal information is greatly complicated by long-
The primary, external calendar that is available distance migration (Aschoff, 1955; Hamner and
to organisms is photoperiod, that is, the annual Stocking, 1970; Rowan, 1926). The very movements
cycle in length of the light fraction of the of migrants, carried out predominantly on a
day. In most global environments, daylength North–South axis, result in substantial modification
markedly changes over the year so that, for a of the daylength they experience. In extreme cases,
given locality, it provides precise predictive infor- transequatorial or interpolar migrations expose
mation (Bradshaw and Holzapfel, 2007; Foster birds to long days for most of the year (Egevang
and Kreitzman, 2009). Birds are highly photope- et al., 2010). This puts high demands on internal
riodic (Dawson et al., 2001; Sharp, 2005) and long-term timing mechanisms, including an ability
may adjust annual cycles in response to subtle to correctly interpret temporal cues at a given
changes in daylength of half an hour or less time and location. Photoperiodism must therefore
461

be complemented by refined mechanisms that regu- circannual rhythms and their phase responses
late its use in a phase-specific way (Bradshaw to the main Zeitgeber, photoperiod, are elaborated
and Holzapfel, 2007; Miyazaki et al., 2012). Such in greatest detail for invertebrates and are
mechanisms have been documented in experimen- summarized in this volume (Miyazaki et al., 2012).
tal and comparative studies. Birds usually become Circannual rhythms in migratory birds belong to
temporarily unresponsive to previously stimulating the earliest, definitive experimental demonstrations
daylengths (photorefractoriness; Dawson et al., of such long-term timing mechanisms (Gwinner,
2001; Hahn et al., 2009; Sharp, 2005). The conditions 1967) and are often particularly robust. For
for regaining responsiveness to stimulatory day- example, in Dark-eyed Juncos (Junco hyemalis),
lengths differ greatly between species and show circannual cycles in migratory restlessness and
clear links to migration (Gwinner, 1988, 1996a,b; related processes continued for several years even
Hahn et al., 2009; Helm et al., 2009). In general, under constant dim light (Holberton and Able,
long-distance migrants do not require short-day 1992). Palearctic long-distance migrants, such as
conditions to regain responsiveness to long days Sylvid warblers (Berthold et al., 1972), can express
(i.e., photosensitivity) but recover it spontaneously robust circannual rhythms under a wide range of
under a wide range of conditions. At the same constant daylengths from LD 10:14 h to 16:8 h.
time, delayed recovery of long-day responsiveness In contrast, in other, less migratory birds, the range
protects migrants from inappropriate seasonal of permissive conditions under which circannual
behavior like breeding on transequatorial winter rhythms are expressed can be very narrow
grounds (Gwinner, 1988, 1996a,b; Hamner and (Dawson et al., 2001). Detailed data for geographic
Stocking, 1970). Such phase-specific and species- differentiation of circannual rhythms are available
specific responses to daylength point to a high from studies of Stonechats (Saxicola torquata;
potential for evolutionary changes of biological time Gwinner, 1996b; Helm et al., 2009). In this songbird
keeping. Different populations within a species may taxon, local populations differ greatly in migratory
differ in the timing of seasonal activities under iden- behavior and correspondingly in seasonal sched-
tical field and captive conditions, and similar varia- uling. A comparison of Siberian long-distance
tion can be observed even within populations migrants and European short-distance migrants
(Battley, 2006; Curry-Lindahl, 1963; Rees, 1989). showed that these differences are partly hard-wired
Several recent studies report substantial heritability in distinct circannual rhythms. Consequently, the
for the timing of migration, which could allow taxa differed markedly in response to a given
rapid evolutionary adjustment to changing environ- daylength: for example, slightly longer simulated
mental conditions (Pulido and Berthold, 2003; van days of more southerly wintering grounds advanced
Noordwijk et al., 2006; Visser et al., 2010). Timing the annual cycle in European Stonechats but
differences between individuals can result from dif- delayed it in Siberian Stonechats (Helm et al.,
ferent responses to cues like daylength and ambient 2009). Such data indicate that the specific ways in
temperature (Bradshaw and Holzapfel, 2007). Such which circannual systems respond to Zeitgeber
differential responses, if innate, can be interpreted stimuli are presumably subject to strong selection
as “reaction norms” to environmental cues (i.e., (Gwinner, 1996a,b).
range of phenotypes expressed by a single genotype Environmental factors modify or fine-tune the
as a function of a given variable) and can be ana- timing and progress of migration (Newton, 2008;
lyzed by quantitative genetic tools (van Noordwijk Ramenofsky et al., 2012; Visser et al., 2010).
et al., 2006; Visser et al., 2010). Within the time window set for migration, its prog-
In various species of migratory birds, annual pro- ress may depend on conducive winds and weather
cesses oscillate as self-sustained circannual rhythms conditions, fuel stores, and feeding conditions
(Gwinner, 1986). The general characteristics of en route (Gwinner, 1999). In many species of
462

migrants, social factors also contribute to the fine- gland and hypothalamic structures and, in some
tuning of timing, so that aggregations may be seen species, also the retina. The pineal gland pro-
as individuals’ compromises between the relative duces and secretes the hormone melatonin into
benefits of joining a group and those of going by the bloodstream in a 24-h rhythm, with generally
own schedule (Helm et al., 2006). In contrast, a role high levels at night and low levels during day-
as Zeitgebers for circannual rhythms of migrants time. Melatonin plays a major signaling role in
appears unlikely for nonphotic factors like tropical the synchronization of avian daily rhythmicity.
seasonality or fluctuations in food availability Its rhythmic secretion persists with circadian
(Gwinner and Helm, 2003). For example, while periodicity even in pineal organs cultured
tropical migrants show highly localized breeding in vitro (Gwinner and Brandstätter, 2001). The
cycles, migrants depart from wintering sites with dif- second major component of the pacemaking sys-
ferent rainfall patterns much more synchronously. tem is a structure in the hypothalamus likened to
Further, in palearctic warblers, departure from the mammalian suprachiasmatic nucleus (SCN)
equatorial winter grounds coincides closely with that shows periodic expression of clock genes
the onset of spring migratory restlessness in captive and generates an autonomous circadian rhythm
conspecifics. Both observations support the idea with neuronal output. The rhythms created by
that spring preparations of equatorial migrants these main pacemakers interact with each other,
are mainly timed by an endogenous circannual as detailed in models of “neuroendocrine feed-
mechanism (Gwinner and Helm, 2003). Con- back loop” and “internal resonance” (Cassone
tributions of programmed timing at one extreme and Menaker, 1984; Gwinner and Brandstätter,
and of environmental factors at the other differ not 2001). The internal-resonance model assumes
only between calendar birds and weather birds but that the pineal rhythm of melatonin secretion
sometimes also within species and over the course amplifies the rhythm in the SCN, and in turn,
of a migration period (Newton, 2011). Clearly, the rhythm of the SCN amplifies that in the
the balance between control of annual cycles by pineal via the hypothalamo-spinal cord-superior
endogenous clocks versus environmental responses cervical ganglion tract. Such mutual augmenta-
must be finely adjusted to match the ecology of a tion would result in a high-amplitude circadian
given population (Visser et al., 2010). rhythm that would be strongly self-sustained
(Abraham et al., 2010; Gwinner and Helm,
2003; Menaker et al., 1997). Contributions of the
Daily time keeping oscillatory components to the overall pacemaking
system differ between the taxa studied in detail
Like annual cycles, daily cycles of birds have long (mostly songbirds and galliforms; Cassone et al.,
been considered precise and reliable, as sym- 2009; Gwinner, 2001; Kumar et al., 2004; Menaker
bolized by the wake-up call of the rooster in et al., 1997). Likewise, the amplitude of the nightly
rural life. The circadian basis of these cycles peak in melatonin differs substantially between
and their precision and photosensitivity were taxa. Rhythmicity of the pineal may be particularly
established by chronobiologists early on (Daan important in songbirds, in which elimination of
and Aschoff, 1975; Foster and Kreitzman, the melatonin rhythm severely disrupts activity
2005), and their physiological basis is known in rhythms (Gwinner et al., 1997). Nocturnally active
some detail (Brandstätter and Abraham, 2003; owls, in contrast, display robust circadian rhythms
Cassone et al., 2009; Gwinner, 2001; Kumar despite a rudimentary pineal gland and no signifi-
et al., 2004). The circadian pacemaking system cant melatonin amplitude (Van’t Hof et al., 1998).
of birds is complex and consists of several major Further, some but not all Arctic bird species for
components that generally include the pineal which melatonin profiles were examined during
463

polar midsummer had attenuated melatonin peaks and the two oscillators can stabilize in antiphase
while still retaining rhythmicity (e.g., Miché et al., (Bartell and Gwinner, 2005; Mukhin et al., 2009).
1991; Reierth et al., 1999; Silverin et al., 2009). Thus, In addition, there is evidence for masking by light
circadian organization is multifaceted and thereby that differs for the day and night components
offers many levels of adjustment for meeting the of locomotor activity (Coverdill et al., 2008).
particular challenges posed by migration. Switches to nocturnal restlessness can result from
In migratory birds, the circadian system masters internal and external factors. In many long-distance
at least two particular challenges. First, migrants migrants, periodic nocturnality is clearly driven by
often experience substantial shifts in the phase circannual rhythms (Gwinner, 1996a; Holberton
and length of the solar day, with implications for and Able, 1992), but coupling of the oscillators is
daily timing and for the correct interpretation of also sensitive to environmental factors, in particu-
time-compensated navigational cues. Surprisingly, lar, food. This plasticity is not restricted to migra-
even when moving quickly across time zones dur- tion and can be triggered by nutritional and
ing East–West flights, migrants can apparently social cues during nonmigratory periods and in
correctly use the sun compass (Kishkinev et al., resident birds (Helm and Gwinner, 2006; Mukhin
2010; Kramer, 1959). A second major challenge et al., 2009; Terrill, 1987). The mechanisms
for the circadian system are regular switches of and sites of the putative oscillators governing
daily activity patterns related to migration in diur- day–night-activity remain obscure. Some experi-
nal species that migrate at night. Considering the ments suggest that the pineal gland is involved
physiological challenges of migratory flight, the in the circadian regulation of migration-related
avian capacity to carry out such additional “shift behavior, as evidenced, for example, by disrupted
work” seems quite remarkable (Rattenborg et al., rhythms of nocturnal restlessness after pinealec-
2004). Research on migratory birds has been tomy (McMillan, 1972). The melatonin amplitude,
motivated by a wish to understand how these in turn, is modulated by conditions affecting
challenges are met. migratory activity. For example, refeeding of
Several lines of evidence show that the daily birds after fasting temporarily increased peak
recurrence of migratory behavior is under direct melatonin levels and interrupted nocturnal activ-
control of a circadian timer. With the exception of ity (Gwinner, 1996a). Recent experimental
one study (Coverdill et al., 2008), nocturnal rest- manipulation of melatonin levels as well as food
lessness of migrants remained clearly rhythmic availability indicated a primary effect of nutrition,
after transferring from synchronizing light–dark rather than of melatonin, on nocturnality (Fusani
cycles to constant dim light (e.g., Coppack et al., et al., 2011).
2008; Gwinner, 1996a; Kumar et al., 2006; Pohl, A role for melatonin in avian migration is
2000). When free-running period lengths were also suggested by changes in its amplitude during
estimated separately, they were shorter for “day migration periods. For example, in one study,
activity” than for migration-specific “night activ- the nocturnal melatonin peak was periodically
ity,” indicating possible involvement of at least damped during migratory season even under con-
two oscillators (Bartell and Gwinner, 2005). A cur- stant, circannual light conditions (Gwinner et al.,
rent conceptual idea therefore suggests that the 1993). Such fluctuations have been interpreted
temporal flexibility of the avian circadian system in the context of different pacemaker components
is partly achieved by two oscillators with variable in birds (Gwinner, 2001; Gwinner et al., 1997).
coupling. During diurnality, the night-activity oscil- As explained above for the internal-resonance
lator is suppressed by or phase-locked to the day- model, mutual augmentation of pacemakers
activity oscillator; if coupling strength of the would result in a highly self-sustained oscillation,
oscillators is modified, night activity appears which in turn makes the system relatively inert
464

in its ability to respond to phase shifts (Abraham lunar cycles of activity and additional coordination
et al., 2010). Conversely, a reduction of amplitude of seasonal migration with lunar and solar cycles
of the output of the pacemaker as a whole, as (Pinet et al., 2011). A role of lunar cycles for sea-
indicated by a reduced peak in melatonin, would sonal behavior has also been suggested for another
make the system more responsive to Zeitgebers sea-dweller, the Sooty Tern (Sterna fuscata; Order
(Gwinner et al., 1997). This has indeed been shown Charadriiformes) (Chapin and Wing, 1959). On a
by experiments that eliminated the pineal melatonin tropical island, this species breeds approximately
rhythm by implants or pinealectomy and resulted every 9 months, displaying what resembles a free-
in greater sensitivity of the circadian system to running circannual cycle in the wild. Closer analysis
weak Zeitgebers such as T-cycles and periodic food of the data suggests that the precise intervals are
availability (Gwinner and Brandstätter, 2001; determined by lunar cycles, so that the birds initiate
Gwinner et al., 1997). Accordingly, temporary reproduction during every 10th lunar cycle, but
reduction of the melatonin amplitude could enable long-term data from individual birds are still lacking
birds to modify their schedules rapidly during for consolidation of this idea (Chapin and Wing,
migration and thereby help them to cope with aris- 1959). Finally, tidal cycles are generally not well
ing phase shifts. This idea has been supported by understood but are clearly relevant for birds that
preliminary evidence for shorter resynchronization specialize on foraging along shorelines, that is,
times during than outside the migratory seasons waders, which are considered below in detail.
(Pohl, 2000). Adjustment of the melatonin ampli-
tude and its interactivity with light is therefore a can-
Temporal multitasking illustrated: Waders
didate mechanism for adaptations of birds for
meeting migration-related challenges (Gwinner
Waders master outstanding challenges of biological
and Helm, 2003).
time keeping. Specialized on coastal and terrestrial
wetlands, they cope with daily as well as tidal cycles,
carry out extreme long-distance migrations, and
Lunar and tidal cycles
breed at high latitudes (van de Kam et al., 2004;
Fig. 1). On their polar breeding grounds, they not
In contrast to solar cycles, lunar and tidal rhythms
only face an extremely short breeding season
of birds are poorly known. Most terrestrial species
but also encounter light conditions in which Zeitge-
are diurnal for most of the year and are believed
ber information is greatly reduced in amplitude.
to be relatively unaffected by moon-related cycles,
Waders are a diverse group that displays stark con-
while in nocturnal species, effects of moonlight on
trasts in lifestyles so that relationships between time
activity are evident (Martin, 1990). For example,
keeping, behavioral ecology, and physiology can be
Nightjars (Order Caprimulgiformes) reduce activ-
readily explored.
ity and drop body temperature when moonlight
is unavailable, resulting in cyclical heterothermy
related to the lunar cycle (Smit et al., 2011). Com- Annual punctuality, daily flexibility
pared to terrestrial species, seabirds have generally
more flexible diel activity patterns, which may be The group of waders contains many primary
linked to moonlight (Martin, 1990). In Boobies examples of highly punctual calendar birds
(Order Pelecaniformes) on the Galapagos Islands, (Piersma et al., 1990) as well as some weather birds
diel profiles of plasma melatonin and corticosterone (e.g., Lapwing, Vanellus vanellus; Schüz, 1971).
were indeed damped during moonlit nights (Tarlow Recent advances in animal tracking technology
et al., 2003). A recent study of tropical Petrels and observation efforts involving citizen scientists
(Order Procellariiformes) confirmed substantial have boosted information on migratory schedules.
465

Remarkable insights into the rigidity of scheduling a crucial aspect of breeding habitat is thaw date,
of annual cycle events come from Bar-Tailed which varies with breeding latitude. A study using
Godwits (Limosa lapponica baueri). These birds geolocators, which calculate position based on
carry out the longest documented nonstop migra- recordings of local daylight (Conklin et al., 2010),
tory flights for any landbird—10,000 km from New confirmed that the latitude the birds attempted to
Zealand to Asia, where they refuel before migrating breed at was closely associated with their schedul-
to their Alaskan breeding grounds, and 11,700 km ing throughout migration (Fig. 1). Even upon first
from Alaska to New Zealand after breeding (Fig. 1; departure from New Zealand, 45% of the variance
Battley et al., 2012; Gill et al., 2009). Preparation in the timing of migration was explained simply
for migration from New Zealand begins as early as by what latitude the bird was heading to. By
late December, when birds may start increasing individuals spending varying amounts of time
mass (P.F. Battley and J.R. Conklin, Massey Univ., staging in Asia, the relationship between departure
unpublished data) and soon afterward start to moult date and latitude became even stronger upon leav-
into breeding plumage. Birds depart on migration ing the Yellow Sea (r2 ¼ 0.80), and by the time
mostly from mid-March until early April (Battley, birds arrived on the breeding grounds, there was
1997; Conklin and Battley, 2011). The first data on almost no variation left to explain (r2 ¼ 0.91).
their repeatable schedules came from a color- Hence, the timing of an individual’s migration
banding study conducted in the Firth of Thames toward the arctic and subarctic breeding grounds,
(Battley, 2006). Individuals were monitored across even from deep into the southern hemisphere,
2 or 3 years, and despite limitations in detecting reflected a repeatable schedule relating to eventual
marked birds, there was a highly significant positive arrival time on the breeding grounds. Surprisingly,
relationship between last sighting in 1 year and the differences in arrival time tended to carry over to
equivalent date in a subsequent year. Most birds the end of the breeding season so that birds
departed within the same week each year, even from the more clement, lower-latitude locations
though the population as a whole departed across departed on their trans-Pacific migration earlier
a month, so that repeatability (Lessells and Boag, than those from the colder high-latitude locations
1987) of apparent departure date was 0.83. This (r2 ¼ 0.32). Taken together, the observed patterns
work was followed up at an isolated estuary where strongly suggest that these Godwits largely rely
a smaller population could be monitored in great on internal timing programs for their migratory
detail (Conklin and Battley, 2011). There, individ- schedules and that these timing programs may dif-
ual departures were generally confirmed to the fer between breeding populations, as reported for
day and mostly directly observed. Quite remark- songbirds (Curry-Lindahl, 1963; Helm et al., 2009).
ably, the calculated repeatability was virtually iden- A second, striking example of punctuality in
tical (0.84) to that from the previous study. Across a waders may give leads to how such local patterns
span of roughly a month, most birds left within the might evolve. In the closely related Icelandic
same week across 3 years; most of those that did Black-Tailed Godwit (L. limosa islandica), the
not were delayed due to avoidance of winds that mates of breeding pairs winter in separate locations
were unsuitable for migration. Clearly, New (Gunnarsson et al., 2004; Fig. 1). Surprisingly, previ-
Zealand Godwits have strongly characterized indi- ously mated birds nonetheless returned to the
vidual migration schedules. breeding sites at very similar times as their mates
One possible interpretation for individual varia- (r ¼ 0.97). Synchrony of arrival times was also
tion was that birds of different schedules bred demonstrated in the lower-latitude sister subspecies
in places with different breeding seasons, as L. limosa limosa (Lourenço et al., 2011). This may
suspected for sympatrically wintering songbirds be highly consequential for mate choice and breed-
(Curry-Lindahl, 1963). For waders like Godwits, ing success because pair formation in migratory
466

birds can be strongly time-dependent. In the other species indicate that light conditions might
Iceland-breeding subspecies, two out of the three affect activity patterns via predation risk
Black-Tailed Godwit pairs whose mates returned (Beauchamp and Ruxton, 2008). In Oystercatchers
8 days or more apart divorced, suggesting that syn- (Haematopus ostralegus), behavioral observations
chronous arrivals are an important factor in strongly suggested periodic movements in anticipa-
mate retention. Similarly, time-dependent pair for- tion of tidal food availability. An experimental
mation is known from other migrants like penguins, approach yielded findings that, although not con-
swans, and songbirds (summarized in Helm et al., clusive, were suggestive of possible (circa-)tidal
2006). Based on such observations and on an clocks in these birds (Daan and Koene, 1981).
assumed partly genetic regulation, the evolution of Temporal flexibility is also reported from
timing could be accelerated by assortative mating Arctic-breeding waders that frequently experi-
(Bearhop et al., 2005). ence continuous polar light from arrival until
In at least two species of wader, the precise migratory departure. The annual time period dur-
annual timing is based on robust circannual rhythms ing which the birds can breed is extremely short so
(Piersma, 2002; Piersma et al., 2008). Birds that that they should reproduce as early as possible after
were kept first under simulated natural daylength arrival to maximize reproduction (Yohannes et al.,
and then under constant light–dark cycles displayed 2010). While seasonal time pressure is high, light
clear circannual cycles of periodic fattening and availability could be permissive for efficient timing
moult of body plumage and flight feathers. Strik- of activities around the clock. Preliminary data
ingly, however, these cycles differed from those in from a single Alaskan study site support this idea
songbirds by an endogenous period length that (Silke Steiger et al., unpublished data). The study
was clearly longer, rather than shorter, than the examined several sympatrically breeding species
solar year (Gwinner, 1986; Piersma et al., 2008). that differed in mating system and in the ways the
Daily activity patterns of waders are generally sexes shared parental care (Székely et al., 2006).
reported to be relatively flexible and are thought Based on this diversity, species-specific activity
to depend only weakly on internal clocks (Krüll patterns were predicted and indeed discovered.
et al., 1985; Martin, 1990). This is based on several Under continuous light, the birds were either briefly
lines of evidence. Behavioral observations indicate active around the clock or showed rhythmic activity,
that waders can forage at night, aided by sophisti- most commonly while incubating (Cresswell et al.,
cated tactile sensing of prey in soft ground. When 2003). However, only some species synchronized
light intensity is increased, for example, by artifi- their rhythmic activities to the 24 h cycle of subtle,
cial illumination along shorelines, the birds possible Zeitgebers such as changes in light intensity,
take advantage of visual opportunities and extend polarization pattern, or sun azimuth (Gwinner,
foraging into the night (Santos et al., 2010). Flexi- 2001; Krüll et al., 1985; Pohl, 1999). Activity rhythms
ble timing is also strongly suggested by activity of other species drifted across the 24-h day, and
patterns of waders in tidal habitats, for which rich in some of these cases, breeding partners showed
descriptive data and, to our knowledge, the only remarkably well-coordinated incubation shifts
experiment on avian tidal cycles are available. which are suggestive of social synchronization (Silke
For example, constant recording of Red Knots Steiger et al., unpublished data).
(Calidris canutus) at a roost in the winter quarters
yielded clear bitidal patterns of attendance. Rest
times of the birds drifted straight across daytime Melatonin cycles of captive waders
and nighttime without any evidence of modulation
or masking by daylight (Leyrer et al., 2006; see van In view of their extensive temporal multitasking,
Gils et al., 2006). However, observations from we used waders to explore mechanisms that may
467

enhance such flexibility. Based on concepts of Zeitgebers. We also asked (3) whether melatonin
internal resonance, Daily time keeping and neuro- profiles changed across the year. We predicted
endocrine feedback loops (see above), we specula- that, in both species, melatonin should be reduced
ted that melatonin cycles of waders should have during life-cycle stages when birds entrain to rap-
low amplitudes, which could be further reduced idly shifting or subtle Zeitgebers, that is, during
during seasons when the birds would entrain to migration and breeding. Alternatively, melatonin
subtle Zeitgebers. To test predictions of circadian profiles could vary seasonally, simply as a reflec-
modulation via adjustment of the melatonin ampli- tion of light conditions. (4) Lastly, we tested for
tude, we compared two wader species that share suppressive effects of moonlight.
many features of migration and the annual cycle Both species were studied in small flocks under
(Fig. 1) (Mendes et al., 2006; Piersma et al., identical conditions at the Royal Netherlands
1996). The Ruff (Philomachus pugnax) and Red Institute for Sea Research (NIOZ) on the island of
Knot are similarly sized basal waders (sandpipers, Texel (53 050 N 4 500 E) in the Dutch Wadden Sea.
family Scolopacidae). Both breed at mid to high Data on annual cycles in various body traits, includ-
latitudes, winter in temperate to tropical zones, ing seasonally changing immune parameters, have
have definitive periods of migration, and like to been published elsewhere (Piersma et al., 2000).
flock. They show robust annual cycles in the wild For some Red Knots, overall seasonal patterns of
(Buehler and Piersma, 2008) and in captivity nightly amplitude and duration of elevated plasma
(Piersma et al., 2008). However, the species also melatonin were compared to immune parameters
represent prominent contrasts in ecology to test for immuno-enhancement (Buehler et al.,
(summarized in Piersma, 2003; Mendes et al., 2009). Here, we newly report melatonin profiles
2006; Piersma et al., 2000). Whereas Ruffs stay at around the year for Ruffs and detailed diel profiles
inland freshwater wetlands and breed at mid for Knots. We relate these patterns to seasonal
latitudes, Knots never leave the seashore except activities and to tentative data on activity collected
to breed on the northernmost tundra (Fig. 1). from small groups because these gregarious birds
Ruffs are lek-breeding, highly sexual dimorphic cannot be well studied individually.
birds, while Knots are socially monogamous with
biparental care. Ruffs typically have strongly pat- Methods
terned diel rhythms of diurnal feeding and noctur-
nal use of communal roosts, whereas Knots lead a Experiments involved eight Red Knots of the sub-
fully tidal life, foraging during day and night, and species islandica (four males, three females, one
gathering at roosts at high tides (Leyrer et al., sex unknown), which winters in Europe, and seven
2006), except for the brief time when they breed Ruffs (all males), which use the Wadden Sea for
in the High Arctic. Based on these differences, staging (Fig. 1; for details, see Buehler et al., 2009;
we expected to find corresponding differences in Piersma et al., 2000). In the Dutch Wadden Sea
the daily and annual profiles of melatonin. Ruffs were caught on March 15, 1997 using a
Specifically, we asked (1) whether melatonin “wilsternet,” while Knots were captured with
rhythms were low in these birds that perform on mistnets near a high-tide roost between 1994 and
different temporal scales and entrain to subtle 1997. At capture, birds were ringed, aged as older
Zeitgebers and (2) whether melatonin profiles dif- than 2 years, and transferred to outdoor aviaries
fered between the two species such that Ruffs (4.5 m  1.5 m  2.3 m high) at the NIOZ. Knots
would show a clear, diel pattern. Knots should were kept in a single aviary, while Ruffs, which
have a lower melatonin amplitude than Ruffs as are more territorial, were divided over two
an adaptation for, or a consequence of, tidal activ- aviaries. One-quarter of the surface contained a
ity cycles to enhance entrainment to nonphotic shallow basin with sand and running seawater
468

(mudflat). Each aviary contained a small freshwa- In Knots, feeding activity was also surveyed by
ter basin for drinking and bathing. The birds were an additional sensor next to the single source of
fed ad libitum portions of protein-rich trout-food food. Ruffs, in contrast, were equipped with sev-
pellets and once per week were screened for physi- eral feeders because of their higher aggression
cal condition, while the aviaries were disinfected so that foraging was not monitored. Locomotion
and the food replenished. Birds experienced local and feeding were surveyed by custom-made pho-
photoperiod and temperature. Compared to wild tocell systems. Beam breaks of an infrared beam
conspecifics, Ruffs were thus exposed to identical were recorded in 2-min bins on a computerized
conditions in spring and fall (Fig. 1). In summer, data-logging system. Activity data originated
they would have experienced slightly longer days from initially eight Knots (six birds from August
at more northerly latitudes and winter days would 1997) and from initially four Ruffs (three birds
also be longer on their African winter grounds. from August 1997). Sensors yielded data through-
In contrast, Knots were exposed to conditions out experimentation, except for brief periods of
that were identical to those experienced by wild technical problems. Because of behavioral
conspecifics that stay in the Netherlands from late differences between the species, we interpreted
summer until spring. During summer, daylength their activity records from the artificial mudflats
was much shorter for the captive Knots than for differently. Red Knots forage on mudflats, and
conspecifics on their Arctic-breeding grounds. even in our aviaries with permanent food supply
Motivation for migration should be similar in elsewhere, they spent 10–20% of each day actively
spring but differ between the species in autumn probing for food on the artificial mudflats. In con-
when Ruffs should be active migrants, whereas trast, they roost on dry land and not on mudflats.
Knots would be already at their wintering grounds. Therefore, recordings from the mudflat in Knots
Behavioral observations of Knots and other species reflect feeding and general activity, and the addi-
suggest that migration may indeed be suppressed tional feeding sensor narrowed down foraging
by site recognition (Ketterson and Nolan, 1988; motivation. Ruffs, in contrast, usually do not feed
Schwabl et al., 1991). To delineate life-cycle stages on mudflats, but they like to roost on wet ground
(Fig. 2), data on phenology were collected for all if available. Therefore, recordings from the mud-
birds once per week. Body mass was measured on flat combine general activity and resting behavior.
an electronic balance to the nearest gram. Moult However, the group recordings of both species
was scored on breast plumage for body moult yield only tentative information.
(0 ¼ no feathers growing, 1 ¼ feathers growing)
and on the primaries for flight feather moult
(scores from 1 ¼ feather dropped to 5 ¼ new Results and discussion
feather for all 10 primaries, thus reaching values
from 0 to 50; Ginn and Melville, 1983). Extent of Daily profiles of melatonin are shown for all sam-
breeding plumage was scored on a scale between pling periods and separately for life-cycle stages
1 ¼ basic, indicating full gray winter plumage, and (Figs. 3 and 4). We conducted two overview ana-
7 ¼ full alternate, indicating full rusty-red breeding lyses (Table 1) that converged in all major points:
plumage. Experiments started for Knots on Febru- melatonin levels strongly differed between day-
ary 20 and for Ruffs on March 24, 1997 and were time versus nighttime (LD) and depended greatly
completed on February 11, 1998. Melatonin was on life-cycle stage. We also found some evidence
sampled biweekly as explained in Fig. 3. that effects of light differed over the annual cycle.
As a rough overview of activity patterns, in one In general, melatonin peaks tended to be higher
aviary group of each species wading in the during late moult and wintering, but there was
simulated mud flats was continuously recorded. no direct relationship to daylength (Figs. 3 and
469

(a) (b)
Knot Winter Spring Breeding state Moult Winter Ruff Winter Spring Breeding state Moult
Fuel Fly Transitory Fuel & fly Fuel & fly Postmigratory

16 16
Daylength (h)

Daylength (h)
14 1 2 3A 3B 4 1 1 2 3 4A 4B 1 14

12 12

10 10
8 8

Mar 1 Apr 1 May 1 Jun 1 Jul 1 Aug 1 Sep 1 Oct 1 Nov 1 Dec 1 Jan 1 Feb 1 Mar 1 Apr 1 May 1 Jun 1 Jul 1 Aug 1 Sep 1 Oct 1 Nov 1 Dec 1 Jan 1 Feb 1

240
180

Body mass (g)


Body mass (g)

220
160
200
140
180
120
160
100
140
Mar 1 Apr 1 May 1 Jun 1 Jul 1 Aug 1 Sep 1 Oct 1 Nov 1 Dec 1 Jan 1 Feb 1 Mar 1 Apr 1 May 1 Jun 1 Jul 1 Aug 1 Sep 1 Oct 1 Nov 1 Dec 1 Jan 1 Feb 1

50 50
Primary moult

Primary moult
40 40
30 30
20 20
10 10
0 0
Mar 1 Apr 1 May 1 Jun 1 Jul 1 Aug 1 Sep 1 Oct 1 Nov 1 Dec 1 Jan 1 Feb 1 Mar 1 Apr 1 May 1 Jun 1 Jul 1 Aug 1 Sep 1 Oct 1 Nov 1 Dec 1 Jan 1 Feb 1
8 8
Plumage color

Plumage color
6 6

4 4

2 2

0 0
Mar 1 Apr 1 May 1 Jun 1 Jul 1 Aug 1 Sep 1 Oct 1 Nov 1 Dec 1 Jan 1 Feb 1 Mar 1 Apr 1 May 1 Jun 1 Jul 1 Aug 1 Sep 1 Oct 1 Nov 1 Dec 1 Jan 1 Feb 1
4 4

Breast moult
3 3
Breast moult

2 2

1 1

0 0

Mar 1 Apr 1 May 1 Jun 1 Jul 1 Aug 1 Sep 1 Oct 1 Nov 1 Dec 1 Jan 1 Feb 1 Mar 1 Apr 1 May 1 Jun 1 Jul 1 Aug 1 Sep 1 Oct 1 Nov 1 Dec 1 Jan 1 Feb 1

Fig. 2. Derivation of life-cycle stages based on phenology of (a) Knots and (b) Ruffs. Top: natural daylength, with stars marking
days of blood sampling and letters indicating life-cycle stage; bottom: mean  SD for body mass and scores of primary moult,
plumage color, and breast moult; vertical lines indicate demarcations of life-cycle stages. Stage definitions Knot: (1) “winter”:
quiescent time. From end of autumn moult until start of spring preparations, that is, fattening and moult into breeding plumage;
(2) “spring”: fueling and prenuptial moult. From start of fattening, moult, and recoloration until reaching full breeding coloration
and peak body mass; (3) “breeding state”: defined by full breeding plumage. From reaching peak body mass and breeding
coloration until the onset of moult; (3A) “flight phase”: ca. 4 weeks of using up fuel stores, from maximum to trough of body
mass. Prediction: melatonin low; (3B) “transitory phase for nonbreeders”: corresponding to breeding in the wild, from end of
body mass peak until start of moult. Prediction: melatonin low; (4) “moult”: from start to completion of primary moult,
corresponding to moult of wild conspecifics on the winter grounds slightly later in the year. Stage definitions Ruff: (1) “winter”:
quiescent time. From end of autumn moult until start of spring preparations, that is, moult into breeding plumage; (2) “spring”:
fueling and prenuptial moult. In contrast to Knots, Ruffs overlap these activities with migratory flight. From start of moult and
recoloration until reaching full breeding coloration and peak body mass. Prediction: melatonin low; (3) “breeding state”: defined
by full breeding plumage. From peak body mass and breeding coloration until the onset of moult; (4) “moult”: renewal of
primaries. From start to completion of primary moult, corresponding to moult of wild birds started during migration and
completed on the African winter grounds; (4A) “fuel, flight and moult phase,” starting with a transitory phase in nonbreeders
with slow and then progressively faster moult and fueling. Unlike in captive Knots, this corresponds to flight phases in the wild.
From starting of color loss and increase of body mass until birds reach peak body mass. Prediction: melatonin low; (4B)
“postmigratory completion of moult”: From end of body mass peak until completion of moult, corresponding to finalizing of
moult in Africa.
470

Knots Ruffs
60
30

60 Mar 1
30

60 60
30 30
60 60
30 Apr 1 30

60 60
30 30

60 60
30 May 1 30

60 60
30 30

60 60
30 Jun 1 30

60 60
30 30

60 60
30 30
Jul 1
60 60
30 30

Sep 1
60 60
30 30

60 60
30 30
Oct 1
60 60
30 30

60 60
30 30
Nov 1
60
30

60 60
30 30
Dec 1
60 60
30 30

60
30
Jan 1 60
60
30 30

60 60
30 30

60 Feb 1 60
30 30

12:00 16:00 20:00 24:00 04:00 08:00 12:00 12:00 16:00 20:00 24:00 04:00 08:00 12:00
471

4). Both models did not support species Actograms were also suggestive of additional,
differences. Results for damping effects of moon- longer periodicities in autumn, which, however,
light on melatonin were equivocal: comparison of could not be ascertained. Surprisingly, the main
elevated versus baseline melatonin by logistic feeding activity of Knots shifted into the night
regression was more supportive of moonlight after moult. In Ruffs, support for rhythmicity
effects than comparison of melatonin levels by lin- was generally much weaker and primary peri-
ear mixed effect models. Within the species, sub- odicities were more variable than in Knots. How-
stantial effects of light (LD) and life-cycle stage ever, during postmigratory moult, Ruffs showed
were confirmed, while there was little evidence pronounced activity cycles that recurred twice
for interactions between these factors. Effects of per day.
moon were clearly more pronounced in Ruffs com- Overall, our study showed rhythmic, but low
pared to Knots. Predictions for stage-dependent melatonin levels in the two wader species that
suppression of melatonin amplitudes were only were of similar magnitude as in other seabirds
partly supported. In Fig. 4, the life-cycle stages (Miché et al., 1991; Tarlow et al., 2003; Wikelski
for which low melatonin was predicted are plotted et al., 2006). Melatonin cycles changed over the
by solid lines. While melatonin was indeed low annual cycle, but the seasonal course did not fully
during these times of year, similarly low support damping during particular life-cycle
amplitudes also occurred at other times of year. stages as had been shown for migratory songbirds
Group activity patterns were overall rhythmic (Gwinner and Brandstätter, 2001; Gwinner et al.,
(Fig. 5; Table 2). In Knots, wading and feeding 1993), nor did it simply reflect the amount of
recurred twice a day, with dominant periodicities available daylight. Our findings add to heteroge-
around 12 h (during the breeding flight phase, a neous evidence for links between light availability
main peak around 22 h had only slightly higher and melatonin in birds (Buehler et al., 2009;
spectral power than one at 10.9 h). These peri- Gwinner and Brandstätter, 2001; Miché et al.,
odicities had relatively high spectral power but 1991; Reierth et al., 1999; Silverin et al., 2009).
were best supported during moult and least Moonlight apparently exerted a mild, damping
supported during the breeding flight phase. influence on melatonin (Tarlow et al., 2003),

Fig. 3. Diel melatonin profiles across the annual cycle for Knot (left) and Ruff (right) (mean  SE; [pg/ml]). Plots are arranged by
corresponding time of year. Vertical curves show time of sunset and sunrise, respectively. The birds were sampled biweekly to
examine diel profiles over an entire year, alternating between phases of full moon and new moon. Sampling took place over a 24-h
session, starting and ending at mid-day with seven bleedings per bird at 4-h intervals (12:00, 16:00, 20:00, 00:00, 04:00, 08:00;
12:00 h; Piersma et al., 2000). On each sampling date, blood was collected for at least six birds of both species. The protocol
followed local time, implying that samples were collected an hour earlier relative to the solar day during summer savings time.
Blood was collected within 15 min after entering the aviary by lightly puncturing the brachial wing vein. We collected 150–250 ml of
blood into heparinized capillary tubes that were stored cool. Tubes were centrifuged within 2 h of sampling at 6900  g for 15 min,
and plasma was subsequently stored at -80  C until transfer on dry ice to the Max-Planck Institute for Ornithology in Andechs,
Germany. Data collection was briefly interrupted in August 1997, and for Ruffs, two sampling periods were missing (12 November;
19 December). Individual birds were sampled up to 22 times during a total of 23 sampling periods, and overall we analyzed 1667
samples. Simultaneous monitoring of health parameters indicated no harmful effects (Piersma et al., 2000). We quantified
melatonin by the established procedure at our laboratory (Silverin et al., 2009; Van’t Hof and Gwinner, 1996) as described for the
Knots (Buehler et al., 2009). Due to the sudden loss of Ebo Gwinner and subsequent complications, we unfortunately have only
long-term information about the melatonin assay at our institute. Accordingly, mean recoveries are at 90%, and intraassay and
interassay variation were 4.5% and 14%, respectively. For further details on Knots, see Buehler et al. (2009).
472

70
Knots Winter
Spring
Breed A
60 Breed B
Moult
Melatonin (pg/ml)

50

40

30

20

12:00 16:00 20:00 24:00 04:00 08:00 12:00

70
Ruffs
Winter
Spring
Breed
60
Moult A
Moult B
Melatonin (pg/ml)

50

40

30

20

12:00 16:00 20:00 24:00 04:00 08:00 12:00


Time of day

Fig. 4. Daily profiles of plasma melatonin (mean  SD; [pg/ml]) by species-specific life-cycle stage (cf. Fig. 2) for Knots (top) and
Ruffs (bottom). Solid lines highlight stages for which low melatonin amplitudes had been predicted.
473

Table 1. Melatonin profiles in Knot and Ruff

Species comparison Knot Ruff

Logistic Linear mixed Logistic Linear mixed Logistic Linear mixed


Factors regression effects regression effects regression effects

Light/dark w42 ¼ 198.4; w42 ¼ 159.3; w52 ¼ 154.8; w52 ¼ 109.9; w52 ¼ 52.9; w52 ¼ 46.3;
(LD) p < 0.001 p < 0.001 p < 0.001 p < 0.001 p < 0.001 p < 0.001
Species w12 ¼ 0.007; w12 ¼ 2.32; – – – –
p ¼ 0.935 p ¼ 0.128
Life-cycle w62 ¼ 138.6; w62 ¼ 102.4; w82 ¼ 69.4; w82 ¼ 47.4; w82 ¼ 110.9; w82 ¼ 103.5;
stage p < 0.001 p < 0.001 p < 0.001 p < 0.001 p < 0.001 p < 0.001
Moon w12 ¼ 5.13; w12 ¼ 0.02; w12 ¼ 0.6; w12 ¼ 5.1; w12 ¼ 18.7; w12 ¼ 8.5;
p ¼ 0.02 p ¼ 0.88 p ¼ 0.449 p ¼ 0.024 p < 0.001 p ¼ 0.003
Light/dark* w32 ¼ 11.3; w32 ¼ 9.5; w42 ¼ 13.2; w42 ¼ 4.1; w42 ¼ 3.9; w42 ¼ 7.0;
Life-cycle p ¼ 0.010 p ¼ 0.023 p ¼ 0.010 p ¼ 0.392 p ¼ 0.426 p ¼ 0.134
stage

Results from analyses using logistic regression and linear mixed effect models; inference based on likelihood-ratio testing. Data on melatonin were not
normally distributed because about two-thirds of the values were at baseline concentrations (lower detection limit: 17 pg/ml). We analyzed these cen-
sored data in two ways: Distinguishing between baseline and elevated melatonin, we performed logistic regression with binominal error distribution,
using procedure GLMER in the software R (R Core Development Team, 2009). We analyzed melatonin grouped by individual bird by the factors
daytime versus nighttime (LD), “moon,” “species,” life-cycle stage (stage), and interaction of LD stage, (LD* stage, whereby asterisk indicate inter-
action). In this model, assumptions were met, but information content was reduced. We then analyzed original (log-converted) melatonin data using
linear mixed effect models (procedure LMER in R), using full information but deviating from distributional assumptions. Significance testing was per-
formed by likelihood-ratio testing (procedure ANOVA in R). In these overview analyses, life-cycle stages (see Fig. 2) were made comparable by
pooling data from stage “breed A” in Knots with those of “spring,” and by pooling data from stage “moult B” in Ruffs with those of winter. Within
species, we analyzed data over their defined, specific life-cycle stages.

particularly in Ruffs. Given the flexible schedules Concluding remarks


of waders and their entrainment to subtle
Zeitgebers, one may ask why waders retained diel We hope to have shown that migratory birds are an
melatonin rhythmicity at all, while some noctur- exciting study system because their wide range of
nal species have no significant melatonin ampli- temporal problem-solving offers opportunities for
tude and may nonetheless show robust circadian addressing important questions in chronobiology.
rhythms (Van’t Hof et al., 1998; Wikelski et al., As first suggested for birds, circannual systems
2006). Presumably, the rise in melatonin in can now progressively be viewed as “periodically
waders could serve physiological functions other changing dispositions to respond to environmental
than those involved in circadian rhythmicity, for cues” (Gwinner, 1999) that are fine-tuned by evo-
example, induction of sleep despite continuous lution. New studies of migrant circadian systems
light and reduction in body temperature (Silverin highlight remarkable plasticity and responsiveness
et al., 2009; Tarlow et al., 2003). An additional to external conditions, but mechanisms behind the
possibility, immuno-enhancement, was found to observed diversity of avian time keeping still
be unlikely for Red Knots (Buehler et al., 2009). remain to be understood. The balance between
Experimental studies involving melatonin treat- rigidity and flexibility of timing, as evident in
ment of different avian species did indeed indi- migratory birds, is of growing interest to chronobi-
cate separate adjustment of several regulatory ology but is also relevant for the general public.
functions of melatonin (Murakami et al., 2001), Global change progresses at a disconcerting rate
but such presumable, diverse effects of melatonin and its implications for seasonal and daily timing
cycles in birds still require further exploration. become increasingly evident. Organisms have
474

May 1

May 1
May 1

Jun 1

Jun 1
Jun 1

Jul 1

Jul 1
Jul 1

Aug 1

Aug 1
Aug 1

Sep 1

Sep 1
Sep 1

Oct 1

Oct 1
Oct 1

Nov 1

Nov 1
Nov 1

Dec 1

Dec 1
Dec 1

Jan 1

Jan 1
Jan 1

Feb 1

Feb 1
Feb 1

00:00 12:00 24:00 12:00 24:00 00:00 12:00 24:00 12:00 24:00 00:00 12:00 24:00 12:00 24:00
Time of day

Fig. 5. Actograms showing activities of groups of waders around the annual cycle; data are double-plotted across the day. Left:
wading of Knots; middle: feeding of Knots; and right: wading of Ruffs on a simulated mudflat.
475

Table 2. Rhythmicity of wading activity of groups of Knots and Ruffs; tabulated data are time intervals of analysis, primary
periodicity, and corresponding maximum of spectral power separated by life-cycle stage

Life-cycle stage Start End Period Spectral power

Knot
1 Winter Nov 18, 1997 Feb 3, 1998 12.1 354.5
2 Spring Apr 5, 1997 May 19, 1997 13.1 953.5
3 Breed A “flight phase” May 20, 1997 Jun 23, 1997 21.8 148.2
3 Breed B “transitory” Jun 24, 1997 Jul 16, 1997 11.8 461.0
4 Moult (1) Aug 7, 1997 Aug 22, 1997 12.0 1851.2
4 Moult (2) Aug 29, 1997 Nov 17, 1997 11.9 1493.7
Ruff
1 Winter Apr 5, 1997 Apr 14, 1997 15.9 52.7
2 Spring Apr 15, 1997 May 26, 1997 46.4 46.3
3 Breed (1) May 27, 1997 Jul 16, 1997 60.5 78.4
3 Breed (2) Jul 25, 1997 Aug 5, 1997 12.2 87.7
4A Moult “fuel and flight” Aug 29, 1997 Nov 23, 1997 11.9 153.9
4B Moult “postmigratory” Nov 24, 1997 Feb 3, 1998 12.0 10082.4

Due to equipment failure, registration from the moult stage in Knots and the breeding stage in Ruffs were split into two parts. We used maximum
entropy spectral analysis (MESA) (Dowse, 2007) to identify the primary periods during which spectral power peaked.

altered their temporal behavior and natural sys- the DEC (Animal Experiment Committee) of the
tems begin to come out of synchrony in what Dutch Royal Academy of Sciences (KNAW).
appears to be a puzzling, largely unexplained mix- Tom Van’t Hof provided valuable endocrine
ture of organismic responses. While many species expertise and ideas for the melatonin experiments.
cannot keep up with changes, others are surpris- Wolfgang Goymann and Josef Habersetzer helped
ingly apt to take advantage of new conditions. to complete analysis of melatonin and activity data.
Changes in temporal patterns, and as a conse- Marina Lehmann, Fränzi Körner, and Dusty
quence, in behavior and species distribution, likely Dowse gave friendly assistance on data handling
affect not only biodiversity per se but also human and analysis, and Edda Starck and Dick Visser cre-
health and agricultural productivity. Understand- ated graphics. We are most grateful to Joop
ing changes in time keeping has therefore become Jukema for catching the Ruffs, Piet Duiven for
of pressing interest. Chronobiologists are in a catching the Knots, and Ciska Raaymakers, Timo
privileged position to address such predicted Verbeek, and Bernard Spaans for all sorts of prac-
changes and the mechanisms that might determine tical help. The shorebird work on Texel was
organismic responses. supported by a PIONIER-grant to T. P. from the
Netherlands Organization for Scientific Research
(NWO). B. H. is supported by the European Social
Acknowledgments Fund in Baden-Württemberg.

We greatly miss our coauthor Ebo Gwinner, who


has been the driving intellectual force behind the References
empirical study but did not live to see its publica-
Abraham, U., Granada, A. E., Westermark, P. O., Heine, M.,
tion. We thank Helga Gwinner for kindly Kramer, A., & Herzel, H. (2010). Coupling governs entrain-
supporting this late publication of his work. ment range of circadian clocks. Molecular Systems Biology,
Research was carried out under the auspices of 6, 438.
476

Aschoff, J. (1955). Jahresperiodik der Fortpflanzung bei paradoxes for avian circadian biology. General and Compar-
Warmblütern. Studium Generale, 8, 742–776. ative Endocrinology, 163, 109–116.
Bartell, P. A., & Gwinner, E. (2005). A separate circadian oscil- Chapin, J. P., & Wing, L. W. (1959). The Wideawake Calen-
lator controls nocturnal migratory restlessness in the songbird dar, 1953 to 1958. The Auk, 76, 153–158.
Sylvia borin. Journal of Biological Rhythms, 20, 538–549. Chovnick, A. (Ed.), (1960). 1960 Biological clocks. Cold
Battley, P. F. (1997). The northward migration of Arctic waders Spring Harbor symposia on quantitative biology, Cold
in New Zealand: Departure behaviour, timing and possible Spring Harbor, L.I., NY: The Biological Laboratory.
migration routes of Red Knots and Bar-tailed Godwits from Conklin, J. R., & Battley, P. F. (2011). Impacts of wind upon
Farewell Spit, North-West Nelson. Emu, 97, 108–120. repeatable individual migration schedules of New Zealand
Battley, P. F. (2006). Consistent annual schedules in a migra- bar-tailed godwits. Behavioral Ecology, 22, 854–861.
tory shorebird. Biology Letters, 2, 517–520. Conklin, J. R., Battley, P. F., Potter, M. A., & Fox, J. W. (2010).
Battley, P. F., Warnock, N., Tibbitts, T. L., Gill, R. E., Jr., Breeding latitude drives individual schedules in a trans-
Piersma, T., Hassell, C. J., et al. (2012). Contrasting extreme hemispheric migrant bird. Nature Communications, 1, 67.
long-distance migration patterns in bar-tailed godwits Coppack, T., Becker, S. F., & Becker, P. J. J. (2008). Circadian
Limosa lapponica. Journal of Avian Biology, 43, 21–32. flight schedules in night-migrating birds caught on migra-
Bearhop, S., Fiedler, W., Furness, R. W., Votier, S. C., tion. Biology Letters, 4, 619–622.
Waldron, S., Newton, J., et al. (2005). Assortative mating Coverdill, A. J., Bentley, G. E., & Ramenofsky, M. (2008).
as a mechanism for rapid evolution of a migratory divide. Circadian and masking control of migratory restlessness in
Science, 310, 502–504. Gambel’s white-crowned sparrow (Zonotrichia leucophrys
Beauchamp, G., & Ruxton, G. D. (2008). Time of day and gambelii). Journal of Biological Rhythms, 23, 59–68.
flightiness in flocks of semipalmated sandpipers. The Cresswell, W., Holt, S., Reid, J. M., Whitfield, D. P., &
Condor, 110, 269–275. Mellanby, R. J. (2003). Do energetic demands constrain
Berthold, P., Gwinner, E., & Klein, H. (1972). Circannuale incubation scheduling in a biparental species? Behavioral
Periodik bei Grasmücken. I. Periodik des Körpergewichtes, Ecology, 14, 97–102.
der Mauser und der Nachtunruhe bei Sylvia atricapilla und Curry-Lindahl, K. (1963). Molt, body weights, gonadal develop-
S. borin unter verschiedenen konstanten Bedingungen. ment, and migration in Motacilla flava. In Proceedings of the
Journal of Ornithology, 113, 170–190. XIIIth International Ornithological Congress, Ithaca, 960–973.
Birkhead, T. R. (2008). The wisdom of birds. An illustrated Daan, S., & Aschoff, J. (1975). Circadian rhythms of locomo-
history of ornithology. London: Bloomsbury. tor activity in captive birds and mammals: Their variations
Bradshaw, W. E., & Holzapfel, C. M. (2007). Evolution of ani- with season and latitude. Oecologia, 18, 269–316.
mal photoperiodism. Annual Review of Ecology, Evolution, Daan, S., & Koene, P. (1981). On the timing of foraging flights
and Systematics, 38, 1–25. by oystercatchers, Haemotopus ostralegus, on tidal mudflats.
Brandstätter, R., & Abraham, U. (2003). Hypothalamic circa- Netherlands Journal of Sea Research, 15, 1–22.
dian organization in birds. I. Anatomy, functional morphol- Dawson, A., King, V. M., Bentley, G. E., & Ball, G. F. (2001).
ogy, and terminology of the suprachiasmatic region. Photoperiodic control of seasonality in birds. Journal of
Chronobiology International: The Journal of Biological & Biological Rhythms, 16, 365–380.
Medical Rhythm Research, 20, 637. Dowse, H. (2007). Statistical analysis of biological rhythm
Buehler, D. M., Koolhaas, A., Van’t Hof, T., Schwabl, I., data. In E. Rosato (Ed.), Circadian rhythms: Methods and
Dekinga, A., Piersma, T., et al. (2009). No evidence for mel- protocols (Vol. 362), (pp. 29–45). Totowa, NJ: Humana
atonin-linked immunoenhancement over the annual cycle Press.
of an avian species. Journal of Comparative Physiology A: Dunlap, J. C., Loros, J. J., & DeCoursey, P. (Eds.), (2004). Chro-
Neuroethology, Sensory, Neural, and Behavioral Physiology, nobiology: Biological timekeeping. Sunderland, MA: Sinauer.
195, 445–451. Egevang, C., Stenhouse, I. J., Phillips, R. A., Petersen, A.,
Buehler, D. M., & Piersma, T. (2008). Travelling on a budget: Fox, J. W., & Silk, J. R. D. (2010). Tracking of Arctic terns
Predictions and ecological evidence for bottlenecks in the Sterna paradisaea reveals longest animal migration. Pro-
annual cycle of long-distance migrants. Philosophical Trans- ceedings of the National Academy of Sciences, 107, 2078–2081.
actions of the Royal Society of London, Series B:, Biological Fernández-Duque, E., de la Iglesia, H., & Erkert, H. G.
Sciences, 363, 247–266. (2010). Moonstruck primates: Owl monkeys (Aotus) need
Cassone, V. M., & Menaker, M. (1984). Is the avian circadian moonlight for nocturnal activity in their natural environ-
system a neuroendocrine loop? The Journal of Experimental ment. PLoS One, 5, e12572.
Zoology, 232, 539–549. Foster, R. G., & Kreitzman, L. (2005). Rhythms of life: The
Cassone, V. M., Paulose, J. K., Whitfield-Rucker, M. G., & biological clocks that control the daily lives of every living
Peters, J. L. (2009). Time’s arrow flies like a bird: Two thing. New Haven, CT: Yale University Press.
477

Foster, R. G., & Kreitzman, L. (2009). Seasons of life: The Gwinner, E., Schwabl-Benzinger, I., Schwabl, H., &
biological rhythms that enable living things to thrive and Dittami, J. (1993). Twenty-four hour melatonin profiles in
survive. New Haven, CT: Yale University Press. a nocturnally migrating bird during and between migratory
Foster, R. G., & Roenneberg, T. (2008). Human responses seasons. General and Comparative Endocrinology, 90,
to the geophysical daily, annual and lunar cycles. Current 119–124.
Biology, 18, R784–R794. Hahn, T. P., Watts, H. E., Cornelius, J. M., Brazeal, K. R., &
Fusani, L., Cardinale, M., Schwabl, I., & Goymann, W. (2011). MacDougall-Shackleton, S. A. (2009). Evolution of environ-
Food availability but not melatonin affects nocturnal rest- mental cue response mechanisms: Adaptive variation in
lessness in a wild migrating passerine. Hormones and Behav- photorefractoriness. General and Comparative Endocrinol-
ior, 59, 187–192. ogy, 163, 193–200.
Gill, R. E., Jr., Tibbitts, T. L., Douglas, D. C., Handel, C. M., Hamner, W. M., & Stocking, J. (1970). Why don’t Bobolinks
Mulcahy, D. M., Gottschalck, J. C., et al. (2009). Extreme breed in Brazil? Ecology, 51, 743–751.
endurance flights by landbirds crossing the Pacific Ocean: Hau, M., Wikelski, M., & Wingfield, J. C. (1998). A neotropi-
Ecological corridor rather than barrier? Proceedings of the cal forest bird can measure the slight changes in tropical
Royal Society B, 276, 447–457. photoperiod. Proceedings of the Royal Society of London,
Ginn, H. B., & Melville, D. S. (1983). Moult in birds. British Series B: Biological Sciences, 265, 89–95.
Trust for Ornithology Guide. Hertfordshire, England: Helm, B., & Gwinner, E. (2006). Migratory restlessness in an
Maunt and Iroine Ltd. equatorial nonmigratory bird. PLoS Biology, 4, 611–614.
Gunnarsson, T. G., Gill, J. A., Sigurbjörnsson, T., & Helm, B., Piersma, T., & Van der Jeugd, H. (2006). Sociable
Sutherland, W. J. (2004). Arrival synchrony in migratory schedules: Interplay between avian seasonal and social
birds. Nature, 431, 646. behaviour. Animal Behaviour, 72, 245–262 1215.
Gwinner, E. (1967). Circannuale Periodik der Mauser und der Helm, B., Schwabl, I., & Gwinner, E. (2009). Circannual basis
Zugunruhe bei einem Vogel. Naturwissenschaften, 54, 447. of geographically distinct bird schedules. Journal of Experi-
Gwinner, E. (1986). Circannual rhythms. Heidelberg, Berlin: mental Biology, 212, 1259–1269.
Springer. Holberton, R. L., & Able, K. P. (1992). Persistence of
Gwinner, E. (1988). Photorefractoriness in equatorial circannual cycles in a migratory bird held in constant dim
migrants. In Proceedings of the 19th International Ornitho- light. Journal of Comparative Physiology, 171, 477–481.
logical Congress, 626–633. Ottawa. Hut, R. A., Kronfeld-Schor, N., Van der Vinne, V., & De la
Gwinner, E. (1996a). Circadian and circannual programmes Iglesia, H. (2012). In search of a temporal niche: Environ-
in avian migration. Journal of Experimental Biology, 199, mental factors. In R. G. Foster, A. Kalsbeek, M. Merrow
39–48. & T. Roenneberg (Eds.), Neurobiology of circadian timing.
Gwinner, E. (1996b). Circannual clocks in avian reproduction Amsterdam: Elsevier.
and migration. Ibis, 138, 47–63. Ketterson, E. D., & Nolan, V. (1988). A possible role for expe-
Gwinner, E. (1999). Rigid and flexible adjustments to a rience in the regulation of the timing of bird migration. In
periodic environment: Role of circadian and circannual pro- Proceedings of the 19th International Ornithological Con-
grams. In N. J. Adams & R. H. Slotow (Eds.), Proceedings gress, 2169–2179.
of the 22nd International Ornithological Congress, Durban, Kishkinev, D., Chernetsov, N., & Mouritsen, H. (2010).
Johannesburg: BirdLife South Africa, pp. 2366–2378. A double-clock or jetlag mechanism is unlikely to be
Gwinner, E. (2001). Diversity and complexity of avian circa- involved in detection of east–west displacements in a long-
dian systems. In K. Honma & S. Honma (Eds.), Zeitgebers, distance avian migrant. Auk, 127, 773–780.
entrainment and masking of the circadian system Kramer, G. (1959). Recent experiments on bird orientation.
(pp. 201–213). Sapporo: Hokkaido University Press. Ibis, 101, 399–416.
Gwinner, E., & Brandstätter, R. (2001). Complex bird clocks. Krüll, F., Demmelmeyer, H., & Remmert, H. (1985). On the
Philosophical Transactions of the Royal Society of London, circadian rhythm of animals in high polar latitudes.
Series B: Biological Sciences, 356, 1801–1810. Naturwissenschaften, 72, 197–203.
Gwinner, E., Hau, M., & Heigl, S. (1997). Melatonin: Genera- Kumar, V., Rani, S., & Singh, B. (2006). Biological clocks help
tion and modulation of avian circadian rhythms. Brain reduce the physiological conflicts in avian migrants. Journal
Research Bulletin, 44, 439–444. of Ornithology, 147, 281–286.
Gwinner, E., & Helm, B. (2003). Circannual and circa- Kumar, V., Singh, B. P., & Rani, S. (2004). The bird clock:
dian contributions to the timing of avian migration. In A complex, multi-oscillatory and highly diversified system.
P. Berthold, E. Gwinner & E. Sonnenschein (Eds.), Biological Rhythm Research, 35, 121–144.
Avian migration (pp. 81–95). Heidelberg: Springer- Lessells, C. M., & Boag, P. T. (1987). Unrepeatable
Verlag Berlin Heidelberg. repeatabilities: A common mistake. Auk, 104, 116–121.
478

Leyrer, J., Spaans, B., Camara, M., & Piersma, T. (2006). Palmer, J. D. (2000). The clocks controlling the tide-associated
Small home ranges and high site fidelity in red knots rhythms of intertidal animals. BioEssays, 22, 32–37.
(Calidris c. canutus) wintering on the Banc d’Arguin, Piersma, T. (2002). When a year takes 18 months: Evidence for
Mauritania. Journal of Ornithology, 147, 376–384. a strong circannual clock in a shorebird. Naturwissenschaften,
Lourenço, P. M., Kentie, R., Schroeder, J., Groen, N. M., 89, 278–279.
Hooijmeijer, J. C. E. W., & Piersma, T. (2011). Repeatable Piersma, T. (2003). “Coastal” versus “inland” shorebird
timing of northward departure, arrival and breeding in species: Interlinked fundamental dichotomies between their
black-tailed godwits Limosa l limosa, but no domino effects. life- and demographic histories? Wader Study Group Bulletin,
Journal of Ornithology, 152, 1023–1032. 100, 5–9.
Martin, G. R. (1990). Birds at night. London: Poyser. Piersma, T., Brugge, M., Spaans, B., & Battley, P. F. (2008).
McMillan, J. P. (1972). Pinealectomy abolishes the circadian Endogenous circannual rhythmicity in body mass, molt,
rhythm of migratory restlessness. Journal of Comparative and plumage of great knots (Calidris tenuirostris). Auk,
Physiology A: Neuroethology, Sensory, Neural, and Behav- 125, 140–148.
ioral Physiology, 79, 105–112. Piersma, T., Klaassen, M., Bruggemann, J. H., Blomert, A.-M.,
Menaker, M. (Ed.), (1971). Biochronometry. Washington, DC: Gueye, A., Ntiamoa-Baidu, Y., et al. (1990). Seasonal timing
Academy of Sciences. of the spring departure of waders from the Banc d’Arguin,
Menaker, M. (2006). Circadian organization in the real world. Mauritania. Ardea, 78, 123–134.
Proceedings of the National Academy of Sciences of the Piersma, T., Koolhaas, A., Dekinga, A., & Gwinner, E. (2000).
United States of America, 103, 3015–3016. Red blood cell and white blood cell counts in sandpipers
Menaker, M., Moreira, L. F., & Tosini, G. (1997). Evolution of (Philomachus pugnax, Calidris canutus): Effects of captivity,
circadian organization in vertebrates. Brazilian Journal of season, nutritional status and frequent bleedings. Canadian
Medical and Biological Research, 30, 305–313. Journal of Zoology, 78, 1349–1355.
Mendes, L., Piersma, T., & Hasselquist, D. (2006). Two Piersma, T., van Gils, J., & Wiersma, P. (1996). Family
estimates of the metabolic costs of antibody production in Scolopacidae (sandpipers, snipes and phalaropes). In J. del
migratory shorebirds: Low costs, internal reallocation, or Hoyo, A. Elliott & J. Sargatal (Eds.), Handbook of the birds
both? Journal of Ornithology, 147, 274–280. of the world. Hoatzin to auks (Vol. 3), (pp. 444–533).
Miché, F., Vivien-Roels, B., Pévet, P., Spehner, C., Robin, J. P., Barcelona: Lynx Edicions.
& Le Maho, Y. (1991). Daily pattern of melatonin secretion Pinet, P., Jaeger, A., Cordier, E., Potin, G., & Le Corre, M.
in an Antarctic bird, the emperor penguin, Aptenodytes (2011). Celestial moderation of tropical seabird behavior.
forsteri: Seasonal variations, effect of constant illumination PLoS One, 6, e27663.
and of administration of isoproterenol or propranolol. Gen- Pittendrigh, C. S., & Daan, S. (1976). A functional analysis of
eral and Comparative Endocrinology, 84, 249–263. circadian pacemakers in nocturnal rodents. V. Pacemaker
Miyazaki, Y., Nisimura, T., & Numata, H. (2012). Circannual structure: A clock for all seasons. Journal of Comparative
rhythm in the varied carpet beetle, Anthrenus verbasci. In Physiology, 106, 333–355.
R. G. Foster, A. Kalsbeek, M. Merrow & T. Roenneberg Pohl, H. (1999). Spectral composition of light as a Zeitgeber
(Eds.), Neurobiology of circadian timing. Amsterdam: for birds living in the high arctic summer. Physiology &
Elsevier. Behaviour, 67, 327–337.
Mrosovsky, N., & Hattar, S. (2005). Diurnal mice (Mus Pohl, H. (2000). Circadian control of migratory restlessness
musculus) and other examples of temporal niche switching. and the effects of exogenous melatonin in the brambling
Journal of Comparative Physiology A: Sensory Neural and (Fringilla montfiringilla). Chronobiology International, 17,
Behavioral Physiology, 191, 1011–1024. 471–488.
Mukhin, A., Grinkevich, V., & Helm, B. (2009). Under cover Pulido, F., & Berthold, P. (2003). Quantitative genetic analysis
of darkness: Nocturnal life of diurnal birds. Journal of of migratory behavior. In P. Berthold, E. Gwinner & E.
Biological Rhythms, 24, 225–231. Sonnenschein (Eds.), Avian Migration (pp. 53–77). Berlin/
Murakami, N., Takanori, K., Nakahara, K., Nasu, T., & Heidelberg: Springer-Verlag Berlin Heidelberg.
Shiota, K. (2001). Effect of melatonin on circadian rhythm, R Core Development Team, (2009). R: A language and envi-
locomotor activity and body temperature in the intact house ronment for statistical computing. Vienna, Austria: R Foun-
sparrow, Japanese quail and owl. Brain Research, 889, dation for Statistical Computing.
220–224. Ramenofsky, M., Cornelius, J. M., & Helm, B. (2012). Physio-
Newton, I. (2008). The migration ecology of birds. London: logical and behavioural responses of migrants to environ-
Academic Press. mental cues. Journal of Ornithology, (in press).
Newton, I. (2011). Obligate and facultative migration in birds: Rattenborg, N. C., Mandt, B. H., Obermeyer, W. H.,
Ecological aspects. Journal of Ornithology, (in press). Winsauer, P. J., Huber, R., Wikelski, M., et al. (2004).
479

Migratory sleeplessness in the white-crowned Sparrow Terrill, S. B. (1987). Social dominance and migratory restless-
(Zonotrichia leucophrys gambelii). PLoS Biology, 2, ness in the dark-eyed junco (Junco hyemalis). Behavioral
924–936. Ecology and Sociobiology, 21, 1–11.
Rees, E. C. (1989). Consistency in the timing of migration for Tessmar-Raible, K., Raible, F., & Arboleda, E. (2011).
individual Bewick’s swans. Animal Behavior, 38, 384–393. Another place, another timer: Marine species and the
Reierth, E., Van’t Hof, T. J., & Stokkan, K.-A. (1999). Sea- rhythms of life. BioEssays, 33, 165–172.
sonal and daily variations in plasma melatonin in the High- van de Kam, J., Ens, B. J., Piersma, T., & Zwarts, L. (2004).
Arctic Svalbard ptarmigan (Lagopus mutus hyperboreus). Shorebirds. An illustrated behavioural ecology. Utrecht:
Journal of Biological Rhythms, 14, 314–319. KNNV Publishers.
Rowan, W. (1926). On photoperiodism, reproductive periodic- van Gils, J. A., Spaans, B., Dekinga, A., & Piersma, T. (2006).
ity, and the annual migrations of birds and certain fishes. Foraging in a tidally structured environment by red
Proceedings of the Boston Society of Natural History, 38, knots (Calidris canutus): Ideal, but not free. Ecology, 87,
147–189. 1189–1202.
Santos, C. D., Miranda, A. C., Granadeiro, J. P., van Noordwijk, A., Pulido, F., Helm, B., Coppack, T.,
Lourenço, P. M., Saraiva, S., & Palmeirim, J. M. (2010). Delingat, J., Dingle, H., et al. (2006). A framework for the
Effects of artificial illumination on the nocturnal foraging study of genetic variation in migratory behaviour. Journal
of waders. Acta Oecologica, 36, 166–172. of Ornithology, 147, 221–233.
Schüz, E. (1971). Grundriss der Vogelzugskunde. Berlin/Ham- Van’t Hof, T. J., & Gwinner, E. (1996). Development of post-
burg: Parey. hatching melatonin rhythm in zebra finches (Poephila gut-
Schwabl, H., Gwinner, E., Benvenuti, S., & Ioalé, P. (1991). tata). Experientia, 52, 249–252.
Exposure of dunnocks (Prunella modularis) to their previ- Van’t Hof, T. J., Gwinner, E., & Wagner, H. (1998). A highly
ous wintering site modifies autumnal activity pattern: Evi- rudimentary circadian melatonin profile in a nocturnal bird,
dence for site recognition? Ethology, 88, 35–45. the barn owl (Tyto alba). Naturwissenschaften, 85, 402–404.
Sharp, P. J. (2005). Photoperiodic regulation of seasonal Visser, M. E., Caro, S. P., van Oers, K., Schaper, S. V., &
breeding in birds. Annals of the New York Academy of Helm, B. (2010). Phenology, seasonal timing and circannual
Sciences, 1040, 189–199. rhythms: Towards a unified framework. Philosophical
Silverin, B., Gwinner, E., Van’t Hof, T. J., Schwabl, I., Transactions of the Royal Society B: Biological Sciences,
Fusani, L., Hau, M., et al. (2009). Persistent diel melatonin 365, 3113–3127.
rhythmicity during the Arctic summer in free-living willow Wikelski, M., Tarlow, E., Eising, C., Groothuis, T., & Gwinner, E.
warblers. Hormones and Behavior, 56, 163–168. (2006). Do night-active birds lack daily melatonin rhythms? A
Smit, B., Boyles, J. G., Brigham, R. M., & McKechnie, A. E. case study comparing a diurnal and a nocturnal-foraging gull
(2011). Torpor in dark times: Patterns of heterothermy are species. Journal of Ornithology, 147, 107–111.
associated with the lunar cycle in a nocturnal bird. Journal Wingfield, J. (2005). Flexibility in annual cycles of birds:
of Biological Rhythms, 26, 241–248. Implications for endocrine control mechanisms. Journal of
Székely, T., Thomas, G. H., & Cuthill, I. C. (2006). Sexual con- Ornithology, 146, 291–304.
flict, ecology, and breeding systems in shorebirds. BioSci- Wirz-Justice, A., Kräuchli, K., & Graw, P. (2001). An underly-
ence, 56, 801–808. ing circannual rhythm in seasonal affective disorder? Chro-
Tarlow, E. M., Hau, M., Anderson, D. J., & Wikelski, M. nobiology International, 18, 309–313.
(2003). Diel changes in plasma melatonin and corticosterone Yohannes, E., Valcu, M., Lee, R. W., & Kempenaers, B.
concentrations in tropical Nazca boobies (Sula granti) in (2010). Resource use for reproduction depends on spring
relation to moon phase and age. General and Comparative arrival time and wintering area in an arctic breeding shore-
Endocrinology, 133, 297–304. bird. Journal of Avian Biology, 41, 580–590.
Subject Index

Note: Page numbers followed by “f” indicate figures, and “t” indicate tables.

ACTH. See Adrenocorticotrophin (ACTH) role, 120


Adenosine monophosphate-activated protein SCN, 121
kinase (AMPK) activation, 164–165, 187 VIP, 132–133
Adrenocorticotrophin (ACTH), 166, 171 Autonomic nervous system (ANS), 184, 185, 189
ANS. See Autonomic nervous system (ANS) Avian migration. See Temporal multitasking
Anteroventro-periventricular (AVPV) nucleus, AVP. See Arginine vasopressin (AVP)
Kp neurons
circadian oscillator, 427 BAC. See Bacterial artificial chromosome (BAC)
circadian signal, 425–427, 426f Bacterial artificial chromosome (BAC), 251,
estrogen signal, 425, 426f 257–258
female rodents, 425, 426f
RFRP-3, 427–428 cAMP. See Cyclic adenosine monophosphate
Anthrenus verbasci, circannual rhythm. (cAMP)
See Circannual rhythm, Anthrenus verbasci Cardiotrophin-like cytokine (CLC)
Arcuate (ARC) nucleus, Kp neurons PK2, 124
melatonin-binding sites, 428 receptor, 123
RFRP neurons, 430–432 rhythm, 123
seasonal reproduction role, 123
GnRH and and gonadotropin release, 429 TGF-a, 122–123
Kiss1 expression, 430 Cardiovascular function and circadian disruption
melatonin regulation, 429–430, 429f cardiomyopathic hamsters, 346
Siberian hamsters, 430 hypertension
T3/RFRP/Kp network, 429–430, 431f clock genes, 347
Syrian hamster, 428 SCN dysfunction, 346–347
T3/RFRP/Kp network, 432–433 misalignment and clock gene mutations,
TSH-driven regulation, tanycyte deiodinases, 345–346
432 shift work, 346
Arginine vasopressin (AVP) Tau mutant hamsters, 346
circadian rhythms, 133 Central and peripheral clock interaction
deficiency, 120–121 ACTH, 171
description, 120 communication routes, circadian timing system,
GnRH, 121 166–167, 167f

481
482

Central and peripheral clock interaction intersubject and intrasubject confound,


(Continued) 386–387
daily fluctuation, 166 PVT and recovery rate, 387–388
description, 163–164 tasks and sleep loss, 386–387
FEO, 169 driving performance, 386
fibroblasts, 170 human performance, regulation
GC secretion (see Glucocorticoid (GC) homeostatic and circadian process, 383
secretion) sleep inertia, 383
glucose metabolism waking performance and sleep propensity,
clock-deficient animal models, 173 383–384
euglycemia, 174 sleepiness and fatigue, 383
hepatic sympathetic denervation, 173 sleep propensity, 384
insulin sensitivity, 173–174 Circadian clocks
maintenance, 172 description, 41–42
neurons, 174 development
orexin, 172–173 aanat2 expression, 46
immortalized rat fibroblast (Rat-1) cells, 170 blastula, 47
MASCO, 169–170 bmal mRNA expression, 47
mediobasal hypothalamus (MBH), 168 circadian rhythms, 46–47
molecular clockworks DD, 46–47
AMPK activation, 164–165 per1b expression, 46–47
circadian pacemaker, SCN light-entrainable peripheral clocks, 47–53
(see Suprachiasmatic nuclei (SCN)) zebrafish (see Zebrafish)
orchestration, 165 Circadian output signals
redox status, 164–165 AVP, 120–121
TTL, 164 dye tracing experiments, 120
pathways, 167f, 170 neuronal firing, 123
Per gene expression, 171 PK2, 121–122
PVN, 166 SCN, 119–120
retina, 167–168 TGF-a and CLC, 122–123
reward system controls and animal behavior, Circadian rhythms
168–169 description, 184
tissue, transplantation, 165–166 nongenetic clock
TTL, 166 AMPK activation, 187
Chronic partial sleep restriction cAMP, 187
accidents and medical errors, 383 heme, 187–188
behavioral alertness, 384–386 NAD, 187
causes peroxiredoxins (PRX), 188
chronotype, 382–383 posttranslational modifications, 186–187
commute/travel, 382 oscillators, 184
leisure, 382 regulation, peripheral tissues, 184–185
paid work, 380–382 SCN (see Suprachiasmatic nuclei (SCN))
cognitive performance social interactions
attention and working memory, 388 abiotic factors, 267
executive function and deficits, 388 Acomys cahirinus and A. russatus, 269
483

American mink, 269 constant routine and forced desynchrony


animals, 273 protocols, 340–341
assays, 274 coronary artery ischemia, 345
Coda, 276 epinephrine and norepinephrine, 344–345
description, 267–268 morning peak and behavioral changes, 340
environment, 273–274 parasympathetic nervous activity, markers,
experimental design, 274–275 345
field and laboratory, 271–272 platelets, 341–344
Gerbillus allenbyi and G. pyramidum, 269 relative frequency histogram, 339–340, 340f
honeybees, 270 resting heart rate, 341, 342f
maternal entrainment, 271 standardized bicycle exercise, 341, 343f
mixed sex interactions, 272 light exposure, 339
Mongolian gerbil, 269 metabolic function (see Metabolic function and
mutualism, 268 circadian disruption)
Nile grass rat, 275 and nutrition (see Nutrition and circadian
parasitism, 268–269 timing system)
predator–prey, 270 SCN and circadian oscillators, 338
wavelet transforms, 275 western societies, 337–338
transcriptional–translational feedback model, Circannual clock hypothesis, 445
185–186 Circannual rhythm, Anthrenus verbasci
WAT (see White adipose tissue (WAT), adenohypophysis, 452
circadian rhythms) biological clocks, mechanisms, 439–440
Circadian system characteristics and environmental cycles, 451–452
clocks, 84 circannual clock hypothesis, 445
distribution, ontogenesis, 97 description, 440
LD cycle, 83–84 DNA microarrays and RNAi, 453
peripheral clocks, 94–95 entrainability, 444–445
SCN, 84 environmental cycles, 439–440
utero, 84–92 frequency demultiplication hypothesis, 445–446
Circadian thermoenergetics (CTE) Japanese population, 440
hypothesis molecular mechanisms, 453
energy balance, 297–298 photoperiodism, entrainment
LD cycle, 298–299 circadian clock, 450–451
proximate mechanisms, 296–297 natural day length, 449–450
temperatures, 297–298, 298f, 299 physiological components, 452, 452f
ultimate mechanism PRCs and biological oscillator, 447–449
ambient temperatures, 297 properties, 440
mice, food intake, 297, 298f self-sustainability
Circadian timing system constant light/darkness, 441–442, 442f
cardiovascular function (see Cardiovascular frequency, pupation times, 441, 441f
function and circadian disruption) inter-and intraindividual variability, 442–443
core clock genes, 338–339 Junco hyemalis and Sturnus vulgaris, 442–443
description, 338 sequence of linked stages hypothesis, 446–447
internal clock and adverse cardiovascular Tamias sibiricus and Ovis aries, 452–453
events temperature compensation, 443–444
484

Clock gene functions, 139 synaptic overload, 221–222


Cryptochromes (CRYs) thalamus-driven process, 223
Drosophila Timeless protein, 234–236 topographical modifications, 222
PERs, 236 description, 220
CRYs. See Cryptochromes (CRYs) implication, medicine
Cyclic adenosine monophosphate (cAMP), 187 clinical practice, 229
dissociation and sleep inertia, 229
Deep brain photoreceptors, 418 hippocampus, 228
Disc1. See Disrupted-in-Schizophrenia-1 (Disc1) neuronal tiredness, 228
Disk-over water methodology, 377–378 NREM sleep, 229
Disrupted-in-Schizophrenia-1 (Disc1), performance, 227
259–260 sleep-like and wake-like activity, 229
Drosophila state-synchronization, 228–229
Clk-gal4, 62–64 Enhanced green fluorescence protein (EGFP),
CRY, 64 121–122
gal4 and uas, 62 EUMODIC. See European Mouse Disease Clinic
LNs and DNs, 62 (EUMODIC)
PDF, 62 European conditional mouse mutagenesis
Pdf-gal80, 64–65 (EUCOMM ) program, 256, 260–261
PER/TIM, 62–64 European Mouse Disease Clinic (EUMODIC),
Dual oscillator model 260–261
Drosophila, 61–65 Evening (E) cells
E cells, 75–76 dominance, 66
LD cycles, 61 light activates output, 66–67
M cells, 76 light decelerates, 67–70
phase angle between M and E peak, 61 oscillators, 70–72
photoperiods, 61
SCN, 61 Familial advanced sleep-phase syndrome
(FASPS), 237f, 241, 257–258
EGFP. See Enhanced green fluorescence protein FASPS. See Familial advanced sleep-phase
(EGFP) syndrome (FASPS)
Electroencephalographic (EEG), local sleep Fast Fourier Transform (FFT) analysis, 134–135,
behavioral transition 136f
cortical neurons, 221 FEO. See Food-entrainable oscillator (FEO)
deactivation, thalamus, 222 Food anticipatory activity (FAA), 50–53
delta activity, 222 Food-entrainable oscillator (FEO), 169
firing pattern, 221 Forward genetics, mouse mutant generation
intracerebral and scalp, recordings, 222–223, description, 248
224f ENU (see N-ethyl-N-nitrosourea (ENU),
NREM sleep (see Non-rapid eye movement mouse mutant generation)
(NREM) sleep) Mus musculus, 248
processes, 220–221 spontaneous mutations, 248–249
slow-delta and sigma frequency range, transversions, 249
222–223, 225f X-ray, 249
SWA and hippocampus, 221–222 Frequency demultiplication hypothesis, 445–446
485

GC secretion. See Glucocorticoid (GC) secretion IKMC. See International Knockout Mouse
Gene knockout, mouse mutant generation Consortium (IKMC)
description, 254–255 IMPC. See International Mouse Phenotyping
inducible and tissue-specific Consortium (IMPC)
gene expression, liver, 256 Individual shift-work load (ISL), 403–404
loxP sites and Cre recombinase, 256 Inner plexiform layer (IPL), 20–21
resources, 256–257, 257t Intergeniculate leaflet (IGL), 317–318
nonparalogous genes, 255–256 International Knockout Mouse Consortium
paralogous genes (IKMC), 257t, 261
Clock and Npas2, 255 International Mouse Phenotyping Consortium
Dec1 and Dec2, 255 (IMPC)
Drosophila Per, 255 challenges, 262
PAR bZIP transcription factors, 255 description, 261
transgenics forward genetics, 261–262
Cry1 overexpression, 258 Intracerebroventricular (ICV) delivery, 121
determination, 258 ISL. See Individual shift-work load (ISL)
FASPS, 257–258
tetracycline transactivator (tTA), 258 Kisspeptin (Kp) neurons
Glucocorticoid (GC) secretion ARC nucleus (see Arcuate (ARC) nucleus, Kp
Addison’s disease, 170–171 neurons)
cortex and medulla, 171 AVPV nucleus (see Anteroventro-
disruption, 170–171 periventricular (AVPV) nucleus,
neuronal signals, 171 Kp neurons)
nycthemeral production, 171 gonadotropic axis
receptors, 171–172 GnRH, 423
GnRH. See Gonadotropin-releasing hormone Kiss1 gene, 423–424, 424f
(GnRH) photoneuroendocrine pathway, 424–425
Gonadotropin-releasing hormone (GnRH) sex steroids and metabolic factors, 424
and AVP, 121 Knockout mouse project (KOMP), 256, 260–261
description, 423 KOMP. See Knockout mouse project (KOMP)
secretion and tanycytes, 432
stimulatory effect, Kp, 429 Lateral hypothalamic area (LHA), 320
gpr50. See G protein-coupled receptor Lateral neurons (LNs), 65
50 (gpr50) LHA. See Lateral hypothalamic area (LHA)
G protein-coupled receptor 50 (gpr50) Light/dark (LD) cycle, 95–97
energy homeostasis, 420 Light-entrainable peripheral clocks
melatonin receptors, evolution, 419–420 bioluminescent, 47–48
G-protein-coupled receptor kinases (GRKs), 32 cavefish
G-protein signaling pathway description, 49–50
Gq/11-type G-proteins, 25–27, 26t FAA, 50–53
phospholipase C activation free-running period, 50–53
PLC, 27 LD cycles, 50–53
PLCb1 and PLCb3, 27–28 opsin genes, 53
cycling clock gene expression, 47–48
IGL. See Intergeniculate leaflet (IGL) gene expression
486

Light-entrainable peripheral clocks (Continued) glucocorticoid signaling, 320–321


D box-binding factors, 49, 50f molecular basis, 316–317
light input pathway, 49, 51f NPY and IGL, 317–318
MAPK signaling, 48–49 orexin and LHA, 320
Per2, 49 serotonin and raphe nuclei, 318–320
SCN, 48–49 photic entrainment
light exposure, 47–48 glutamate role, 308
peripheral photoreceptors, 48 negative masking, light, 308
photoreceptors, 48 phase shift data, laboratory rodent, 306–307,
Local sleep 307f
behavioral transition, 220–227 physiological correlates, 315–316
description, 220 research techniques, 324–325
EEG (see Electroencephalographic (EEG), species generalization, 324
local sleep) zeitgeber vs. zeitnehmer, 325–326
implication, medicine, 227–230 Luminance and temporal niche switching
Locomotor activity, feedback actions effects, moonlight, 288–289
aging, 322 Mediterranean lobster, 288
arousal and motivation, 315
benzodiazepine-induced circadian effects, 309 Mammalian circadian clock
circadian phase, alteration, 314–315 conceptual model
circadian responses, discrete “pulses,” 309–310 assumptions, implications, and uncertainties,
core clock gene mutations, 322–323 9–14
endogenous circadian oscillators, 305–306 cones, role, 10–11
entrainment to scheduled locomotor activity melanopsin, 11–12
free-running rhythms, 312 photoreceptor, 12–14
mouse model, 312–313, 314f rods, cones and melanopsin, 9, 10f
phase angle and nonphotic PRC, 313–314 cones, 7–8
running wheel and tau shortening, 312 description, 1
signaling mutants, 323–324 environment, 1–2
extra-SCN oscillators, 324 ipRGCs, 2
interactions, nonphotic and photic entraining melanopsin, 8–9, 11–12
stimuli photoentrainment pathway, 1–3, 2f
dark pulses, 310–311 photoreceptor, 3
light–dark (LD) cycle shift and conflict red-cone knockin, 4, 6f
studies, 311–312 RHT, 14
laboratory rodents, 309 SCN, 3
mammalian circadian system, 306, 306f Mammalian nocturnal bottleneck, temporal niche
nonphotic entrainment switching
activity/arousal-associated stimuli, 309 arguments, 285–286
description, 308 ectothermic predators and competitors, 285
diurnal species, 321–322 evidence, 287
restricted feeding and food-entrainable extinction, dinosaurs, 285
oscillator, 308–309 factors, 285
nonphotic phase shifting and entrainment MASCO. See Methamphetamine-sensitive
electrical output, SCN, 317 oscillator (MASCO)
487

Meal patterns and human health gpr50, 419–420


breakfast, 367 subtypes, 419
clinical nutrition, 367–368 rhythmical gene expression, 417–418
day and night secretion, 416–417
NES and clinical trials, 366 Metabolic function and circadian disruption
shift workers and short sleepers, 366 adverse endocrine effects, night work, 351
wakefulness, food intake and anabolic human endogenous circadian rhythms, 347–348
metabolic processes, 365–366, 366f impaired glucose metabolism, 350
enteral tube feeding, 368 isocaloric meals and leptin, 348–349
meal frequency, 366–367 leptin, glucose and insulin, 348, 349f
parenteral nutrition, 368–369 lipopolysaccharide (LPS) challenge, 349–350
Melanopsin, 8–9 molecular clock genes, 350–351
Melanopsin phototransduction organ-specific clocks, 351
Gb subunits, 33 SCN, 347–348
GMP, 23 shift work, 348
G-protein signaling pathway, 25–27 Methamphetamine-sensitive oscillator (MASCO),
ion channels and potentials firing, 29–30 169–170
light absorption Morning and evening oscillator model
GPCR, 24–25 ablate clock neurons, 65
Opn4L and Opn4S, 25 complexity, 76–77
spectrophotometry analysis, 25 dominance, 66
PKA, 30–32 dual oscillator model, 61–62
pRGCs, 19–20 flies adaptation
rods and cones, 23 clock, 72
scaffold proteins, 33 dawn and dusk, 72
signaling pathway, 23, 24f PDF-positive l-LNvs, 72–73
TRPC6 and TRPC7 channels, 30 fruit flies, 65
TRP channel activation, 28–29 gal4, 65–66
Melatonin LNs, 65
actions, vertebrate lineages temperature and photoperiod influence, 59–60,
circadian coordination, 419 60f
extra-retinal photoreception, 418–419 temperature effects
receptor distribution, 419 flies, natural-like temperature cycles, 74
cycles, captive waders light and, 74–75
daily profiles, plasma melatonin, 468–471, Morning (M) cells
472f dominance, 66
diel melatonin profiles, 468–471, 470f inhibit output, 66–67
feeding activity, 468 light accelerates, 67–70
group activity patterns, 471, 474f, 475t oscillators, 70–72
immuno-enhancement, 467 Mouse mutant generation
life-cycle stages, derivation, 467–468, 469f description, 247–248
light availability and physiological functions, EUCOMM, KOMP and IKMC, 261
471–473 EUMODIC, 261
Ruff vs. Red Knot, 466–467 forward genetics (see Forward genetics, mouse
receptors, divergent evolution mutant generation)
488

Mouse mutant generation (Continued) cyclical appearance, 225–226, 227f


gene knockout, reverse genetics (see Gene description, 223–225
knockout, mouse mutant generation) intracerebral and scalp, recordings, 225–226,
IMPC, 261–262 226f
Multi-electrode array dish (MED), 129–130 sensory-motor processing, 226
Mutualism, 268 time–frequency analysis, 226, 228f
Process S, 210
NAD. See Nicotinamide adenine dinucleotide SCN neuronal activity, 208–210
(NAD) NPY. See Neuropeptide Y (NPY)
Nanda–Hamner protocol, 450–451 NREM sleep. See Non-rapid eye movement
NES. See Night eating syndrome (NES) (NREM) sleep
N-ethyl-N-nitrosourea (ENU), mouse mutant Nutrition and circadian timing system
generation central clock, 360
description, 249 digestion and metabolism
gene-driven approach, reverse genetics carbohydrates, 363–365
description, 259, 260f food intake, 361–363
Disc1 and Srr, 259–260 gastrointestinal system, 363
identification, 259 lipids, 365
mutation detection, DNA, 259, 260f disturbed rhythms and type 2 diabetes, 369
real-time monitoring assays, 259 food availability, 359–360
mapping and positional cloning meal patterns and human health (see Meal
BAC, 251 patterns and human health)
candidate region, 251 peripheral clocks
principles, 251 autonomic cycle, 361
redundancy and epistasis, investigation, 254 gene oscillations, 360
screening physiological function, 360–361
affect, clock circuitry, 254 SCN, 361, 362f
Drosophila, 253–254
Fbxl3, 253 Ontogenetic development
flybase, 253–254 circadian systems, 105–106
identification, Clock gene, 251–253 endogenous rhythm generator
phenodeviants, circadian phenotypes, AANAT activity, 108–109
250–251 adrenal gland, 111
strategies, 249–250 Avp expression, 108
Neuropeptide Y (NPY), 317–318 description, 108
Nicotinamide adenine dinucleotide (NAD), 187 environmental lighting, 110
Night eating syndrome (NES), 366 feeding and social interaction, 110
Non-rapid eye movement (NREM) sleep liver, 111
cortical neurons, 221 Per2 and MT1, 109–110
Cry1 and Cry2 double-knockout mice, 210–211 peripheral oscillators, 110
description, 204 PT, 110–111
2 h/2 h protocol, 212–214, 213f SCN, 108
homeostatic process, 214 Per and Cry genes, 105–106
motor cortex (Mc) local activations SCN, 105–106
cortical deactivation, 226–227 synchronization mechanisms
489

HPA, 107–108 phase resetting and state variables, 447–448


melatonin, 107 2-week long-day pulse, 447
Per expression, 106–107 4-week long-day pulse, 447, 449f
PVN, 107 nonphotic, 313–314
SCN, 107 Phase-shifting capacity, SCN neurons
adjustment, 152–153
Parasitism, 268–269 constant darkness, 151
Paraventricular nucleus (PVN), 107–108, 166, coupled oscillator theory, 152
168, 171, 172, 184, 185, 190f curve fitting analysis, 151–152, 152f
Pars tuberalis (PT) desynchrony, advances, 153
melatonin, 110–111 differences, synchronization, 153
SCN, 110–111 GABA, 151
PCEs. See Photoreceptor conserved elements peaks, multiunit activity rhythm, 151
(PCEs) resynchronization, 153
PERIOD2 (PER2) protein shift work and jet lag, description, 150–151
brain, 239–240 Photoperiodism, vertebrates
circadian oscillator deep brain photoreceptors and birds, 418
histone deacetlyases (HDAC), 236–237 energy budgets, life history strategies and
nonrhythmic overexpression, 236–237 seasonality
protein–protein interaction regions, 236–237, breeding and offspring survival, 413–414
237f circadian rhythm generators, 414
CRYs, 236 environmental variability, 414
cyanobacteria, 233–234 species survival, 413
history, 234–236 gpr50 and energy homeostasis, 420
liver metabolism melatonin
blood glucose concentration, 238–239 actions, vertebrate lineages, 418–419
detoxification enzymes, 238–239 receptors, divergent evolution, 419–420
nuclear receptors, 239, 240f photoperiodic axis, mammals
mechanisms, 234 Eya3-dependent transcription, TSHb,
posttranslational modifications, 241 417–418, 417f
sleep, 240–241 hypothalamic deiodinase expression, 416–417
stabilizing loop, 238 melatonin secretion, 416–417
synchronization, 242 thyroid hormone (TH) signaling
Zeitgebers, 233–234 deiodinase gene expression, 416
PER2 protein. See PERIOD2 (PER2) protein endostyle and comparative genomics
Phase-resetting capacity, SCN neurons approaches, 415–416
description, 150 evolutionary appearance, elements, 415–416,
distribution, PRCs, 149–150, 150f 416f
low-and high-amplitude rhythm, 149 metabolism, hypothalamus, 414–415, 415f
NMDA receptors, 149 nutrient-sensing pathway, 415–416
short and long photoperiods, 149 Photoreceptor
Phase response curves (PRCs) function, 4
biological oscillator in vitro, 12
circadian and circannual rhythms, 447, 448f photoentrainment, 12
Cricetus cricetus, 448–449 rod phase, 13
490

Photoreceptor (Continued) circadian timing system, 400


saturation point, 13 description, 406–407
sensitivity ranges, 13–14 health problems, 399–400
Photoreceptor conserved elements (PCEs), 43–44 internal time, 400–401
Photosensitive retinal ganglion cell (pRGCs) laboratory studies
functional differences, 21–22 advantages, 404–406
IPL, 20–21 disadvantages, 406
melanopsin expressing, 20–21, 20f real-life settings
M1-type cells, 20–21 cortisol awakening response and stress,
physiological responses, 22–23 402–403
retinal connections, 22 ISL, 403–404
variable responses, 34–35 shift type and incidence rate, 404
PKA. See Protein kinases activity (PKA) stress and adverse effects, health, 401–402
PRCs. See Phase response curves (PRCs) “shift-clock-work,” 407f, 407
pRGCs. See Photosensitive retinal ganglion cell Sleep restriction
(pRGCs) chronic partial (see Chronic partial sleep
Primates, temporal niche switching, 287–288 restriction)
Prokineticins (PK2) description, 378
description, 121 disk-over water methodology, 377–378
EGFP, 121–122 healthy human sleep
ICV delivery, 122 circadian process, 379
rhythm, 122 duration, 379–380
Protein kinases activity (PKA) homeostasis, 378–379
cAMP pathway, 32 states and characteristics, 378
DAG, 30–32 interindividual differences
GRKs, 32 biomarkers and behavioral predictors, 390–391
PKCz, 32 genetic polymorphisms, 390
TRPC6 and TRPC7, 30–32 neurobehavioral performance, 388–389, 389f
Psychomotor Vigilance Test (PVT), 387–388 nocturnal sleep, 390–391, 391f
PVN. See Paraventricular nucleus (PVN) task dependent and cognitive performance,
PVT. See Psychomotor Vigilance Test (PVT) 389–390
neurobehavioral impairment, management,
Retinohypothalamic tract (RHT) activity, 8 391–392
RF-amide-related peptide (RFRP) Sleep, rodents
mediobasal hypothalamus, 432–433 circadian–homeostatic interactions
neurons, 430–432 conceptual model, 208–210
RFRP-3 and circadian regulation, 427–428 phase-shifting experiments, humans, 208
RFRP. See RF-amide-related peptide (RFRP) pressure, changes, 208
Rodents, temporal niche switching, 287 SCN neuronal activity, 208–210
time course, Process S, 210
SCN. See Suprachiasmatic nuclei (SCN) clock genes
Sequence of linked stages hypothesis, 446–447 Cry1 and Cry2, 210–211
Serine racemase (Srr), 259–260 description, 210
Shift-work research mPer3, 211
chronotype-specific differences, 400–401 mPer1 and mPer2, 211
491

clocks, photoperiod, running wheels and light, Per1 and Per2 expressions, 93
207–208 rabbit pups, 93–94
comparison, mouse strains, 206t, 207 MED, 129–130
crepuscular behavior, 204 metabolic-related factors, 347–348
description, 204 neuron, 105–106
12 h:12 h L:D cycle, 204, 205t ontogenesis, 85, 85f
modeling, 214–215 oscillators and non-oscillator cells
protocols and techniques, 214 AVP and VIP, 132–133
separation, circadian and homeostatic Per1-luc rhythms, 131–132, 132f
processes pacemaker phase, 321
daily changes, 211 periventricular cell groups, 85–86
desynchrony protocol, 211–212 population, 92
22-h dim light–dark cycle, 212 regional oscillators
2 h/2 h short-sleep deprivation protocol, 212 bioluminescence images, 134–135, 135f
“90-min day” protocol, 212 FFT, 134–135, 136f
NREM sleep (see Non-rapid eye movement LD cycle, 133–134
(NREM) sleep) organotypic slice culture, 133
SCN lesions, 211 per-luc rhythms, 133–134, 134f
TST (see Total sleep time (TST), rodents) serotonin signaling, 319
Srr. See Serine racemase (Srr) sleep, rodents
Suprachiasmatic nuclei (SCN) Cry1,2 knockout mice, 210–211
cAMP-signaling pathway, 87 neuronal activity, 208–210
cellular oscillators and coupling subregions, 306
behavioral rhythms, 130–131 TTX, 136–138
spontaneous firing rhythms, 130, 131f VIP, 135–136
central and peripheral clock interaction (see WAT, circadian rhythms (see White adipose
Central and peripheral clock interaction) tissue (WAT), circadian rhythms)
circadian molecular clock mechanism, 86 Suprachiasmatic nuclei (SCN) neurons
circadian rhythms, regulation, 338 aging and disease, effects, 153–154
circadian timing system constant light effects
central clock, 360 coupling factors, 146–147
peripheral clocks, 361, 362f in vitro and in vivo recordings, 145–146, 146f
clock gene functions, 139 description, 143–144
description, 84 generation, circadian rhythms, 144
development, 87–88 intercellular coupling, circuit, 154–156
dysfunction and hypertension, 346–347 phase-resetting capacity (see Phase-resetting
electrical output, suppression, 317 capacity, SCN neurons)
extra-SCN oscillators, 324 phase-shifting capacity (see Phase-shifting
food intake, 363 capacity, SCN neurons)
functional network, 86 seasonal changes, waveform
in vitro, 86–87 brain slice preparation, 147–148
lesioned hamsters, 312–313 clock gene expression, in situ analysis, 147
maternal entrainment C57 mice, 147
lactation, 93 distribution, single-unit electrical activity
LD cycle, 92–93 patterns, 147–148, 148f
492

Suprachiasmatic nuclei (SCN) neurons nocturnal bottleneck (see Mammalian


(Continued) nocturnal bottleneck, temporal niche
longer duration activity, 148 switching)
photoperiodic encoding, 147 primates, 287–288
waveform, electrical activity rodents, 287
distribution, 145, 145f mechanisms
in vitro and in vivo electrophysiological circadian oscillator properties, 293–295, 295f
recordings, 144 diurnal and nocturnal degus, 293–295
onset–offset, behavioral patterns, 144–145 expression, behavioral program, 293–295,
295f
Temporal multitasking golden spiny mice, 293–295
annual time keeping locomotor activity, Drosophila melanogaster,
circannual rhythms, 461 295–296
environmental factors, 461–462 Zeitgebers, 285f, 296
long-distance migrants, 460–461 microchiroptera bats, 299
photoperiodism, 460–461 species, description, 282, 283t
southward migration pathways, 459–460, 460f visual adaptations (see Visual adaptations,
Sylvid warblers, 461 temporal niche switching)
avian long-distance migration, 458–459 Tetrodotoxin (TTX)
challenges, avian migration, 459 effects, 136–138, 137f
chronobiology and activity patterns, 457–458 resistant circadian rhythms, 138–139
daily time keeping treatment, 136–138
challenges, 463 Total sleep time (TST), rodents
circadian pacemaking system, birds, 462–463 C57BL/6 mouse, 204–205, 205t
internal-resonance model, 462–463 comparison, mouse strains, 206t, 207
melatonin role, 463–464 description, 204
nocturnal restlessness, 463 hourly distribution, activity, 207
entrainment mechanisms, 458 light vs. dark period, 207
lunar and tidal cycles, 464 mice vs. rats, 206t, 207
waders (see Waders and temporal multitasking) photoperiods, 206t, 208
Temporal niche switching running wheel, 207–208
ambient temperature Transcriptional–translational feedback loop
colder water, 289 (TTL), 164, 166
diurnal rhythms and locomotor activity, 289 TSHb. See TSH beta subunit (TSHb)
effects, 289 TSH beta subunit (TSHb)
behavioral programs, 282, 285f Eya3-dependent transcription, 417–418, 417f
CTE hypothesis (see Circadian photoperiodic signaling, birds, 418
thermoenergetics (CTE) hypothesis) TTL. See Transcriptional–translational feedback
description, 281–282 loop (TTL)
golden spiny mouse, 288, 288f
limitation, factors, 289–290 Utero
locomotor activity, 282 clocks, entrainment
luminance, 288–289 dopamine/ melatonin, 91–92
lunar cycle and light properties, 299 LD cycle, 91–92
mammals molecular clock mechanism, 89–91
493

synchronized circadian rhythms, 89–91, 90f Waders and temporal multitasking


peripheral clocks annual punctuality and daily flexibility
heart and kidney, 89 assortative mating, 465–466
hepatocytes, 88–89 behavioral observations, 466
liver, 88 breeding habitat and geolocators, 465
Rev-erba expression, 88–89 Calidris canutus and Haematopus ostralegus,
prefetal development 466
Bmall, 85 continuous polar light, 466
multipotent cells, 84–85 internal timing programs, 465
SCN, 85–88 Limosa lapponica baueri, 464–465
L. limosa islandica, 465–466
Varied carpet beetle. See Circannual rhythm, description, 464
Anthrenus verbasci melatonin cycles
Vasoactive intestinal peptide (VIP) daily profiles, plasma melatonin, 468–471,
AVP, 132–133 472f
SCN, 135–136 diel melatonin profiles, 468–471, 470f
Vertebrate core clock mechanism, zebrafish feeding activity, 468
multiple clock genes group activity patterns, 471, 474f, 475t
cry genes, 45–46 immuno-enhancement, 467
genome duplication, 45–46 life-cycle stages, derivation, 467–468, 469f
mammals, 45 light availability and physiological functions,
per genes, 45–46 471–473
new clock genes Ruff vs. Red Knot, 466–467
asparagine mutation, 45 White adipose tissue (WAT), circadian rhythms
bioluminescence, 44 adipokines, description, 194–195
Drosophila, 44 adiponectin, 194
melatonin, 44 catecholamines, 184–185
microsatellite markers, 44–45 depots, 189
Visual adaptations, temporal niche switching description, 183
ascorbic acid concentration, 292–293 development, 188
night-and day-adapted vision feeding–fasting cycle, 185
foraging, 290–291 gene expression, 192–193
mouse, cones, 286t, 291 innervation
physiological and anatomical determinants, lipodystrophy and hypothalamus, 189
290 parasympathetic, 191
reef fish, 290 sensory nerves, 191
rod-cone densities, visual acuity and sensitivity sympathetic nerve fibers, 189–191
golden spiny mice, 291–292 leptin, 193–194
ground squirrels, 291 lipogenesis, 192
macula and fovea, 292 lipolysis, 192
tapetum, 292 non-neural signals, 185
spectral filtering and UV protection obesity
coloration, lenses, 292 db/db mice, 195–196
ground squirrel, 292, 299 ob/ob mice, 195
properties, photoreceptors, 292, 293t, 294f subcutaneous WAT (sWAT), 195
494

White adipose tissue (WAT), circadian rhythms Zebrafish


(Continued) chronobiology
Zucker rat, 196 circadian clock genes, 42–43
oscillators, 184 circadian timing system, 43
pathways and timing information, 185, 186f PCEs, 43–44
plasma corticosterone concentrations, 185 pineal cells, 43
PVN and ANS, 184 embryology, 42
resistin, 194–195 vertebrate core clock mechanism,
types, adipose tissue, 188 44–46
Other volumes in PROGRESS IN BRAIN RESEARCH

Volume 149: Cortical Function: A View from the Thalamus, by V.A. Casagrande, R.W. Guillery and S.M. Sherman (Eds.) – 2005
ISBN 0-444-51679-4.
Volume 150: The Boundaries of Consciousness: Neurobiology and Neuropathology, by Steven Laureys (Ed.) – 2005,
ISBN 0-444-51851-7.
Volume 151: Neuroanatomy of the Oculomotor System, by J.A. Büttner-Ennever (Ed.) – 2006, ISBN 0-444-51696-4.
Volume 152: Autonomic Dysfunction after Spinal Cord Injury, by L.C. Weaver and C. Polosa (Eds.) – 2006, ISBN 0-444-51925-4.
Volume 153: Hypothalamic Integration of Energy Metabolism, by A. Kalsbeek, E. Fliers, M.A. Hofman, D.F. Swaab, E.J.W. Van
Someren and R.M. Buijs (Eds.) – 2006, ISBN 978-0-444-52261-0.
Volume 154: Visual Perception, Part 1, Fundamentals of Vision: Low and Mid-Level Processes in Perception, by
S. Martinez-Conde, S.L. Macknik, L.M. Martinez, J.M. Alonso and P.U. Tse (Eds.) – 2006, ISBN 978-0-444-52966-4.
Volume 155: Visual Perception, Part 2, Fundamentals of Awareness, Multi-Sensory Integration and High-Order Perception, by
S. Martinez-Conde, S.L. Macknik, L.M. Martinez, J.M. Alonso and P.U. Tse (Eds.) – 2006, ISBN 978-0-444-51927-6.
Volume 156: Understanding Emotions, by S. Anders, G. Ende, M. Junghofer, J. Kissler and D. Wildgruber (Eds.) – 2006,
ISBN 978-0-444-52182-8.
Volume 157: Reprogramming of the Brain, by A.R. Mller (Ed.) – 2006, ISBN 978-0-444-51602-2.
Volume 158: Functional Genomics and Proteomics in the Clinical Neurosciences, by S.E. Hemby and S. Bahn (Eds.) – 2006,
ISBN 978-0-444-51853-8.
Volume 159: Event-Related Dynamics of Brain Oscillations, by C. Neuper and W. Klimesch (Eds.) – 2006, ISBN 978-0-444-52183-5.
Volume 160: GABA and the Basal Ganglia: From Molecules to Systems, by J.M. Tepper, E.D. Abercrombie and J.P. Bolam
(Eds.) – 2007, ISBN 978-0-444-52184-2.
Volume 161: Neurotrauma: New Insights into Pathology and Treatment, by J.T. Weber and A.I.R. Maas (Eds.) – 2007,
ISBN 978-0-444-53017-2.
Volume 162: Neurobiology of Hyperthermia, by H.S. Sharma (Ed.) – 2007, ISBN 978-0-444-51926-9.
Volume 163: The Dentate Gyrus: A Comprehensive Guide to Structure, Function, and Clinical Implications, by H.E. Scharfman
(Ed.) – 2007, ISBN 978-0-444-53015-8.
Volume 164: From Action to Cognition, by C. von Hofsten and K. Rosander (Eds.) – 2007, ISBN 978-0-444-53016-5.
Volume 165: Computational Neuroscience: Theoretical Insights into Brain Function, by P. Cisek, T. Drew and J.F. Kalaska (Eds.) –
2007, ISBN 978-0-444-52823-0.
Volume 166: Tinnitus: Pathophysiology and Treatment, by B. Langguth, G. Hajak, T. Kleinjung, A. Cacace and A.R. Mller
(Eds.) – 2007, ISBN 978-0-444-53167-4.
Volume 167: Stress Hormones and Post Traumatic Stress Disorder: Basic Studies and Clinical Perspectives, by E.R. de Kloet,
M.S. Oitzl and E. Vermetten (Eds.) – 2008, ISBN 978-0-444-53140-7.
Volume 168: Models of Brain and Mind: Physical, Computational and Psychological Approaches, by R. Banerjee and
B.K. Chakrabarti (Eds.) – 2008, ISBN 978-0-444-53050-9.
Volume 169: Essence of Memory, by W.S. Sossin, J.-C. Lacaille, V.F. Castellucci and S. Belleville (Eds.) – 2008,
ISBN 978-0-444-53164-3.
Volume 170: Advances in Vasopressin and Oxytocin – From Genes to Behaviour to Disease, by I.D. Neumann and R. Landgraf
(Eds.) – 2008, ISBN 978-0-444-53201-5.
Volume 171: Using Eye Movements as an Experimental Probe of Brain Function—A Symposium in Honor of Jean Büttner-
Ennever, by Christopher Kennard and R. John Leigh (Eds.) – 2008, ISBN 978-0-444-53163-6.
Volume 172: Serotonin–Dopamine Interaction: Experimental Evidence and Therapeutic Relevance, by Giuseppe Di Giovanni,
Vincenzo Di Matteo and Ennio Esposito (Eds.) – 2008, ISBN 978-0-444-53235-0.
Volume 173: Glaucoma: An Open Window to Neurodegeneration and Neuroprotection, by Carlo Nucci, Neville N. Osborne,
Giacinto Bagetta and Luciano Cerulli (Eds.) – 2008, ISBN 978-0-444-53256-5.
Volume 174: Mind and Motion: The Bidirectional Link Between Thought and Action, by Markus Raab, Joseph G. Johnson and
Hauke R. Heekeren (Eds.) – 2009, 978-0-444-53356-2.
Volume 175: Neurotherapy: Progress in Restorative Neuroscience and Neurology — Proceedings of the 25th International
Summer School of Brain Research, held at the Royal Netherlands Academy of Arts and Sciences, Amsterdam,
The Netherlands, August 25–28, 2008, by J. Verhaagen, E.M. Hol, I. Huitinga, J. Wijnholds, A.A. Bergen, G.J. Boer
and D.F. Swaab (Eds.) –2009, ISBN 978-0-12-374511-8.
Volume 176: Attention, by Narayanan Srinivasan (Ed.) – 2009, ISBN 978-0-444-53426-2.
Volume 177: Coma Science: Clinical and Ethical Implications, by Steven Laureys, Nicholas D. Schiff and Adrian M. Owen (Eds.) –
2009, 978-0-444-53432-3.
Volume 178: Cultural Neuroscience: Cultural Influences On Brain Function, by Joan Y. Chiao (Ed.) – 2009, 978-0-444-53361-6.
Volume 179: Genetic models of schizophrenia, by Akira Sawa (Ed.) – 2009, 978-0-444-53430-9.
Volume 180: Nanoneuroscience and Nanoneuropharmacology, by Hari Shanker Sharma (Ed.) – 2009, 978-0-444-53431-6.
496 Other volumes in PROGRESS IN BRAIN RESEARCH

Volume 181: Neuroendocrinology: The Normal Neuroendocrine System, by Luciano Martini, George P. Chrousos, Fernand
Labrie, Karel Pacak and Donald W. Pfaff (Eds.) – 2010, 978-0-444-53617-4.
Volume 182: Neuroendocrinology: Pathological Situations and Diseases, by Luciano Martini, George P. Chrousos, Fernand
Labrie, Karel Pacak and Donald W. Pfaff (Eds.) – 2010, 978-0-444-53616-7.
Volume 183: Recent Advances in Parkinson's Disease: Basic Research, by Anders Björklund and M. Angela Cenci (Eds.) – 2010,
978-0-444-53614-3.
Volume 184: Recent Advances in Parkinson's Disease: Translational and Clinical Research, by Anders Björklund and M. Angela
Cenci (Eds.) – 2010, 978-0-444-53750-8.
Volume 185: Human Sleep and Cognition Part I: Basic Research, by Gerard A. Kerkhof and Hans P.A. Van Dongen (Eds.) –
2010, 978-0-444-53702-7.
Volume 186: Sex Differences in the Human Brain, their Underpinnings and Implications, by Ivanka Savic (Ed.) – 2010, 978-0-444-
53630-3.
Volume 187: Breathe, Walk and Chew: The Neural Challenge: Part I, by Jean-Pierre Gossard, Réjean Dubuc and Arlette Kolta
(Eds.) – 2010, 978-0-444-53613-6.
Volume 188: Breathe, Walk and Chew; The Neural Challenge: Part II, by Jean-Pierre Gossard, Réjean Dubuc and Arlette Kolta
(Eds.) – 2011, 978-0-444-53825-3.
Volume 189: Gene Expression to Neurobiology and Behaviour: Human Brain Development and Developmental Disorders, by
Oliver Braddick, Janette Atkinson and Giorgio M. Innocenti (Eds.) – 2011, 978-0-444-53884-0.
Volume 190: Human Sleep and Cognition Part II: Clinical and Applied Research, by Hans P.A. Van Dongen and
Gerard A. Kerkhof (Eds.) – 2011, 978-0-444-53817-8.
Volume 191: Enhancing Performance for Action and perception: Multisensory Integration, Neuroplasticity and Neuroprosthetics:
Part I, by Andrea M. Green, C. Elaine Chapman, John F. Kalaska and Franco Lepore (Eds.) – 2011, 978-0-444-
53752-2.
Volume 192: Enhancing Performance for Action and Perception: Multisensory Integration, Neuroplasticity and Neuroprosthetics:
Part II, by Andrea M. Green, C. Elaine Chapman, John F. Kalaska and Franco Lepore (Eds.) – 2011, 978-0-444-
53355-5.
Volume 193: Slow Brain Oscillations of Sleep, Resting State and Vigilance, by Eus J.W. Van Someren, Ysbrand D. Van Der Werf,
Pieter R. Roelfsema, Huibert D. Mansvelder and Fernando H. Lopes da Silva (Eds.) – 2011, 978-0-444-53839-0.
Volume 194: Brain Machine Interfaces: Implications For Science, Clinical Practice And Society, by Jens Schouenborg, Martin
Garwicz and Nils Danielsen (Eds.) – 2011, 978-0-444-53815-4.
Volume 195: Evolution of the Primate Brain: From Neuron to Behavior, by Michel A. Hofman and Dean Falk (Eds.) – 2012,
978-0-444-53860-4.
Volume 196: Optogenetics: Tools for Controlling and Monitoring Neuronal Activity, by Thomas Knöpfel and Edward S. Boyden
(Eds.) – 2012, 978-0-444-59426-6.
Volume 197: Down Syndrome: From Understanding the Neurobiology to Therapy, by Mara Dierssen and Rafael De La Torre
(Eds.) – 2012, 978-0-444-54299-1.
Volume 198: Orexin/Hypocretin System, by Anantha Shekhar (Ed.) – 2012, 978-0-444-59489-1.

You might also like